Você está na página 1de 31

Version of Record: https://www.sciencedirect.

com/science/article/pii/S0950061821007480
Manuscript_d147d3847f5d2b13041f7bcdb7b3c264
1 Thermo-mechanical treatment of sugarcane bagasse ash with very
2 high LOI: A pozzolanic paradigm
3 Amol Kamalakar Malia*, Prakash Nanthagopalanb
4 aPh.D. Research Scholar, Department of Civil Engineering, IIT Bombay, Mumbai - 400 076, India
5 bAssociate Professor, Department of Civil Engineering, IIT Bombay, Mumbai - 400 076, India

6 Abstract

7 The sugarcane bagasse ash (SCBA) with a very high loss on ignition (LOI) (> 20%), deficient
8 in silica, has remained inattentive and needs sincere attention instead of getting dumped in
9 landfills. With this motivation, the attempts were accomplished to transmute the sugarcane
10 bagasse ash with very high LOI (3 samples with diverse LOI up to 50%) as a pozzolanic
11 material through the thermo-mechanical treatment. The optimum level of re-calcination and
12 particle size was ascertained based on characteristics and pozzolanic reactivity assessed
13 through four different methods. The results of this study ensured that the re-calcination
14 temperature of 600-700oC preserves the parental amorphous state of SCBA with enhancement
15 in silica. Based on the LOI of as-such received/untreated SCBA, the residence time of re-
16 calcination needs to be decided to fulfill the chemical requirement of ASTM C618.
17 Furthermore, the grinding of re-calcined SCBA until the particle size D50 finer than 35 µm, and
18 particles below 45 µm higher than 60% are the favorable grinding levels for achieving strength
19 activity index (SAI) higher than 75%, provided amorphous silica is higher than 25% and LOI
20 less than 20%. The association of modified Chapelle and Frattini test methods with SAI
21 evident to be an affirmative way to predict the ingenuous pozzolanic reactivity of SCBA.

22 Keywords: Sugarcane bagasse ash; strength activity index; pozzolanic reactivity; grinding; re-
23 calcination; amorphous silica.

24 1.0 Introduction

25 The sugarcane bagasse has an average gross calorific value of 17.64 MJ/kg, comparable to
26 lignite coal and wood [1]. The attractive calorific value of bagasse as a carbon-neutral fuel
27 [1,2] resulted in fuel feedstock in the co-generation plant of the sugar industry to run the
28 boilers and generate electricity. Further, the excess generated energy has become an income
29 source for these sugar industries. These outlooks dragged sugar industries for the
30 implementation of cogeneration systems in India as well as other sugar-producing countries
31 [3]. However, with the implementation of co-generation plants, the challenges for the disposal
32 of resultant ash (sugarcane bagasse ash) were rolled out as well. This ash is merely getting
33 disposed of in landfills, polluting water and air, escalating a lot of burden on environmental

*
Corresponding author.

1
© 2021 published by Elsevier. This manuscript is made available under the Elsevier user license
https://www.elsevier.com/open-access/userlicense/1.0/
34 implications. India processed 414.2 million metric tonnes of sugarcane with a total of 532 mills
35 in operation during 2018-19 [4], leading to a generation of sugarcane bagasse ash (SCBA) to
36 the tune of 12.42 million tonnes per annum considering 3% ash content [4,5].
37 On the other hand, cement industries contributing to an average of 842 kg CO2 per tonne of
38 clinker in the production of ordinary Portland cement [6], and with the infrastructural
39 developments, the demand for cement is increasing day by day. This demands an
40 environmentally sustainable and economically viable solution to deal with these issues.
41 Subsequently, the incorporation of sugarcane bagasse ash (SCBA) as a pozzolanic material in
42 the cement concrete/mortar looks to be one of the appropriate solutions to cope with these
43 concerns.
44 The ample silica in the amorphous form is the prime potential of SCBA to be a pozzolanic
45 material. However, the pretreatment before utilization is vital on account of the higher loss on
46 ignition (LOI) and larger particle sizes [7–13]. Several authors reported LOI higher than the
47 limiting criteria (10%) [11–16] of the ASTM C618 for class N [17], and very high LOI to the
48 level of 59.20 [11] and 84.80 [12] were reported as well. The utilization of SCBA with such a
49 level of LOI demands thermal pretreatment in view of enhancement in silica content and
50 amorphous silica. Moreover, the strength activity index of as-such received (raw
51 state/untreated) SCBA was observed to be as low as 49% and high as 79% [16,18]. However,
52 in most cases, the minimum requirement (75%) of ASTM C618 [17] couldn’t be achieved
53 [7,10,16]. The variability in the characteristics and pozzolanic reactivity of SCBA confines the
54 treatment adopted to a specific SCBA sample and cannot be generalised.
55 The thermal conditioning followed by grinding as a treatment for as-such received SCBA was
56 addressed by numerous authors to transform SCBA as pozzolanic material [10,19,20].
57 However, a challenging area in the field of as-such received SCBA with very high LOI (>
58 20%) is almost neglected. Nevertheless, few researchers [11,12] have put forth two-stage re-
59 calcination for very high LOI SCBA samples, followed by grinding to the level of D50 as 10
60 µm. The suitability of these treatments is concerned with a definite SCBA sample and cannot
61 be a generalised protocol in terms of energy-intensive treatment.
62 In view of the axiom mentioned above, a holistic approach dictating the most generic
63 guidelines for conversion SCBA even with high LOI as a pozzolanic material need to be
64 addressed. The study aimed to broaden knowledge on the method of treatment for SCBA with
65 very high LOI. Three SCBA samples with LOI in the range of 20% to 50% were chosen for the
66 present investigation to figure out the level of thermo-mechanical treatment eventuating
67 towards the pozzolanicity. The optimum level of thermo-mechanical treatment was assured
68 based on the characteristics and pozzolanic reactivity of pretreated SCBA samples assessed
69 with four different test methods. The present research work will provide a generalised protocol
2
70 of pozzolanicity through a synergistic effect of thermo-mechanical treatment for as-such
71 received SCBA with very high LOI.

72 2.0 Materials and methodology


73 The ordinary Portland Cement (OPC-53 grade) complying with IS 269 [21], standard sand
74 complying with IS 650 [22], and quartz powder (QP) (as an inert material) were used. The
75 three as-such received samples of SCBA having very high LOI (> 20%) were selected from
76 three sugar industries with distinct collection systems from Maharashtra province of India. The
77 obtained information of these samples is tabulated in Table 1.
78 Table 1: Details of SCBA samples collection

Capacity Furnace Ash


Sample
of furnace temperature collection
No.
(Tonnes) range (oC) system

25 35 600-700 VWS
19 125 750-850 ESP
8 50 825-860 VWS
79 VWS: Ventury wet scrubber, ESP: Electrostatic precipitator,

80 The current study is contemplated in 3 phases. Initially, the characterisation of as-such received
81 SCBA was carried out through several sophisticated instrumental techniques. Afterward, as-
82 such received SCBA were re-calcined in a laboratory muffle furnace at various calcination
83 temperatures. The optimum re-calcination temperature was decided based on the characteristics
84 and pozzolanic reactivity of re-calcined SCBA samples. Subsequently, the SCBA samples re-
85 calcined at optimum temperature were ground in a laboratory ball mill at various durations to
86 evaluate the effect of grinding on the physical attributes of SCBA. The pozzolanic reactivity of
87 pretreated (re-calcined+ground) samples was determined by four methods, in view of a realistic
88 assessment of pozzolanic reactivity as well as comprehension of the level of particle size and
89 amorphous silica content in the pozzolanic reactivity. Finally, the optimum level of re-
90 calcination temperature and particle size was proposed to reach the minimum level of
91 pozzolanicity for SCBA with a very high LOI.

92 2.1 Characterization of Sugarcane Bagasse Ash

93 The physical characteristics of as-such received and pretreated SCBA viz. particle size and
94 specific surface area (SSA) were assessed through laser particle size analyzer and Brunauer–
95 Emmett–Teller (BET) (N2 adsorption) technique, respectively. The mineralogical and chemical
96 composition of SCBA was determined with the X-ray diffraction (XRD) and inductively
97 coupled plasma (ICP) (atomic emission spectrometer) technique, with a prior grinding of
98 SCBA samples in the planetary ball mill at 400 rpm for 20 minutes to get a homogenous
99 sample for valid results. The X-ray diffraction (XRD) scans were executed with a 2θ range
100 from 5 to 70°, step size - 0.013°, and scan step time - 29.05 s. The peaks were identified

3
101 (mineralogical composition) by Rietveld refinement analysis. The total amorphous
102 content/unidentified phases in SCBA samples were quantified, employing an external standard
103 approach (XRD). The phase quantification was carried out by comparison of the phase scale
104 factors of the SCBA sample to the scale factor of standard material (ZnO) measured under
105 identical diffractometer conditions. Subsequently, the amorphous silica was calculated from
106 total silica content and percentage of crystalline phases of silica in SCBA samples. Fourier-
107 transform infrared spectroscopy was used to analyse the amorphous phases of the SCBA
108 samples, as well. The ASTM C188 [23] and ASTM C114 [24] guidelines were followed in the
109 determination of specific gravity and loss on ignition (LOI) of SCBA samples, respectively.
110 The changes in the morphology of SCBA with pretreatment were captured via microscopic
111 images using scanning electron microscopy (SEM).

112 2.2 Re-calcination of as-such received SCBA

113 The presence of silica and alumina from pozzolanic material is essential for the pozzolanic
114 reaction with Ca(OH)2 (CH) from hydrated cement. However, on account of very high LOI, as
115 a result of partial/incomplete combustion of bagasse, the amount of silica proportionally lowers
116 down. Consequently, the necessity of re-calcination of SCBA arises to enhance the silica
117 content. Although the re-calcination appears to be vital for SCBA with higher LOI, the level of
118 re-calcination needs to be judged depending upon the LOI content, initial combustion
119 temperature at the sugar industry, and crystallization temperature. With these parameters as
120 reference (Table 1), the SCBA samples were re-calcined from 500 to 800oC with an increment
121 of 100oC for 1 h to maintain the inherent amorphous nature. The re-calcination was performed
122 in a laboratory muffle furnace (ventilated) with a constant ratio of the volume of sample to
123 furnace as 0.03 [25]. The SCBA sample was kept in the furnace once the furnace reached at
124 required temperature to mimic the industrial combustion process and, after completion of
125 residence time (1 h), cooled to room temperature. The respective re-calcined samples were
126 suffixed with C500 for sample re-calcined at 500oC for 1 h and so on.

127 2.3 Mechanical grinding of re-calcined SCBA

128 Apart from the thermal treatment, the reaction kinetics is commanded by the particle size and
129 specific surface area of the pozzolanic material. Consequently, the SCBA samples re-calcined
130 at optimum temperature were ground in the laboratory ball mill for different residence times to
131 reach the datum of particle size necessary for achieving the minimum level of pozzolanic
132 reactivity. Prior to grinding of re-calcined samples, the optimum grinding charge configuration
133 and charge to feed ratio were analysed by performing several trials run on SCBA samples. The
134 binary and ternary combinations of spherical steel grinding charges of diameter 12.5, 18.0, and
135 25.0 mm were assessed with the feed to charge ratio of 1:4 and 1:5 (by weight). In line with
4
136 industrial ball mills, the volume of feed + charge was kept constant between 35 to 40% of the
137 ball mill volume. The critical speed (Nc) (ball gets attached to the wall due to centrifugation)
138 was determined with eq. 1 [26], and accordingly, the optimum speed of the ball mill adjusted to
139 65-70 RPM (65% of critical speed).
42.3
= Eq.1

140 Where Dm and db are the mill diameter and the ball diameters in metres, respectively.
141 Depending upon the results of trial runs, feed to charge ratio of 1:4 and binary grinding charge
142 configuration of 12.5 and 18 mm (3:1) was finalized, which was in line with the literature [27].
143 The higher energy for breakage of bigger charge is efficient for coarser particles, whereas the
144 larger surface area of smaller charge is more effective for smaller particles [28]. The SCBA
145 samples (8, 19, and 25) were ground for 30, 60, 90, and 120 minutes to understand the effect
146 on particle size and SSA with the optimum grinding parameters, as given in Table 2. The
147 respective ground samples were suffixed with G30 for 30 minutes of grinding, and so on.
148 Table 2: Optimum grinding parameters adopted for grinding of SCBA

Proportion (by weight) SCBA: Charge Volume of Speed


ratio (by weight) charge + feed (rpm)
SCBA Charge (steel ball)
(%)
12.5 mm 18.0 mm
1 3 1 1:4 35-40 65-70

149 2.4 Pozzolanic reactivity of SCBA

150 The pozzolanic reactivity (PR) of SCBA samples was evaluated by three direct methods viz.
151 modified Chapelle test (MCT), Frattini test (FT), and thermo-gravimetric analysis (TGA) and
152 one indirect method, strength activity index (SAI) from the available test methods [29]. The
153 direct test methods like the Frattini test [29–31] and the modified Chapelle test [11,32–34]
154 predict real pozzolanic performance through the quantification of CH concentration of
155 pozzolana-cement and pozzolana-CH mixtures, respectively. The SAI test method is a
156 collective output of pozzolanic reaction as well as filler effect imparted by finer particles,
157 reflecting as comparable compressive strength. The feasibility, reliability, and simplicity of
158 these test methods (except TGA) in the field are the critical features for adopting these test
159 methods. Whereas the thermo-gravimetric analysis (TGA) [30,35,36] as a sophisticated
160 technique for precise quantification of residual CH content in a cement-pozzolana system.
161 Thus, rather than a single test method, four different test methods were adopted to figure out
162 the conclusive pozzolanic performance of pretreated SCBA.
163 The guidelines of ASTM C311 [37] were followed in the performance of the strength activity
164 index (SAI) test method. The flow (ASTM C1437 [38]) of the SCBA blended test mixture was
165 maintained equivalent to the flow of the control mortar (CM) by adjusting the amount of water
166 content. The mortar cubes (50 mm) were cured for 7, 28, and 56 days at a temperature of 23 ±
5
167 2oC and 95% RH. The modified Chapelle test was carried out as per the IS 16354 [39]. Frattini
168 test method was performed as per the BS EN 196-5 [40] recommendations on SCBA (20% by
169 weight) blended cement. The thermo-gravimetric analysis (TGA) of hydrated cement-SCBA
170 (20% by weight) pastes cured for 28 days was performed through a simultaneous thermal
171 analyzer with prior freeze-drying after completion of the curing period.

172 2.5 Analysis of hydrated SCBA blended cement paste

173 The SCBA (20% by weight) blended cement paste samples cured for 28 days were freeze-dried
174 and ground before analyzing through X-ray diffraction techniques for qualitative analysis of
175 cementitious phases.

176 3.0 Results and Discussions

177 3.1 Characteristics of SCBA

178 3.1.1 Chemical and Mineralogical Composition

179 The chemical composition, total amorphous content, and amorphous silica of as-such received
180 (ASR), and re-calcined SCBA samples are demonstrated in Table 3.
181 Table 3: Chemical composition and amorphous content of cement, QP, and SCBA samples (%)

Sample SiO2 Al2O3 Fe2O3 CaO K2O MgO Na2O P2O5 SO3 LOI Sum Total amorphous Amorphous
(S) (A) (F) (A+F+S) content/unidentified silica
phases
Cement 19.55 5.10 4.21 63.92 0.46 0.90 0.21 0.26 3.33 2.08 28.86 - -
QP 99.52 - 0.24 - - - - - - 0.25 99.76 0.00 0.00
8-ASR 61.16 1.70 2.01 1.93 1.90 1.65 0.14 0.94 0.26 28.31 64.87 54.70 15.76
8-C500 73.08 1.85 2.27 2.49 2.76 2.03 0.13 0.00 0.35 15.04 77.20 59.40 32.38
8-C600 74.09 1.95 2.70 2.75 3.09 2.21 0.15 0.01 0.31 12.73 78.74 57.90 31.99
8-C700 75.43 1.99 2.68 2.80 3.02 2.29 0.15 0.00 0.19 11.45 80.10 57.30 32.73
8-C800 77.72 2.20 3.59 2.74 2.94 2.34 0.16 0.00 0.10 8.19 83.51 60.50 38.22

19-ASR 34.76 0.53 0.87 2.71 6.44 2.81 0.22 2.07 1.36 48.24 36.16 64.20 0.00
19-C500 58.74 0.66 1.21 2.50 5.16 2.33 0.15 0.00 0.65 28.60 60.61 43.20 1.94
19-C600 61.11 0.69 1.32 2.42 4.90 2.34 0.17 0.00 0.62 26.43 63.12 44.60 5.71
19-C700 67.36 0.85 1.61 2.88 5.36 2.71 0.24 0.01 0.47 18.51 69.82 54.90 22.36
19-C800 73.33 0.92 1.69 3.18 5.45 3.01 0.33 0.00 0.48 11.61 75.94 43.90 17.23

25-ASR 53.24 0.31 0.75 2.39 2.64 2.33 0.16 1.64 1.19 35.35 54.30 55.70 9.04
25-C500 68.06 0.34 1.15 2.66 3.21 2.46 0.13 0.00 1.19 20.80 69.55 55.00 23.06
25-C600 69.85 0.35 1.34 2.59 2.93 2.47 0.13 0.00 0.96 19.38 71.53 55.30 25.15
25-C700 72.33 0.31 1.19 2.62 2.92 2.61 0.16 0.00 0.87 16.99 73.83 55.20 27.53
25-C800 75.61 0.38 1.44 3.19 3.17 2.96 0.22 0.01 0.78 12.25 77.43 48.60 24.31

182 The LOI of SCBA (ASR) samples was in the range of 28 to 48%. This was enormously greater
183 than the limiting value (10%) specified by ASTM C618 [17] for class N. The lower thermal
184 efficiency of the sugar industry furnace and the high initial moisture content and residual sugar
185 content of SCBA result in partial combustion of bagasse, despite higher combustion
186 temperature, in turn, substantial LOI [9,16,41]. As a consequence of higher LOI of SCBA
6
187 (ASR), the silica content lowered significantly, in turn, the sum of (Al2O3 (A) + Fe2O3 (F) +
188 SiO2 (S)) (AFS) decreased as well. The AFS content of SCBA (ASR) was in the range of 36-
189 64%, which was significantly lower than the acceptable limit (70%) of ASTM C618 [17]. The
190 prior re-calcination is inevitable to reach this permissible limit of AFS. As expected, the re-
191 calcination of SCBA (ASR) improved the combustion attributes of SCBA, in turn, the LOI was
192 decreased progressively with an increase in re-calcination temperature. Successively, the silica
193 and AFS content of SCBA samples was enhanced. The optimum temperature of re-calcination
194 to reach the acceptable level of AFS was augmented, conferring to the LOI content, like 500,
195 600, and 700oC for sample 8, 25 and 19, respectively.
196 The X-ray diffraction patterns of as-such received and re-calcined SCBA samples (8, 19, and
197 25) are portrayed in Fig. 1 to Fig. 3. The polymorphs of silica in the form of crystalline phases
198 of Cristoballite, Quartz [7–9,19,42], and Tridymite [10,43,44] were observed in all SCBA
199 samples. The Moganite, as another polymorph of quartz, was detected in sample 25 as well.
200 The different polymorphs of quartz as Cristoballite and Tridymite confirm the prolonged
201 combustion at high temperatures in the sugar industry [9,16,45]. However, a broad hump at an
202 angle (2θ) between 20-40o confirms the presence of amorphous phases, dictating the positive
203 sign for a pozzolanic reaction [16,19,46,47]. As per the literature, the optimum calcination
204 temperature required to maintain the amorphous state of SCBA lies in the range of 550 to
205 800oC depending upon the residence time of combustion [10,20,25]. Incidentally, the initial
206 combustion temperature of bagasse in the sugar industry was in the same range (Table 1). In
207 view of maintaining the amorphous state of SCBA, the proposed re-calcination temperature
208 range and residence time were selected. In line with the hypothesis, there was not much change
209 in the total amorphous content of SCBA samples 8 and 25 (Table 3), conferring to the literature
210 [48]. The carbon content of the SCBA is part and parcel of total amorphous content [11,12].
211 The reduction in carbon content with an increase in re-calcination temperature reflected as a
212 reduction in total amorphous content for sample 19 having exceptionally higher LOI (48%).
213 However, with re-calcination, the amorphous silica enhanced, compensating total amorphous
214 content for all SCBA samples.
215 Amorphous silica is the ruling component of pozzolanic reactivity with portlandite [8],
216 compared to total amorphous content. The presence of amorphous silica was considerably on
217 the lower side for all as-such received SCBA samples. Subsequently, the amorphous silica was
218 enhanced with an increment in re-calcination temperature up to 700oC. However,
219 crystallisation of polymorphs of silica at 800oC presented a decrement in amorphous silica for
220 samples 19 and 25. However, this phenomenon was not observed for comparatively lower LOI
221 SCBA sample 8, and the amorphous silica was almost similar at all temperatures.

7
222 The Plants consume orthosilicic acid from groundwater, which is later polymerized as
223 amorphous silica in the plant cells, and reflected in residual ash after burning [49]. The
224 decomposition of organic matters, viz. cellulose, lignin, etc., into carbon happens due to
225 calcination. Subsequently, the oxidation of carbon takes place with the rise in the temperature
226 [50]. The resultant ash content increases with re-calcination. As the re-calcination temperature
227 is below the crystallization temperature, the sudden drop in the temperature to room
228 temperature converts silica into an amorphous state, in turn, escalation in amorphous silica.
229 The optimum level of re-calcination comes out as 500, 700, and 700oC for samples 8, 19, and
230 25, respectively, concerning the amorphous silica. The amorphous silica at the optimum level
231 of re-calcination temperature ranges from 22 to 32%, which was higher than the minimum
232 requirement of reactive silica (20%) as per IS 3812 (part-I) [51], exhibiting an exceptional
233 pozzolanic potential for SCBA.
234 The re-calcination temperature in the range of 600-700oC is suitable to enhance silica and
235 amorphous silica of the high LOI SCBA samples. Nevertheless, based on the furnace
236 characteristics and LOI in the raw state, the residence time can be decided. Indeed, the re-
237 calcination in the industrial furnace with a forced-air system would be more appropriate for
238 such high LOI SCBA samples.
Cr: Cristoballite Cr: Cristoballite
Q: Quartz Q: Quartz
T: Tridymite Cr
Cr T: Tridymite

Q QT
T T
T Cr
Cr T
Intensity

Intensity

8-C800 19-C800
8-C700 19-C700
8-C600 19-C600
8-C500 19-C500
Cr
8-ASR 19-ASR

0 10 20 30 40 50 60 70 80 0 10 20 30 40 50 60 70 80
2θ (Degree) 2θ (Degree)
Fig. 1: X-ray diffraction patterns of SCBA sample 8 Fig. 2: X-ray diffraction patterns of SCBA sample 19

Cr: Cristoballite
Q: Quartz 8-C800
Cr T: Tridymite
8-C700
M: Mognite
8-C600
QCr
Transmittance

8-C500
Cr
8-ASR
Cr 3733.51 2358.52 2300.66
799.35
Intensity

1545.67
25-C800

25-C700
T
M 25-C600

25-C500 465.72

25-ASR
1104.05

0 10 20 30 40 50 60 70 80 4500 4000 3500 3000 2500 2000 1500 1000 500


2θ (Degree) -1
Wavenumber (cm )
Fig. 3: X-ray diffraction patterns of SCBA sample 25 Fig. 4: FTIR spectra of SCBA sample 8

8
239 3.1.2 Fourier-transform infrared spectroscopy (FTIR) of SCBA

240 The FTIR spectra of SCBA samples re-calcined at different temperatures are showcased in Fig.
241 4 to Fig. 6. The SCBA samples presented quite comparable peak positions regardless of
242 differences in characteristics, but the transmittance was notable. The broad bands located near
243 1104 cm-1 and bands with smaller intensity near 799 and 465 cm-1 are the vibrations of the Si–
244 O bonds of the quartz [52–54]. The band located at 1104 cm-1 got broaden as per the presence
245 of amorphous silica phases in these SCBA samples and showcased differences in transmittance
246 intensity as well. The enhancement in amorphous silica as a result of re-calcination is evident
247 in FTIR, as well. The wavelength band of 1104 cm-1 shows that quartz in the form of SiO4 is
248 asymmetric vibration, which is the tetrahedral shape [53]. The band stretching near 3733 cm-1
249 indicates the O–H molecule stretching [54] as a result of moisture on account of porous nature.
250 The presence of carbon trace is affirmed by the bands at 2853 and 2922 cm-1 [52]. Moreover,
251 the occurrence of minor bands around 2300 and 2356 cm-1 could be on account of HCO3- ions
252 [55]. Further, the band at 1545 cm-1 representing C–O vibrations was quite cognizable in
253 samples 19 and 25. However, with re-calcination, the bands at these positions were diminished
254 considerably due to efficient combustion.

19-C800
25-C800
19-C700
25-C700
19-C600
25-C600
19-C500
651.82
Transmittance

Transmittance

651.82 25-C500
19-ASR
2853.17 2300.66 25-ASR 2300.66
3733.51 3733.51 2922.59 678.82
2922.59 2358.52 1545.67 797.42 2853.17
1545.67 799.35
2356.59

465.72

465.72
1104.05

1104.05
4500 4000 3500 3000 2500 2000 1500 1000 500 4500 4000 3500 3000 2500 2000 1500 1000 500
-1 -1
Wavenumber (cm ) Wavenumber (cm )
Fig. 5: FTIR spectra of SCBA sample 19 Fig. 6: FTIR spectra of SCBA sample 25

255 3.1.3 Particles Morphology of SCBA

256 For the sake of understanding the effect of re-calcination and grinding on the morphology of
257 SCBA samples, scanning electron microscopy (SEM) images of two SCBA samples (distinct
258 LOI) are depicted in Fig. 7 and Fig. 8. The LOI of SCBA indirectly dictates the efficiency of
259 the combustion of bagasse in the sugar industry. Sugarcane has a fibrous tubular structure to
260 store sugar juice. However, this cellular porous tubular structure is retained even after
261 combustion [44,45,48,56], and the porosity of particles depends upon the efficiency of
262 combustion.

9
A B C D

Fig. 7: SEM micrographs of SCBA 8 (A. as-such received, B. re-calcined at the optimum temperature (600oC), C. & D. ground for 90 minutes)

A B D
C

Fig. 8: SEM micrographs of SCBA 19 (A. as-such received, B. re-calcined at the optimum temperature (700oC), C. & D. ground for 90 minutes)

263 According to LOI, the differences in the porous structure of samples were observed. The
264 disintegration of particle surfaces due to comparatively efficient combustion at the sugar
265 industry was observed, but the internal porosity was still maintained (Fig. 7A) at lower LOI
266 SCBA. In contrast, the parental morphology of bagasse was almost maintained (Fig. 8A) at
267 higher LOI. This highly porous nature of SCBA is contributing to a higher specific surface
268 area. Subsequently, the fragmentation of particles was quite evident in both samples as an
269 effect of re-calcination, which was more pronounced for the higher LOI sample (Fig. 7B and
270 Fig. 8B). However, the internal porous structure was relatively intact to a certain level even
271 after re-calcination. In general, the SCBA particle does not have any definite shape. The
272 fragmentation and destruction of parental porous, cellular structure with grinding are clearly
273 visible from the SEM micrographs (Fig. 7C and Fig. 8C). The particle size was considerably
274 reduced as an effect of grinding, but the irregular shape of particles was observed even after
275 grinding. The intrinsic details got revealed at higher magnifications (Fig. 7D and Fig. 8D).
276 With the formation of new finer particles, the consolidation of the internal porous structure, to
277 a certain extent, was observed as well. The fibrous porous cell-like structure, distinctly visible
278 in as-such received samples, was not found to that level in ground samples. It seems that the
279 impact energy of the grinding charge partially compressed the porous tubular structure.

280 3.1.4 Physical Properties

281 The particle size distribution of as-such received, re-calcined, and ground SCBA samples (8,
282 19, and 25), cement, and QP is illustrated in Fig. 9 to Fig. 11. The combustion characteristics
283 of bagasse are governing the particle size of the SCBA. No doubt, the particle size of SCBA
284 (ASR) is much larger compared to cement. In general, the role of efficient combustion is

10
285 evident from the particle size. Sample 19 with the highest LOI presents a considerably larger
286 particle size, approximately equal to fine aggregate (sand).

100 100
Cement Cement
90 QP 90 QP
8-ASR 19-ASR
80 8-C500 80 19-C500

Cumulative Volume (%)


Cumulative Volume (%)

8-C600 19-C600
70 70
8-C700 19-C700
60 8-C800 60 19-C800
8-C600-G30 19-C700-G30
50 8-C600-G60 50 19-C700-G60
8-C600-G90 40 19-C700-G90
40
8-C600-G120 19-C700-G120
30 30

20 20

10 10

0 0

-10 -10
0.1 1 10 100 1000 0.1 1 10 100 1000
Particle Size (µm) Particle Size (µm)
Fig. 9: Particle size distribution of SCBA Sample No. 8, cement and Fig. 10: Particle size distribution of SCBA Sample No. 19, cement and QP
QP

100 300 Specific Surface Area (BET) (m²/g) 2.4


Cement
90 QP 275 Specific Gravity 2.2
Specific Surface Area (BET) (m²/g)

25-ASR
80 250 2.0
25-C500
R2 = 0.79
Cumulative Volume (%)

25-C600 225 1.8


70
25-C700

Specific Gravity
1.6
60 25-C800 200
25-C700-G30 1.4
50 175
25-C700-G60
1.2
40 25-C700-G90 150
25-C700-G120 R2 = 0.85 1.0
30 125
0.8
20 100
0.6
10 75 0.4

0 50 0.2

-10 25 0.0
0.1 1 10 100 1000 0 10 20 30 40 50
Particle Size (µm) LOI (%)
Fig. 11: Particle size distribution of SCBA Sample No. 25, cement Fig. 12: Correlation of specific surface area and specific gravity with LOI
and QP

287 The physical properties of cement, QP, and SCBA samples are demonstrated in Table 4. The
288 ASTM C618 [17] dictates fineness criteria in percentage passing through 45 μm > 66.0%.
289 However, none of the SCBA (ASR) clears this criterion, in line with the literature [15,57]. The
290 particle size D50 of as-such received SCBA samples is in the range of 74 to 290 μm,
291 corroborating with the literature [56,58]. The reduction in particle size with the re-calcination
292 of SCBA was quite cognizable for coarser particles (D90); however, it was not that efficient for
293 smaller particles (D50). Nevertheless, even after re-calcination, the coarser particles compared
294 to cement are apparent.
295 The as-such received SCBA samples present significantly larger SSA contributed by the
296 internal porous structure, ranging from 109 to 275 m2/g, corroborating with literature (173 -
297 288 m2/g) [41]. A combination of meso and micropores from unburned carbon confers the
298 tremendously higher specific surface area [41]. As discussed earlier, efficient combustion plays
299 a vital role in the breakdown of the porous, cellular structure of SCBA, ensuing reduction in

11
300 SSA with an increase in calcination temperature [10,58]. Subsequently, the correlation between
301 SSA and LOI presented in Fig. 12 is attesting to this fact. The reduction of porous nature is an
302 appreciative entity, which will curtail the water requirement of the system.
303 Table 4: Physical properties of cement, QP, and SCBA samples

Sample No. D10 D50 D90 Particles below SSA Mean Total pore Specific
(μm) (μm) (μm) 45 μm (%) (BET) Pore Dia. volume (cm3/g) gravity
(m²/g) (nm)
Cement 6.84 16.60 65.71 83.00 0.567 36.10 0.0051 3.140
QP 2.02 17.02 54.97 84.21 1.180 21.06 0.0062 2.797
8-ASR 22.36 74.47 312.60 29.09 124.15 3.50 0.1085 1.779
8-C500 17.61 61.74 241.05 36.75 73.81 5.06 0.0934 2.057
8-C600 17.03 57.64 202.03 39.08 61.85 5.84 0.0902 2.076
8-C700 17.24 57.07 184.56 39.23 70.56 5.07 0.0894 2.234
8-C800 18.75 57.84 174.75 37.97 59.03 4.82 0.0670 2.312
8-C600-G30 5.82 27.76 71.51 72.74 19.61 13.10 0.0642 2.155
8-C600-G60 4.47 22.65 59.06 80.70 23.70 9.34 0.0553 2.351
8-C600-G90 3.65 18.95 50.52 86.46 25.83 8.81 0.0569 2.429
8-C600-G120 3.23 16.87 45.84 89.45 24.45 9.27 0.0567 2.449
19-ASR 33.28 290.17 1788.34 13.34 275.79 2.21 0.1523 0.598
19-C500 29.08 135.70 1006.34 17.48 125.38 2.85 0.0894 1.518
19-C600 26.52 123.43 889.81 19.54 131.30 2.84 0.0931 1.520
19-C700 27.43 112.87 736.46 19.61 123.31 2.69 0.0829 1.598
19-C800 29.22 104.13 403.14 18.75 81.37 3.26 0.0637 1.884
19-C700-G30 6.98 33.20 104.32 61.90 90.37 3.10 0.0699 2.243
19-C700-G60 5.32 25.51 78.21 72.39 89.17 2.99 0.0667 2.318
19-C700-G90 4.47 21.33 64.76 79.00 88.66 2.97 0.0658 2.350
19-C700-G120 3.86 18.67 57.07 83.26 90.72 2.98 0.0675 2.409
25-ASR 19.78 74.12 363.17 31.09 109.39 3.80 0.1039 1.530
25-C500 14.58 53.78 201.83 42.71 76.99 5.07 0.0976 1.860
25-C600 14.66 52.38 182.88 43.59 81.61 4.41 0.0900 1.886
25-C700 15.08 53.57 198.18 42.68 79.08 3.86 0.0764 1.956
25-C800 15.67 52.17 166.29 43.28 52.68 7.73 0.0638 1.959
25-C700-G30 6.73 28.76 84.56 69.36 56.32 4.83 0.0680 2.349
25-C700-G60 5.02 22.37 63.80 79.38 59.07 5.91 0.0873 2.399
25-C700-G90 4.15 18.84 54.38 84.80 60.50 5.87 0.0888 2.450
25-C700-G120 3.71 16.85 48.99 87.85 65.37 4.30 0.0703 2.464

304 Despite diversified feed particle size, there were almost similar end products in terms of
305 particle size as an effect of grinding. A shift towards the particle size distribution of cement
306 with the increase in the grinding duration for SCBA is apparent from the particle size
307 distribution curves. The particle size D50 was progressively decreased with an increase in
308 grinding duration (Fig. 13). However, the rate of fragmentation of particles seems to be
309 reached at a saturation level at 90 minutes of grinding, as the change in the particle size (D50) at
310 90 and 120 minutes was insignificant. This could be attributed to the difficulty of the sub-
311 microns particle being broken by grinding media [2]. The SCBA samples (8 and 25) re-
312 calcined at optimum temperature are almost reaching the level of fineness of cement after 90
313 minutes of grinding, whereas sample 19 took 120 minutes. Similarly, it seems that 30 minutes
314 of grinding is just sufficient for reaching the fineness criterion of ASTM C618 [17] for samples

12
315 8 and 25, whereas sample 19 required 60 minutes. The modest difference in grinding time was
316 on account of feed particle size.
317 It is a well-accepted fact that the surface area will increase with the formation of new surfaces
318 and declination in particle size. On the contrary, it doesn’t seem to be applicable for porous
319 material like SCBA. There was a considerable drop in SSA of all SCBA samples re-calcined at
320 optimum temperature with grinding for 30 minutes (Fig. 13). However, at the subsequent
321 grinding time, there was a modest enhancement in the SSA for samples 8 and 25. The modest
322 rise in SSA could be the contribution from the new finers particles formed due to grinding.
323 Nevertheless, this modest change was not observed for sample 19, maybe on account of
324 retention of a certain level of porous nature even after thermo-mechanical pretreatment. The
325 surface area of SCBA samples 8, 19, and 25 after grinding for 120 minutes was observed as
326 24.45, 90.72, and 65.37 m2/g, respectively, significantly lower than the as-such received
327 samples. The breakdown of soft silica walls due to mechanical action resulted in the collapse
328 of porous structure and filled with submicron particles, which resulted in a decrease in SSA
329 [59,60]. The impact force of grinding media compressed the porous tubular channels of SCBA.
330 From this observation, it is clear that the breakdown of porous, cellular structure material may
331 not always result in the creation of new surfaces (as in the case of most solid materials) [59],
332 but a reduction of surface area as well. These porous channels formerly reachable for N2
333 adsorption might be flattened due to grinding media [61]. Nevertheless, the grinding is
334 considerably effective in reducing particle size; however, diminution in SSA must be
335 simultaneously observed.
D50 300
300 8 19 25
275
Specific surface area (BET) (m2/g)

SSA (BET)
275 8 19 25 250
225
100 200
D50 (µm)

175
75
150
125
50
100
25 75
50
0
25
0
ASR 0 (Opt. Cal.) 30 60 90 120

336 Duration of Grinding (minutes)


337 Fig. 13: Effect of grinding on specific surface area and particle size (D50)

338 The as-such received SCBA samples exhibited significantly lower specific gravity in the range
339 of 0.6 to 1.8. The existence of carbon, partially burnt, and un-burnt material contributing to
340 LOI decides the specific gravity (SG) of SCBA [62]. With re-calcination, the LOI was
341 significantly reduced, and as a result, the SG of SCBA was enhanced (1.8 to 2.3). The
342 correlation between LOI and SG, as illustrated in Fig. 12 is attesting to this fact. Consequently,
13
343 there was a further improvement in the SG with a decrease in particle size of SCBA. The
344 concept discussed for a reduction in SSA could be the probable aspect behind enhancement in
345 the SG. Eventually, the specific gravity was still much lower than the cement.

346 3.2 Strength activity index

347 A test mixture composed of SCBA (ASR) demands exceptionally higher water to maintain the
348 flow of mortar. The water requirement was varied from 126 to 156%, approximately 26 to 56%
349 higher than the control mixture. This water requirement was significantly reduced with a surge
350 in re-calcination temperature for all SCBA samples. Subsequently, with a reduction in particle
351 size as an output of grinding, there was a further reduction in water requirement. The SCBA
352 samples 8 and 25, and 19 are reaching the limiting threshold (115%) values of water
353 requirement at 30 minutes and 60 minutes of grinding, respectively.
354 The role of a few characteristics of SCBA in water requirements dictates the rationale behind
355 this scenario. The correlation of LOI, SSA, and SG with water requirement gave insights into
356 this behaviour of SCBA (Fig. 14 and Fig. 15). The SCBA has a cognizably greater SSA on
357 account of its porous tubular channel morphology and, at the same time, significantly lower
358 specific gravity. The high specific surface area on account of porous structure escalated water
359 requirement [10,16,63]. Whereas the lower specific gravity of SCBA intensifies the volume of
360 the system when mass replacement is followed, in turn, more water will be demanded as per
361 the specific gravity [62].
Specific Surface Area (BET) (m²/g) 2.6
160 300
Specific Gravity 2.4
Water Requirement % of Control (%)

Specific Surface Area (BET) (m²/g)

2
R = 0.91
2.2
150 250 R2 = 0.83
2.0
1.8

Specific Gravity
200
140
1.6

150 1.4
130 1.2

100 1.0
120 R2 = 0.91
0.8
50 0.6
110 0.4
0 0.2
100 0.0
0 10 20 30 40 50 100 110 120 130 140 150 160
LOI (%) Water Requirement % of Control (%)
Fig. 14: Correlation between LOI and water requirement of SCBA Fig. 15: Correlation of SG and SSA with water requirement of SCBA
blended mortar blended mortar (re-calcined and ground)

362 Further, the inter-particle resistance offered by the disintegrated particles and irregular shapes
363 reduces the flow of mortar, leading to more water requirements [63]. With the re-calcination of
364 SCBA (ASR), carbon content decreased considerably and the collapse of porous morphology.
365 Consequently, the water requirement of the system was decreased with a reduction in SSA as a
366 result of the destruction of porous morphology [10,63]. As discussed earlier (section 3.1),
367 previously accessible pores are filled with sub-microns particles and compressed to a certain
14
368 extent (decrement in SSA); as a repercussion, the water requirement is downscaled with
369 grinding. Moreover, the increase in specific gravity has lowered the volume of binder content
370 compared to volume with as-such received SCBA. Further, the conversion to a definite shape
371 from disintegrated nature reduced the frictional resistance. As an overall effect, the water
372 requirement of ground SCBA was decreased with a reduction in particle size.
373 The strength activity index and compressive strength of re-calcined SCBA samples at various
374 curing ages are illustrated in Fig. 16 to Fig. 18. The SAI of SCBA (ASR) samples is much
375 lower than the threshold line (75%) set by IS 3812 (Part-I) [51] and ASTM C618 [17]. The
376 lower SAI concerns to meager silica content, higher water/binder ratio on account of high SSA,
377 larger particle size imparting loosening effect, dilution effect (a result of 20% lower cement
378 content in the test mixture), and lower amorphous silica content [2,16,31]. To overcome all
379 these hindrances, the SCBA (ASR) was first re-calcined at various temperatures.
Avg. Compressive Strength (CS) (MPa)

Avg. Compressive Strength (CS) (MPa)


90 CS 7D CS 28D CS 56D 210 90 CS 7D CS 28D CS 56D 210
SAI 7D SAI 28D SAI 56D 195 SAI 7D SAI 28D SAI 56D 195
75 75
180 180
60 165 60 165
150 150
45 45
135 135
30 120 30 120
100
100

100
100

100
100

15 105 15 105
78.57

77.88
75.97

76.44

75.49
75.49

90 90
73.43
73.43

73.25

71.27

70.15
70.15

70.99
68.02

67.68
0 0
63.68

63.05
61.02
64.3
59.32

59.05

75

SAI (%)
59.39

75
SAI (%)
53.37

58.77

57.09

55.86
55.25
54.62

53.76
56.7

52.33
52.77
51.56
-15 60 -15 60
30.87
34.4

45 45
27.43

-30 -30
30 30
-45 -45
15 15
-60 0 -60 0
CM QP 8 ASR 8-C500 8-C600 8-C700 8-C800 CM QP 19 ASR 19-C500 19-C600 19-C700 19-C800

Sample Name Sample Name


Fig. 16: Strength activity index and avg. comp. str. of sample 8 re- Fig. 17: Strength activity index and avg. comp. str. of sample 19 re-calcined
calcined at different temperatures at 7, 28 and 56 days of curing period at different temperatures at 7, 28 and 56 days of curing period
Avg. Compressive Strength (CS) (MPa)

90 CS 7D CS 28D CS 56D 210 100 Silica


SAI 7D SAI 28D SAI 56D 195 Amorphous Silica
75 90
180 LOI
60 165 80 R2 = 0.94
150 70
45
135
30 60
120
%
100
100
100

15 105 50
71.18
75.49

90
73.05
73.43

73.08
70.15

70.49
70.29
70.04

0 40
66.77
66.27

R2 = 0.78
61.16

61.02

75
SAI (%)
58.02
57.53
55.89

54.7

-15 60 30
46.1

-30 45
20
30
-45 R2 = 0.89
15 10
-60 0 0
CM QP 25 ASR 25-C500 25-C600 25-C700 25-C800 30 40 50 60 70 80 90
Sample Name SAI 56D (%)
Fig. 18: Strength activity index and avg. comp. str. of sample 25 re- Fig. 19: Correlation of Silica content, amorphous silica, and LOI with SAI
calcined at different temperatures at 7, 28 and 56 days of curing period at 56 days

380 The compressive strength of re-calcined SCBA samples was augmented with a rise in re-
381 calcination temperature up to an optimum temperature of re-calcination and beyond which
382 there was a trivial drop in the compressive strength. The optimum level of re-calcination

15
383 temperature was governed by the original state of silica content, amorphous silica, and LOI.
384 There was a progressive increment in the SAI up to 700oC for the SCBA samples 19 and 25.
385 This progress was in line with total silica content and amorphous silica. However, beyond
386 700oC, there was a marginal downtrend in SAI, on account of the initiation of crystallization of
387 silica phases and confirmed through the reduction in amorphous silica at 800oC.
388 The SAI of sample 8 at 28 days was improved up to 600oC, and subsequently, modest
389 dwindling observed. However, the SAI at 56 days was almost consistent at the re-calcination
390 temperature as a result of the appreciable amorphous silica, which was virtually similar up to
391 700oC. The correlation of silica content, LOI, and amorphous silica with SAI at 56 days is
392 presented in Fig. 19. The SAI 56D was increased with amorphous silica as well as total silica
393 content. It can be inferred that the presence of amorphous silica, as well as total silica, are key
394 factors governing the compressive strength of test mortar. The LOI is indirectly proportional to
395 silica content, in turn, commands the SAI. Further, the porosity of SCBA was decreased with a
396 reduction in LOI, lowering the water requirement, and finally reflected as strength
397 improvement.
CS 7D CS 28D CS 56D 210 CS 7D CS 28D CS 56D 210
90 90
Avg. Compressive Strength (CS) (MPa)

Avg. Compressive Strength (CS) (MPa)

SAI 7D SAI 28D SAI 56D 195 SAI 7D SAI 28D SAI 56D 195
75 75
180 180
60 165 60 165
150 150
45 45
135 135
30 120 30 120
100

100
100

100
100

100
90.54
91.92

15 105 15 105
81.46
82.29

82.25
80.78

81.52
79.07

79.01
77.25
81.7

78.31
77.09
76.56
76.44
75.49

75.49

75.79
75.07

73.99
90 90
73.43

73.43
73.25

76.5
71.68
70.15

70.15

67.44
68.56

69.18
0 0 63.65
63.68

SAI (%)

SAI (%)
63.05

63.53
59.39

75 75
58.77
53.37

57.09
53.76

-15 60 -15 60
30.87
34.4

45 45
27.43

-30 -30
30 30
-45 -45
15 15
-60 0 -60 0
R 0 30 60 90 120 SR 9-C700 30 60 90 120
CM QP 8 AS 8-C60 8-C600-G 8-C600-G 8-C600-G8-C600-G CM QP 19 A 1 700-G 9-C700-G 9-C700-G9-C700-G
19-C 1 1 1
Sample Name Sample Name
Fig. 20: Strength activity index and avg. comp. str. of re-calcined sample Fig. 21: Strength activity index and avg. comp. str. of re-calcined sample 19
8 ground for various durations at 7, 28, and 56 days of curing period ground for various durations at 7, 28, and 56 days of curing period

90 CS 7D CS 28D CS 56D 210 CM QP 8-C600-G120


90
Avg. Compressive Strength (CS) (MPa)

SAI 7D SAI 28D SAI 56D 195 19-C700-G120 25-C700-G120


75
180
80
Avg. Comp. Strength (MPa)

60 165
150
45 70
135 Dilution Effect
30 120
60
100
100
100

Pozzolanic Effect
15 105
83.16
81.87
79.71

80.18
78.84
78.52
77.44
75.49

75.68

75.42
74.49

90
73.05
73.43

73.08

76.5
70.15

70.45

0 50
SAI (%)

75
58.02
57.53
55.89

-15 60
46.1

40
-30 45
30 30
-45
15
-60 0
0 20
0 0 0 -G12
QP 25 AS
R 700 0-G3 0-G6 0-G9
CM 25-C 25-C70 25-C70 25-C70 25-C700 7 28 56
Sample Name Curing Period (Days)
Fig. 22: Strength activity index and avg. comp. str. of re-calcined sample Fig. 23: Differentiation of Pozzolanic effect and dilution effect
25 ground for various durations at 7, 28 and 56 days of curing period

16
398 Despite re-calcination, none of the SCBA samples reach the limiting threshold line of SAI on
399 account of retained larger and porous particles. This inculcates the essentiality of mechanical
400 grinding of SCBA re-calcined at optimum temperature. Based on the characteristics (AFS,
401 amorphous silica, and LOI), SAI at 28 days, and energy perspective, the re-calcination
402 temperature 700oC for samples 19 and 25, whereas 600oC for sample 8 was selected as
403 optimum temperature. The SCBA samples re-calcined at optimum temperature were ground
404 for various durations to convert SCBA as a pozzolanic material. The effect of reduction in
405 particle size of SCBA samples re-calcined at the optimum temperature on SAI and
406 compressive strength at 7, 28, and 56 days of the curing period is depicted in Fig. 20 to Fig. 22.
407 With the reduction in particle size as an effect of grinding, the compressive strength of the
408 ground SCBA mortar was improved. As well as with the curing age, there was a continuous
409 increment of the compressive strength; nevertheless, the rate of increase was governed by total
410 silica content and amorphous silica from SCBA.
411 The higher SSA and considerable amorphous silica remarkably enhance the compressive
412 strength up to 28 days of the curing period, beyond which the rate of increment in strength was
413 considerably lowered. The dissolution of amorphous silica seems to be more prominent in the
414 early period due to higher surface area, and according to the availability of calcium hydroxide,
415 the pozzolanic reaction could have taken place. However, at a later period, the available
416 amorphous silica might have already converted into a secondary cementitious gel; thus, the
417 difference in compressive strength at 28 days and 56 days is meager. Accordingly, SAI at 56
418 days for all SCBA samples is insignificant compared to SAI at 28 days.
419 The SAI at 28 days was overstepped the threshold line of 75% with grinding for just 30
420 minutes for samples 8 and 25, and on the contrary, sample 19 taken 90 minutes of grinding.
421 The dominance of amorphous silica and the level of particle size are imparting differences in
422 optimum grinding time. However, the particle size required to reach the minimum threshold
423 (75%) of SAI was slightly varied in proportion with amorphous silica.
424 The transition of ground SCBA as a pozzolanic material results from improvement in the
425 characteristics and behaviour of re-calcined SCBA with grinding. Nevertheless, the differences
426 in SAI were on account of the diversified properties of SCBA samples. The rate of dissolution
427 of amorphous silica is governed by particle size and SSA [2,10], and with grinding, it was
428 improved to promote the pozzolanic reaction among the calcium hydroxide and reactive silica
429 to form secondary calcium silicate hydrate (CSH) gel. Subsequently, the filler effect
430 eventuated with finer particles facilitates pore refinement, which could be another facet of
431 improvements in SAI [64]. Eventually, the considerable reduction in water requirement on
432 account of consolidation porous nature, in turn, reduction water/binder ratio, is another aspect
433 for enhancement in the SAI. Nevertheless, the presence of considerable LOI at this level of re-
17
434 calcination seems to a stumbling block in reaching equivalent strength as that of control
435 mortar.
436 By comparing the SAI of control mortar and QP (inert material) blended test mortar, it can be
437 inferred that the dilution effect has reduced SAI by more than 25%. However, the dilution
438 effect may be much more than 25%, as the particles of QP are 50% finer than the cement. The
439 SAI of QP probably escalated due to the filler effect promoted by finer QP particles.
440 In view of differentiating the collective results of the pozzolanic reaction and filler effect, the
441 compressive strength of SCBA mortar is related to the compressive strength of cement and QP
442 (Fig. 23). The particle size D50 and particle size distribution of SCBA samples ground for 120
443 minutes are comparable with QP. The thermo-mechanical transmutation of SCBA samples as a
444 pozzolanic material is distinguishable from Fig. 23. Although the level of particle size of
445 SCBA samples was almost similar at this grinding period, the change on account of the
446 pozzolanic reaction was quite different. The level of amorphous silica and LOI could be
447 dominating the pozzolanic reaction.
448 The augmentation in SAI with a reduction in particle size D50 and an increase in percentage
449 passing through 45 µm is technically a well-known fact observed for SCBA samples as well
450 (Fig. 24). Further, the reduction in SSA lowers the water demand of the system, which reduces
451 the water/binder ratio, in turn, improvement in the SAI. Even though there was a reduction in
452 the SSA, it was significantly higher than the cement promoting the pozzolanic reaction.
D50 (µm) 100 100
300
Particles below 45 μm (%) 90
90
250 R2 = 0.89
Particles below 45 μm (%)

80
R2 = 0.80
80
70
200 R2 = 0.81
SAI 28D (%)
D50 (µm)

60 70
150
50
60
40
100
50
30
50
20 40

10
0 30
116.42
0
30 40 50 60 70 80 90 100 100 110 120 130 140 150 160

SAI 28D (%) Water Requirement % of Control (%)


Fig. 24: Correlation between SAI at 28 days with D50 and Particles below Fig. 25: Correlation between the water requirement % of control mixture
45 µm and SAI at 28 days of curing period

453 The correlation between SAI at 28 days and water required to maintain the standard flow of
454 test mortar is depicted in Fig. 25. According to this relationship, the maximum allowable water
455 requirement percent of the control mixture to achieve 75% of SAI comes out as 116%,
456 corroborating the limit (115%) set by ASTM C618 [17] for Class N pozzolana.
457 The thermal pretreatment in terms of re-calcination between 600 to 700oC seems to be
458 sufficient for fulfilling the chemical requirements with enhancement in amorphous silica.

18
459 Further, the residence time, which brings down LOI below 20% as an upper level of LOI,
460 appears to be conclusive from an energy prospect, provided AFS content is as per the
461 requirement of ASTM C618 [17] and amorphous silica higher than 25%. Subsequently, the
462 results assert that the particle size D50 lesser than 35 µm and particles below 45 µm higher than
463 60% look to be conclusive in the transmutation of re-calcined SCBA as a pozzolanic material,
464 which corresponds to 30 minutes of grinding. However, the SCBA with lower amorphous
465 silica (< 25%) demands ultra-fine particles to reach 75% of SAI. The attainment of SAI to the
466 level of control mixture can be achieved through the combined effect of pozzolanic reaction
467 and pore refinement. This demands re-calcination to the level of LOI less than 10% as per the
468 suggestion of ASTM C618 [17] and ultra-fine grinding.

469 3.3 Modified Chapelle Test

470 The results modified Chapelle test (MCT) for re-calcined SCBA and ground SCBA test
471 samples are portrayed in Fig. 26 and Fig. 27, respectively. The IS 16534 [21] set a line of
472 differentiation for a pozzolanic and non-pozzolanic material based on calcium hydroxide (CH)
473 consumption to the tune of 800 mg Ca(OH)2/g of pozzolana (for Metakaolin). However, it
474 doesn't seem to be promising for porous SCBA samples, and the as-such received samples 8
475 and 25 are quickly reaching the differentiating line of pozzolanicity on account of the
476 availability of amorphous silica as well as substantial SSA. Nevertheless, the inertness of QP
477 was validated with a low consumption CH (176 mg/g).
1800 1800
ASR C500 C600 C700 C800 ASR Opt. Cal. G30 G60 G90 G120
1650 1650

1437
1395
1344
1396
1500 1500

1301
1282
1301

1259
1259

1227
1259

1229
1208

1243

1204

1192

1350
1171
1204

1204

1350
Ca(OH)2 Fixed (mg/g)
Ca(OH)2 Fixed (mg/g)

1140
1095

1072
1093

1072

1200
1040

1200
983

1050
885

1050
892

885

888
888

900 900
800 800
750 750
464
464

600 600

450 450

300 300

150 150

0 0
8 19 25 8 19 25
Sample No. Sample No.

Fig. 26: Influence of re-calcination of SCBA samples on CH fixation Fig. 27: Influence of grinding of re-calcined SCBA samples on CH fixation in
in Modified Chapelle test Modified Chapelle test

478 The amount of CH fixed by SCBA samples was enhanced up to the optimum re-calcination
479 temperature and beyond which there was a minor decrement. On the track of SAI at 28 days
480 results, the CH consumption was surged up to 600oC for sample 8 and up to 700oC for samples
481 19 and 25, confirming the optimum re-calcination temperature. The increase in amorphous
482 silica with re-calcination escalated CH consumption approximately by 50% compared to as-

19
483 such received samples. The recrystallization of silica beyond optimum temperature and
484 dwindling in SSA are the possible root causes behind the reduction in CH consumption.
485 The effect of grinding in terms of reduction in particle size on CH consumption in MCT is
486 depicted in Fig. 27. The modest improvement in CH consumption with grinding for 30 minutes
487 is evident. However, the subsequent grinding up to 120 minutes does not show a notable
488 difference in the amount of CH fixation. Although the particle size of re-calcined SCBA was
489 dropped with grinding, the SSA decreased as well. In short, the finer particles are
490 compensating the effect of the decreased SSA. Nevertheless, there was a considerable
491 difference in CH consumption in these ground samples of SCBA, even though the level of
492 particle size was quite comparable. These differences are in line with the presence of
493 amorphous silica and surface area [11,12,34]. Sample 19 having comparatively higher SSA,
494 but amorphous silica was the lowest; hence the CH consumption was relatively lesser than the
495 samples 8 and 25. The necessity of finer particles in view of enhancing the pozzolanic
496 reactivity is evident through the MCT results as well. Further, the surface area in the range of
497 25 to 50 m2/g seems reasonable for a pozzolanic reaction. Nonetheless, the amount of
498 amorphous silica governing the pozzolanic reaction needs to be observed as well before the
499 finalization of the level of particle size. The significantly higher consumption of CH (> 1200
500 mg/g) by ground SCBA samples was in tune with the literature [11,12,34]. The amorphous
501 phases, solubility, and high SSA governing factors of incredible CH fixation in MCT
502 [11,12,34].
503 The re-calcination protocol, as suggested in the SAI section, is attested by MCT as well. The
504 enhancement in the silica content with re-calcination intensified amorphous silica as well.
505 Consequently, the pozzolanic reaction boosted on account of exposure to the larger surface
506 area. However, the coarser nature of particles as hindrances doesn't show pozzolanic reactivity
507 to this extent in SAI. The grinding to achieve a minimum level particle size below 35 µm
508 appears to be essential to boost the CH consumption to a further level. Finally, the pretreatment
509 dictated in the SAI section looks to be promising in converting SCBA with very high LOI as a
510 pozzolanic material.
511 It is evident from the MCT results that the line distinguishing the pozzolanicity of SCBA can
512 be fixed at a lower limit of CH fixation, as 1100 mg/g to get generalized indication of
513 pozzolanicity of SCBA, provided chemical requirements are in line with ASTM C618 [17],
514 and amorphous silica higher than 25%.

515 3.4 Frattini Test

516 As per the recommendation of BS EN 196-5 [40], the representative point of residual calcium
517 ions and hydroxyl ions under the curve of calcium ion saturation concentration exhibits its

20
518 pozzolanic nature (Fig. 28). According to this guideline, as-such received SCBA samples itself
519 show pozzolanic character, which is not at all corroborating with the SAI. The significantly
520 larger SSA and availability of amorphous silica [65] controlling the reaction kinetics in FT.
521 The coarser particle was exhibiting a loosening effect and large surface area demanding more
522 water escalating water to binder ratio, leading the drop in SAI even though exceptional
523 pozzolanic behaviour in as-such received state. On the contrary, the porous nature of SCBA
524 facilitates the ingress of cement pore solution quickly. The SSA area of SCBA is at such a
525 level that the size of particles is redundant in the pozzolanic reaction between CH and silica
526 (amorphous), in turn, augmentation in reactivity [60]. Nevertheless, the crystalline silica with a
527 significantly finer nature unable to react with CH, as observed for QP.
Saturation Concentration of Ca ions Saturation Concentration of Ca ions
8-ASR 8-C500 8-C600 8-ASR 8-C600
14 8-C700 8-C800 14 8-C600-G30 8-C600-G60
19-ASR 19-C500 19-C600 8-C600-G90 8-C600-G120
19-C700 19-C800 19-ASR 19-C700
12 25-ASR 25-C500 25-C600 12 19-C700-G30 19-700-G60
25-C700 25-C800

CaO (mmol/l)
19-C700-G90 19-C700-G120
CaO (mmol/l)

QP 25-ASR 25-C700
10 10 25-C700-G30 25-C700-G60
25-C700-G90 25-C700-G120

8 8

6 6

4 4

30 40 50 60 70 80 90 100 30 40 50 60 70 80 90 100
[OH]- (mmol/l) [OH]- (mmol/l)

Fig. 28: Frattini test results of SCBA samples re-calcined at various Fig. 29: Frattini test results of re-calcined SCBA samples ground for different
temperatures and QP duration

528 The Frattini test results of SCBA samples re-calcined at various temperatures, and re-calcined
529 samples ground for different duration are depicted in Fig. 28 and Fig. 29, respectively. The
530 SCBA samples in the as-such received state are pozzolanic, and regardless of re-calcination
531 temperature, all samples lie below the Ca ion saturation curve, except sample 19-C800. The
532 recrystallization of silica, i.e., reduction in amorphous content, pushes sample 19-C800 above
533 the line, indicating the non-pozzolanic nature. Further, the SCBA re-calcined at optimum
534 temperature and subsequently ground at various durations is below the Ca ion saturation curve
535 and confirms the pozzolanic nature regardless of the differences in the characteristics.
536 The influence of re-calcination and grinding on the residual concentration of Ca ions is
537 portrayed in Fig. 30 and Fig. 31. With an increase in re-calcination temperature, the amount of
538 residual Ca ions progressively declined up to the optimum re-calcination temperature of the
539 respective SCBA sample, and subsequent re-calcination exhibited higher residual Ca ions. The
540 optimum level of re-calcination was in line with MCT and SAI. Comparatively better
541 performance in terms of pozzolanic reaction is evident for samples 8 and 25. Considerably
542 higher amorphous silica with a larger surface area is the prospective of exceptional pozzolanic

21
543 reactivity. A quick diffusion process of the reaction product layer occurs in the intercellular
544 channel of SCBA on account of its porous nature [66]. Despite substantial surface area, the
545 lower amorphous phases in re-calcined samples of 19 are responsible for comparatively modest
546 performance.
13 13
8 8
12 19 12 19
25 25
11 11
Residual CaO (mmol/l)

Residual CaO (mmol/l)


10 10

9 9

8 8

7 7

6 6

5 5
ASR 500 600 700 800 ASR Opt. Cal. 30 60 90 120
Calcination temperature (oC) Grinding Duration (Minutes)

Fig. 30: Influence of re-calcination of SCBA samples on residual CaO Fig. 31: Influence of grinding of re-calcined SCBA samples on residual CaO
ions in Frattini test ions in Frattini test

547 The influence of the fineness of particles was quite distinct for these re-calcined SCBA
548 samples (Fig. 31). In principle, the residual Ca ions must decline with a decrease in particle
549 size. However, the effect of particle size doesn't seem to be appreciable in terms of reduction in
550 residual Ca ions. Nevertheless, the amount of residual Ca ions was decreased with the grinding
551 of re-calcined samples 19 and 25 for 30 minutes. However, the subsequent finer particles
552 appear to be insignificant in the further enhancement of pozzolanic reactivity. Statistically, the
553 residual Ca ions were consistent beyond 30 minutes of grinding. Although the particle became
554 finer with grinding, the SSA decreased; in turn, the reactivity was compensated and reflected
555 through consistent results. There could be a possibility that the system has reached its
556 equilibrium state due to very high SSA and amorphous silica, in turn, consistency in results.
557 Further, sample 19 presents a significant decline in residual Ca ions with grinding for 30
558 minutes, which confers that the finer particles are essential to utilize available amorphous
559 silica. Surprisingly, the residual Ca ions were increased with grinding of re-calcined SCBA
560 sample 8 for 30 minutes, and further grinding presented progressive declination, yet higher
561 than the re-calcined sample. Pozzolanic reactivity is governed by mainly amorphous silica and
562 SSA, and for sample 8, the surface area dropped significantly with grinding, which has reduced
563 reactivity, in turn, escalation in residual Ca ions. In short, the pores easily accessible for pore
564 solution to form cementitious products are compressed with grinding. However, the MCT
565 could not capture this exciting phenomenon, which may be due to the substantial difference in
566 the test temperature of MCT (90 oC) and FT (40 oC).

22
567 Furthermore, the variability in the concentration of Ca ions and differences in saturated CH
568 solution and cement pore solution possibly render the different portraits. Moreover, the pH of
569 cement solutions is higher (pH >13) than the equilibrium value (~pH 12.4) of CH alone,
570 compensated by Na+ and K+ ions [31]. Probably, this could lead to distinctions in the
571 pozzolanic reactivity of SCBA through these methods.
572 The FT is a sole representation of the pozzolanicity of pozzolanic material. However, SAI is
573 the collaborative effect of pozzolanic reactivity and filler effect [65,67]. The principal
574 differences make it challenging to correlate FT with SAI in the case of SCBA. The above
575 hypothesis suggests assessment with two or more test methods of pozzolanicity to render a real
576 picture of pozzolanic reactivity. The pretreatment indicated in the SAI section is endorsed by
577 the Frattini test as well. However, it enlightens that the level of particle size must be decided
578 with due consideration of amorphous silica and SSA of re-calcined SCBA.

579 3.5 Thermo-gravimetric analysis

580 The effect of thermo-mechanical pretreatment of SCBA on pozzolanic reactivity was analysed
581 through the quantification of portlandite using thermo-gravimetric analysis (TGA). The
582 tangential approach was followed to calculate residual calcium hydroxide. This is assumed to
583 resolve the background loss of hydroxyls by other hydrates over the calcium hydroxide
584 dehydroxylation interval [30,35,36].
585 The dehydroxylation of portlandite generally occurs in the temperature range of 400 to 460oC
586 [68]. The residual portlandite was determined by analyzing the loss in the range of 411oC to
587 479oC, depending upon the onset peak and endothermic peak position in the differential
588 thermo-gravimetric (DTG) curve by the tangential method. The weight loss of CH (WLCH) due
589 to the evaporation of water was used to calculate the amount of CH present with due
590 consideration of the dilution effect as 20%, using the molecular masses of CH (mCH= 74
591 g/mol) and water (mH2O = 18 g/mol) as given in Eq. 2.
74
CH = WL × Eq. 2
18

592 Fig. 32 shows the reduction of portlandite content for the SCBA blended cement paste re-
593 calcined at various temperatures. TGA results are partially in line with observations made for
594 re-calcined SCBA samples in SAI, MCT, and FT. The consumption of CH from the hydrated
595 cement system was escalated with re-calcination temperature up to 700oC for sample 19 and
596 subsequently reduced at 800oC, in line with development in silica and amorphous silica. In the
597 case of sample 8, the consumption of CH was enhanced at re-calcination temperature 500oC
598 and consistent up to 700oC as the amorphous silica was almost similar. However, at 800oC, the
599 inability to improve the CH consumption even though amorphous silica increased could be the

23
600 effect of declination in SSA. For sample 25, as expected, the consumption CH content was
601 enhanced at re-calcination temperature 500oC. However, it was decreased modestly with
602 subsequent re-calcination despite enhancement in the amorphous, and at 800oC, the
603 recrystallization of silica shows inferior performance. Nevertheless, the inert nature of QP was
604 confirmed in TGA as well, with a very negligible reduction in CH as 0.72%.
45 8 50

40 19 45
25
35 40
Reduction in CH (%)

Reduction in CH (%)
35
30
30
25
25
20
20
R2 = 0.76
15
15
10
10
5
5
0 0
ASR 500 600 700 800 -25 0 25 50 75 100 125 150 175 200 225 250 275 300
Calcination temperature (oC) D50 (µm)

Fig. 32: Influence of re-calcination of SCBA samples on the Fig. 33: Influence of particle size (D50) of re-calcined SCBA samples on the
consumption of CH from cement system in TGA consumption of CH from cement system in TGA

605 The influence of grinding on reduction in portlandite content of the SCBA samples re-calcined
606 at optimum temperature is illustrated in Table 5. The mixed outlook came up with TGA
607 compared to the results of SAI, MCT, and FT. As per the observation made in FT for sample 8,
608 the CH consumption with grinding for 30 minutes was dropped, and at 60 minutes, it was
609 consistent, which could be on account of reduction in SSA with grinding. On the other hand,
610 CH consumption was escalated with subsequent grinding at 90 and 120 minutes, emphasizing
611 the necessity of finer particles for a pozzolanic reaction.
612 Table 5: Influence of grinding on reduction in Portlandite content in TGA for
613 re-calcined SCBA samples
Grinding duration Reduction in Portlandite (%)
(minutes)
Samples
8 19 25
Opt. Cal. (no 36.28 31.30 35.17
grinding)
30 33.69 39.41 34.43

60 33.92 33.11 39.15

90 38.08 34.77 34.82

120 41.73 40.39 38.16

614 Similarly, for sample 25, the consumption CH dwindled at 30 minutes of grinding on account
615 of a decrease in SSA. However, at progressive levels of grinding, the CH consumption surged
616 as a result of finer particles (except 90 minutes). In the case of sample 19, there was a
617 significant rise in consumption of CH with grinding for 30 minutes, dictating the importance of
618 smaller particles in lower amorphous silica SCBA samples, and at 120 minutes, it was almost

24
619 similar. Fig. 33 depicts a correlation between particle size (D50) and reduction in CH,
620 exhibiting the role of finer particles in the consumption of CH from the cement system.
621 The literature supports the significant reduction in the portlandite content as 20% [69] and
622 32.05% [56]. The enhanced solubility resulting from the finest particles could be prospective
623 for a considerable pozzolanic reaction [11,12]. No doubt, improvements in total silica and
624 amorphous silica as a result of re-calcination are significantly enhancing CH consumption.
625 Consequently, when results at 120 minutes of grinding are compared with a re-calcined
626 sample, the necessity of finer particles is apparent to achieve the fullest potential of the SCBA
627 samples. Eventually, the thermo-gravimetric analysis also validating the thermo-mechanical
628 pretreatment suggested in the SAI section.

629 3.6 Analysis of hydrated SCBA blended cement paste

630 X-ray diffraction as a qualitative method was explored for understanding the effect of re-
631 calcination and reduction in particle size of SCBA on the formation of hydration products. The
632 few XRD spectrums of hydrated cement paste (HCP) blended with SCBA samples 8, 19, and
633 25 and QP are shown in Fig. 34 to Fig. 36.
P Q: Quartz P: Portlandite P Q: Quartz P: Portlandite
P C: Calcite A: Alite P C: Calcite A: Alite
T: Tobermorite L: Larnite T: Tobermorite L: Larnite
C E: Ettringite Td: Tridymite
T L E: Ettringite Td: Tridymite
Td Q A P C L, T
C,Q T P Td Q T A L C, Q P
P PP P
8-C600-G120 T P PP
19-C700-G120
Intensity
Intensity

8-C600-G30
19-C700-G30
8-C600
19-C700
Q
8-ASR
QQ Q Q
QQ QQ
19-ASR
E QP E

HCP HCP

0 10 20 30 40 50 60 70 80 90 0 10 20 30 40 50 60 70 80 90
2θ (Degree) 2θ (Degree)

Fig. 34: XRD pattern of hydrated cement paste (HCP) blended with SCBA Fig. 35: XRD pattern of hydrated cement paste (HCP) blended with
sample 8 and QP and HCP SCBA sample 19 and HCP

Q: Quartz P: Portlandite
P C: Calcite A: Alite
P
T: Tobermorite L: Larnite
C E: Ettringite Td: Tridymite
Td Q T A LC, Q P P
T P
PP
25-C700-G120
Intensity

25-C700-G30

25-C700

25-ASR
E

HCP

0 10 20 30 40 50 60 70 80 90
2θ (Degree)

Fig. 36: XRD pattern of hydrated cement paste (HCP) blended with SCBA sample 25 and HCP

25
634 The inability of QP to react with CH, on account of crystalline silica observed in FT, MCT,
635 and TGA, is confirmed through the crystalline peaks of quartz. This affirms that the SAI of QP
636 blended mortar to the level of 75% is the result of the filler effect. The distinctive tobermorite
637 [20] peaks at 2θ - 28.76, 31.09, and 40.97 are a sign of the formation of cementitious phases in
638 the presence of re-calcined as well as ground SCBA. The presence of cement compounds Alite
639 and Belite at 2θ - 32.15 and 32.59 in all the samples indicating incomplete hydration of
640 blended cement at the age of 28 days. However, the peak intensity was observed to diminish
641 with the incorporation of re-calcined samples and further with ground SCBA samples,
642 indicating an improvement in the hydration of the cement or could be the effect of dilution in
643 cement content. The portlandite peak intensity corresponding to 2θ - 18.10 was relatively
644 higher than pretreated SCBA samples (8, 19, and 25). However, with the incorporation of
645 treated SCBA, there was a decrement in the intensity.
646 Nevertheless, there were no significant changes in the peak intensity compared with the re-
647 calcined and ground samples. The presence of Portlandite indicates the time-dependent
648 phenomenon of pozzolanic reactivity. Simultaneously, the unreacted silica in the form of
649 quartz and Tridymite (2θ - 26.68, 39.47, and 22.98) was detected in all samples, which was in
650 crystalline form. In consolidation, the significant rise in compressive strength at 56 days for
651 control mortar could be correlated to this phenomenon. The peaks of calcite confirming the
652 carbonation of cement paste to some extent. For sample 19-C700-G30, the calcite peak
653 comparatively intensified, and which could probably be reflected in the escalation in the CH
654 consumption in TGA. The Rietveld refinement analysis shows a reduction in the portlandite
655 content with re-calcination as well as with grinding; however, it doesn't provide conclusive
656 differences to comment on the optimum re-calcination as well as the level of particle size. The
657 enormous amorphous silica and SSA are presenting considerable pozzolanic reactions. The
658 improvements in the formation of cementitious products with re-calcination and grinding are
659 confirmed through X-ray diffractograms.

660 4.0 Conclusions

661 The efforts made through thermo-mechanical pretreatment of SCBA with very high LOI in
662 transmutation to be a pozzolanic material are eventuated to the following conclusions:
663 • The re-calcination temperature 600-700oC is optimum for maintaining the inherently
664 amorphous nature of SCBA with enhancement in the silica content to fulfill the chemical
665 requirements of ASTM C618. However, the residence time and temperature in the
666 suggested range shall be fixed depending upon furnace characteristics and the level of LOI
667 in the raw state of SCBA to bring down at least below 20%.

26
668 • The particle size D50 finer than 35 µm, and particles passing through 45 µm higher than
669 60% are the favorable levels of grinding for achieving the SAI greater than 75%, provided
670 the amorphous silica and AFS content is higher than 25% and 70%, respectively.
671 • The highly pozzolanic nature of re-calcined SCBA is validated through MCT, FT, and
672 TGA test methods, on account of very high SSA and amorphous silica. The ultra-fine
673 grinding of SCBA may not be required to boost the reactivity, as it has a negative impact
674 on the SSA. However, to overcome the effect of loosening and improve the filler effect for
675 enhancement in compressive strength, grinding is essential.
676 • The association of MCT and FT method with SAI is an affirmative way to predict the
677 ingenuous pozzolanic reactivity of SCBA.
678 • The exceptionally lower SAI and AFS content of as-such received SCBA with very high
679 LOI mandates thermo-mechanical treatment before its utilization in cement-based systems.
680 The suggested thermo-mechanical treatment is an effective protocol for the conversion of
681 SCBA with very high LOI to a siliceous pozzolanic material.

682 Acknowledgment

683 The authors sincerely acknowledge all sugar industries for providing samples of sugarcane
684 bagasse ash, as well as giving insights into the process followed in the generation of SCBA.
685 Materials characterisation support from Sophisticated Analytical Instrument Facility,
686 Department of Earth Sciences, and Department of Metallurgical Engineering and Materials
687 Science, IIT Bombay, is acknowledged. The author also wishes to thanks the Transportation
688 engineering laboratory of Walchand College of Engineering, Sangli, Maharashtra, for
689 providing the facility of laser particle size analyser and simultaneous thermal analyzer. Finally,
690 the authors acknowledge the infrastructure and support of TEQIP-III and Walchand College of
691 Engineering, Sangli, Maharashtra.

692 References
693 [1] A. Gopinath, A. Bahurudeen, S. Appari, P. Nanthagopalan, A circular framework for the
694 valorisation of sugar industry wastes: Review on the industrial symbiosis between sugar,
695 construction and energy industries, J. Clean. Prod. 203 (2018) 89–108.
696 [2] G.C. Cordeiro, T.R. Barroso, R.D. Toledo Filho, Enhancement the Properties of Sugar
697 Cane Bagasse Ash with High Carbon Content by a Controlled Re-calcination Process,
698 KSCE J. Civ. Eng. 22 (2018) 1250–1257.
699 [3] S. Deepika, G. Anand, A. Bahurudeen, M. Santhanam, Construction Products with
700 Sugarcane Bagasse Ash Binder, J. Mater. Civ. Eng. 29 (2017) 04017189.
701 [4] ISMA, Indian Sugar Mills Association, (2020).
702 https://www.indiansugar.com/Statics.aspx (accessed April 15, 2020).
703 [5] G. Athira, A. Bahurudeen, P.K. Sahu, M. Santhanam, P. Nanthagopalan, S. Lalu,

27
704 Effective utilization of sugar industry waste in Indian construction sector : A geospatial
705 approach, J. Mater. Cycles Waste Manag. 22 (2020) 724–736.
706 [6] K.L. Scrivener, V.M. John, E.M. Gartner, Eco-efficient cements : Potential
707 economically viable solutions for a low-CO2 cement-based materials industry, Cem.
708 Concr. Res. 114 (2018) 2–26.
709 [7] A. Bahurudeen, K. Wani, M.A. Basit, M. Santhanam, Assessment of Pozzolanic
710 Performance of Sugarcane Bagasse Ash, J. Mater. Civ. Eng. 28 (2016) 04015095.
711 [8] G.C. Cordeiro, L.M. Tavares, R.D. Toledo Filho, Improved pozzolanic activity of sugar
712 cane bagasse ash by selective grinding and classification, Cem. Concr. Res. 89 (2016)
713 269–275.
714 [9] G.C. Cordeiro, R.D. Toledo Filho, L.M. Tavares, E. de M.R. Fairbairn, Ultrafine
715 grinding of sugar cane bagasse ash for application as pozzolanic admixture in concrete,
716 Cem. Concr. Res. 39 (2009) 110–115.
717 [10] A. Bahurudeen, M. Santhanam, Influence of different processing methods on the
718 pozzolanic performance of sugarcane bagasse ash, Cem. Concr. Compos. 56 (2015) 32–
719 45.
720 [11] G.C. Cordeiro, P.V. Andreão, L.M. Tavares, Pozzolanic properties of ultrafine sugar
721 cane bagasse ash produced by controlled burning, Heliyon. 5 (2019) 0–5.
722 [12] P.V. Andreão, A.R. Suleiman, G.C. Cordeiro, M.L. Nehdi, Sustainable use of sugarcane
723 bagasse ash in cement-based materials, Green Mater. 7 (2019) 61–70.
724 [13] R. Mohan, G. Athira, A.K. Mali, A. Bahurudeen, P. Nanthagopalan, Systematic
725 Pretreatment Process and Optimization of Sugarcane Bagasse Ash Dosage for Use in
726 Cement-Based Products, J. Mater. Civ. Eng. 33 (2021) 04021045.
727 [14] L.M.S. Souza, E.M.R. Fairbairn, G.C. Cordeiro, R.D.T. Filho, Hydration Study of Sugar
728 Cane Bagasse Ash and Calcium Hydroxide Pastes of Various Initial C/S Ratios, in:
729 Second Int. Conf. Microstruct. Durab. Cem. Compos. 11-13 April 2012, Amsterdam,
730 Netherlands, .
731 [15] R. Somna, C. Jaturapitakkul, A.M. Amde, Effect of ground fly ash and ground bagasse
732 ash on the durability of recycled aggregate concrete, Cem. Concr. Compos. 34 (2012)
733 848–854.
734 [16] A.K. Mali, P. Nanthagopalan, A systematic assessment for the determination of
735 sugarcane bagasse ash variation potential throughout the harvesting season, Mater.
736 Today Proc. 32 (2020) 888–895.
737 [17] ASTM C618, Standard Specification for Coal Fly Ash and Raw or Calcined Natural
738 Pozzolan for Use, ASTM International. (2019).
739 [18] G.C. Cordeiro, R.D.T. Filho, R.S. de Almeida, Influence of ultrafine wet grinding on
740 pozzolanic activity of submicrometre sugar cane bagasse ash, Adv. Appl. Ceram. 110
741 (2011) 453–456.
742 [19] K. Ganesan, K. Rajagopal, K. Thangavel, Evaluation of bagasse ash as supplementary
743 cementitious material, Cem. Concr. Compos. 29 (2007) 515–524.
744 [20] A. Rajasekar, K. Arunachalam, M. Kottaisamy, V. Saraswathy, Durability
745 characteristics of Ultra High Strength Concrete with treated sugarcane bagasse ash,
746 Constr. Build. Mater. 171 (2018) 350–356.
747 [21] IS 269, Ordinary Portland Cement- Specification, Bureau of Indian Standards. (2015).
748 [22] IS 650, Standard Sand for Testing Cement-Specification, Bureau of Indian Standards.
28
749 (2018).
750 [23] ASTM C188, Standard Test Method for Density of Hydraulic Cement, ASTM
751 International. (2017).
752 [24] ASTM C114, Standard Test Methods for Chemical Analysis of Hydraulic Cement,
753 ASTM International. (2018).
754 [25] G.C. Cordeiro, R.D. Toledo Filho, E.M.R. Fairbairn, Effect of calcination temperature
755 on the pozzolanic activity of sugar cane bagasse ash, Constr. Build. Mater. 23 (2009)
756 3301–3303.
757 [26] Y. Kanda, N. Kotake, Chapter 12 Comminution Energy and Evaluation in Fine
758 Grinding, in: Handb. Powder Technol., Elsevier B.V., : pp. 529–550.
759 [27] H. Cho, J. Kwon, K. Kim, M. Mun, Optimum choice of the make-up ball sizes for
760 maximum throughput in tumbling ball mills, Powder Technol. 246 (2013) 625–634.
761 [28] A.S. Erdem, Ş.L. Ergün, The effect of ball size on breakage rate parameter in a pilot
762 scale ball mill, Miner. Eng. 22 (2009) 660–664.
763 [29] X. Li, R. Snellings, M. Antoni, N.M. Alderete, M. Ben Haha, S. Bishnoi, O. Cizer, M.
764 Cyr, K. De Weerdt, Y. Dhandapani, J. Duchesne, J. Haufe, D. Hooton, M. Juenger, S.
765 Kamali-Bernard, S. Kramar, M. Marroccoli, A.M. Joseph, A. Parashar, C. Patapy, J.L.
766 Provis, S. Sabio, M. Santhanam, L. Steger, T. Sui, A. Telesca, A. Vollpracht, F. Vargas,
767 B. Walkley, F. Winnefeld, G. Ye, M. Zajac, S. Zhang, K.L. Scrivener, Reactivity tests
768 for supplementary cementitious materials : RILEM TC 267-TRM phase 1, Mater. Struct.
769 51:151 (2018) 1–14.
770 [30] G. Mertens, R. Snellings, K. Van Balen, B. Bicer-simsir, P. Verlooy, J. Elsen,
771 Pozzolanic reactions of common natural zeolites with lime and parameters affecting
772 their reactivity, Cem. Concr. Res. 39 (2009) 233–240.
773 [31] S. Donatello, M. Tyrer, C.R. Cheeseman, Comparison of test methods to assess
774 pozzolanic activity, Cem. Concr. Compos. 32 (2010) 121–127.
775 [32] F. Avet, R. Snellings, A. Alujas, M. Ben, K. Scrivener, Cement and Concrete Research
776 Development of a new rapid , relevant and reliable ( R3 ) test method to evaluate the
777 pozzolanic reactivity of calcined kaolinitic clays, Cem. Concr. Res. 85 (2016) 1–11.
778 [33] V.A. Quarcioni, F.F. Chotoli, A.C. V Coelho, M.A. Cincotto, Indirect and direct
779 Chapelle’s methods for the determination of lime consumption in pozzolanic materials,
780 IBRACON Struct. Mater. 8 (2015) 1–7.
781 [34] P. V. Andreão, A.R. Suleiman, G.C. Cordeiro, M.L. Nehdi, Beneficiation of Sugarcane
782 Bagasse Ash: Pozzolanic Activity and Leaching Behavior, Waste and Biomass
783 Valorization. (2019).
784 [35] P.T. Durdzinski, M. Ben Haha, S.A. Bernal, N. De Belie, E. Gruyaert, B. Lothenbach,
785 E.M. Mendez, J.L. Provis, S. Axel, C. Stabler, Z. Tan, Y. Villagran-Zaccardi, A.
786 Vollpracht, F. Winnefeld, M. Zajac, K.L. Scrivener, Outcomes of the RILEM round
787 robin on degree of reaction of slag and fly ash in blended cements, Mater. Struct. 50
788 (2017) 1–15.
789 [36] R. Snellings, J. Chwast, O. Cizer, N. De Belie, Y. Dhandapani, P. Durdzinski, J. Elsen,
790 J. Haufe, D. Hooton, C. Patapy, M. Santhanam, K. Scrivener, D. Snoeck, L. Steger, S.
791 Tongbo, A. Vollpracht, F. Winnefeld, B. Lothenbach, Report of TC 238-SCM :
792 hydration stoppage methods for phase assemblage studies of blended cements — results
793 of a round robin test, Mater. Struct. 51:111 (2018) 1–12.
29
794 [37] ASTM C311, Standard Test Methods for Sampling and Testing Fly Ash or Natural
795 Pozzolans for Use, ASTM International. (2019).
796 [38] ASTM C1437, Standard Test Method for Flow of Hydraulic Cement Mortar, ASTM
797 International. (2015).
798 [39] IS 16534, Metakaolin for Use in Cement, Cement Mortar and Concrete Specification,
799 Bureau of Indian Standards. (2015).
800 [40] BS EN 196-5, Methods of testing cement Part 5 : Pozzolanicity test for pozzolanic
801 cement, BSI Stand. Publ. (2011).
802 [41] V.S. Batra, S. Urbonaite, G. Svensson, Characterization of unburned carbon in bagasse
803 fly ash, Fuel. 87 (2008) 2972–2976.
804 [42] P. Jagadesh, A. Ramachandramurthy, R. Murugesan, Evaluation of mechanical
805 properties of Sugar Cane Bagasse Ash concrete, Constr. Build. Mater. 176 (2018) 608–
806 617.
807 [43] E. Villar-Cociña, M. Frías, J. Hernández-Ruiz, H. Savastano, Pozzolanic behaviour of a
808 bagasse ash from the boiler of a Cuban sugar factory, Adv. Cem. Res. 25 (2012) 136–
809 142.
810 [44] E. Arif, M.W. Clark, N. Lake, Sugar cane bagasse ash from a high efficiency co-
811 generation boiler: Applications in cement and mortar production, Constr. Build. Mater.
812 128 (2016) 287–297.
813 [45] J.F.M. Hernandez, B. Middendorf, M. Gehrke, H. Budelmann, Use of Wastes of the
814 Sugar Industry as Pozzolana in Lime-Pozzolana Binders: Study of the Reaction, Cem.
815 Concr. Res. 28 (1998) 1525–1536.
816 [46] G.C. Cordeiro, O.A. Paiva, R.D. Toledo Filho, E.M.R. Fairbairn, L.M. Tavares, Long-
817 Term Compressive Behavior of Concretes with Sugarcane Bagasse Ash as a
818 Supplementary Cementitious Material, J. Test. Eval. 46 (2018) 20160316.
819 [47] A. Bahurudeen, D. Kanraj, V. Gokul Dev, M. Santhanam, Performance evaluation of
820 sugarcane bagasse ash blended cement in concrete, Cem. Concr. Compos. 59 (2015) 77–
821 88.
822 [48] R.A. Berenguer, A.P.B. Capraro, M.H.F. de Medeiros, A.M.P. Carneiro, R.A. De
823 Oliveira, Sugar cane bagasse ash as a partial substitute of Portland cement: Effect on
824 mechanical properties and emission of carbon dioxide, J. Environ. Chem. Eng. 8 (2020)
825 103655.
826 [49] A. Bahurudeen, A.V. Marckson, A. Kishore, M. Santhanam, Development of sugarcane
827 bagasse ash based Portland pozzolana cement and evaluation of compatibility with
828 superplasticizers, Constr. Build. Mater. 68 (2014) 465–475.
829 [50] J.K.T. Krishnarao R.V., Subrahmanyam J., Studies on the formation of black particles in
830 rice husk silica ash, J. Eur. Ceram. Soc. 21 (2001) 99–104.
831 [51] IS 3812 (Part 1), Pulverized fuel ash- specification (Part 1 for use as pozzolana in
832 cement, cement mortar and concrete), Bureau of Indian Standards. (2017).
833 [52] M. Frías, E. Villar, H. Savastano, Brazilian sugar cane bagasse ashes from the
834 cogeneration industry as active pozzolans for cement manufacture, Cem. Concr.
835 Compos. 33 (2011) 490–496.
836 [53] P. Jagadesh, A. Ramachandramurthy, R. Murugesan, K. Sarayu, Micro-analytical
837 studies on sugar cane bagasse ash, Sadhana - Acad. Proc. Eng. Sci. 40 (2015) 1629–
838 1638.
30
839 [54] P. Jagadesh, A. Ramachandramurthy, R. Murugesan, T. Karthik Prabhu, Adaptability of
840 Sugar Cane Bagasse Ash in Mortar, J. Inst. Eng. Ser. A. 100 (2019) 225–240.
841 [55] S. Kumar, F. Kristály, G. Mucsi, Geopolymerisation behaviour of size fractioned fly
842 ash, Adv. Powder Technol. 26 (2015) 24–30.
843 [56] J. Payá, J. Monzó, M. V. Borrachero, L. Díaz-Pinzón, L.M. Ordónez, Sugar-cane
844 bagasse ash (SCBA): Studies on its properties for reusing in concrete production, J.
845 Chem. Technol. Biotechnol. 77 (2002) 321–325.
846 [57] K. Montakarntiwong, N. Chusilp, W. Tangchirapat, C. Jaturapitakkul, Strength and heat
847 evolution of concretes containing bagasse ash from thermal power plants in sugar
848 industry, Mater. Des. 49 (2013) 414–420.
849 [58] D.V. Ribeiro, M.R. Morelli, Effect of Calcination Temperature on the Pozzolanic
850 Activity of Brazilian Sugar Cane Bagasse Ash (SCBA), Mater. Res. 17 (2014) 974–981.
851 [59] G.C. Cordeiro, R.D.T. Filho, L.M. Tavares, E.D.M.R. Fairbairn, S. Hempel, Influence
852 of particle size and specific surface area on the pozzolanic activity of residual rice husk
853 ash, Cem. Concr. Compos. 33 (2011) 529–534.
854 [60] V.T.A. Van, C. Rößler, D.D. Bui, H.M. Ludwig, Mesoporous structure and pozzolanic
855 reactivity of rice husk ash in cementitious system, Constr. Build. Mater. 43 (2013) 208–
856 216.
857 [61] G.C. Cordeiro, R.D.T. Filho, E.D.M.R. Fairbairn, Use of ultrafine rice husk ash with
858 high-carbon content as pozzolan in high performance concrete, Mater. Struct. Constr. 42
859 (2009) 983–992.
860 [62] E. Arif, M.W. Clark, N. Lake, Sugar cane bagasse ash from a high-efficiency co-
861 generation boiler as filler in concrete, Constr. Build. Mater. 151 (2017) 692–703.
862 [63] N. Chusilp, C. Jaturapitakkul, K. Kiattikomol, Effects of LOI of ground bagasse ash on
863 the compressive strength and sulfate resistance of mortars, Constr. Build. Mater. 23
864 (2009) 3523–3531.
865 [64] G.C. Cordeiro, R.D. Toledo Filho, L.M. Tavares, E.M.R. Fairbairn, Pozzolanic activity
866 and filler effect of sugar cane bagasse ash in Portland cement and lime mortars, Cem.
867 Concr. Compos. 30 (2008) 410–418.
868 [65] A. Tironi, M.A. Trezza, A.N. Scian, E.F. Irassar, Assessment of pozzolanic activity of
869 different calcined clays, Cem. Concr. Compos. 37 (2013) 319–327.
870 [66] E. V. Morales, E. Villar-Cociña, M. Frías, S.F. Santos, H. Savastano, Effects of
871 calcining conditions on the microstructure of sugar cane waste ashes (SCWA): Influence
872 in the pozzolanic activation, Cem. Concr. Compos. 31 (2009) 22–28.
873 [67] J. Skibsted, R. Snellings, Reactivity of supplementary cementitious materials (SCMs) in
874 cement blends, Cem. Concr. Res. 124 (2019) 105799.
875 [68] R. Maddalena, A. Hamilton, A.K. Mali, Low-pressure nano-silica injection on cement
876 for crack-healing and water transport, in: Proc. 10th Int. Conf. Struct. Anal. Hist. Constr.
877 (SAHC, Leuven, Belgium, 13-15 Sept. 2016), CRC Press, : pp. 579–584.
878 [69] G.C. Cordeiro, K.E. Kurtis, Effect of mechanical processing on sugar cane bagasse ash
879 pozzolanicity, Cem. Concr. Res. 97 (2017) 41–49.

31

Você também pode gostar