You are on page 1of 39

Accepted Manuscript

Title: Algae derived biodiesel using nanocatalytic


transesterification process

Author: Siow Hwa Teo Aminul Islam Yun Hin Taufiq-Yap

PII: S0263-8762(16)30051-X
DOI: http://dx.doi.org/doi:10.1016/j.cherd.2016.04.012
Reference: CHERD 2258

To appear in:

Received date: 15-10-2015


Revised date: 4-3-2016
Accepted date: 13-4-2016

Please cite this article as: Teo, S.H., Islam, A., Taufiq-Yap, Y.H.,Algae derived biodiesel
using nanocatalytic transesterification process, Chemical Engineering Research and
Design (2016), http://dx.doi.org/10.1016/j.cherd.2016.04.012

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.
1 Graphic abstract

t
ip
6

cr
8

us
9

an
M
d
p te
ce
Ac

1
Page 1 of 38
9 Research highlights:
10

11 Solid methoxide catalyst was developed via an in situ hydrothermal process.

12 Transesterification of photoautotrophic Nannochloropsis sp. algae oil to

13 biodiesel.

t
ip
14 Optimization of the process parameters.

15 Kinetic analysis of the process involved in biodiesel production.

cr
16

us
17
18

an
19
20
M
21
22
23
d

24
te

25
26
p

27
28
ce

29
30
Ac

31
32
33
34
35

2
Page 2 of 38
35 Algae derived biodiesel using nanocatalytic transesterification process

36
37 Siow Hwa Teo1,2, Aminul Islam1,2,3 and Yun Hin Taufiq-Yap*1,2,4
1
38 Catalysis Science and Technology Research Centre, Faculty of Science, Universiti Putra

39 Malaysia, 43400, Serdang, Selangor, Malaysia.

t
ip
2
40 Department of Chemistry, Faculty of Science, Universiti Putra Malaysia, 43400, Serdang,

41 Selangor, Malaysia.

cr
3
42 Faculty of Engineering, Universiti Malaysia Sabah, 88400 Kota Kinabalu, Sabah, Malaysia.

us
4
43 Curtin Sarawak Research Institute, Curtin University, Miri, Sarawak, Malaysia.

44

an
45

46
M
47

48
d

49
te

50
p

51
ce

52 Y. H. Taufiq-Yap (Corresponding author)

53 Catalysis Science and Technology Research Centre,


Ac

54 Faculty of Science, Universiti Putra Malaysia, 43400,

55 UPM Serdang, Selangor, Malaysia.

56 Tel: +603-89466809, Fax: +603-89466758,

57 Email: taufiq@upm.edu.my

58

59

60

3
Page 3 of 38
61 Abstract

62 This work investigates the nanocatalytic biodiesel production from algae (Nannochloropsis

63 sp.). The hydrothermal synthesis route was used in this study to produce nano Ca(OCH3)2

64 (calcium methoxide) as a model catalyst. The effect of the main reaction parameters i.e.

t
65 catalyst dosage, temperatures under constant pressure, methanol molar ratio and reaction time

ip
66 on the yield of FAME (fatty acid methyl ester) were examined. Kinetic study of biodiesel

cr
67 synthesis from crude microalgae oil using nanocatalytic transesterification reaction was

68 appraised. The results indicate that CH3O- species (a cluster of tiny plate-like architectures) in

us
69 Ca(OCH3)2 catalyst, and acted as main active sites for transesterification process. In

an
70 additional, Ca(OCH3)2 catalyst has excellent catalytic performance in production of biodiesel.

71 The highest FAME yield of 99.0 % was obtained over 3 wt.% of Ca(OCH3)2 catalyst loading
M
72 at methanol to oil molar ratio of 30:1 and reaction time of 3 h at 80 °C. Moreover, the catalyst

73 displays a good stability and reutilization. A satisfactory FAME yield of 96 % was achieved
d

74 after use for five consecutive cycles without significant deactivation. The activation energy
te

75 (Ea) of the transesterification reaction of crude N. oculata oil with methanol over Ca(OCH3)2

76 nanocatalyst was obtained as 58.62 KJ mol-1. The results revealed that the yield of methyl
p

77 esters obtained from algae-based triglycerides was follows a pseudo first order mechanism
ce

78 for the forward reaction. These results suggest that the nanocatalyst is a promising for a green

79 biodiesel production process from algae.


Ac

80

81

82 Keywords: Nannochloropsis oculata; Biodiesel; Transesterification; Calcium methoxide;

83 Kinetic; Thermodynamics

84

85

4
Page 4 of 38
86 1. Introduction

87 One of the major environmental problems associated with the consumption of fossil

88 fuel resources is global warming and climate change. Hence, many of the developed country

89 including Italy, USA, France and Brazil have taken initiative to enhance the alternative

t
90 energy research. It was reported that the reserves of coal, oil and natural gas will be

ip
91 exhausted by 122, 42 and 60 years, respectively (Islam et al., 2014). Therefore, researchers

cr
92 have directed towards the alternative renewable energy sources to reduce and replace

93 petroleum derived diesel fuel.

us
94 Recently, biodiesel is made entirely from a wide spectrum of feedstocks i.e. edible

an
95 (virgin oil) non-edible or waste cooking, waste frying, animal or fish waste and microalgae

96 oils, and its merits as an alternative, renewable energy source, biodegradable, environmental
M
97 friendly (Birla et al., 2012; Palligarnai and Michael, 2008). Edible oil contributes 95 % of the

98 feedstock constituent for biodiesel production (Hassan and Vinjamur, 2014). Nevertheless,
d

99 the use of food-based feedstock extensively for production of biodiesel is shrinking arable
te

100 land availability and the worldwide concern over the fuel versus food debate. The shortage of

101 the edible oil in many developing and undeveloped countries remains a question mark to
p

102 further production of biodiesel from edible oils (Hassan and Vinjamur, 2014). Hence, it does
ce

103 not seem to be an ideal justification to divert edible oil to sustainable biofuels production. On

104 the other hand, discovering the potentials of the non-edible feedstocks such as Jatropha
Ac

105 curcas and Pongamia pinnata to be high efficient for making biodiesel (Boucher et al., 2008).

106 Apart from this, the other plant species including microalgae oil also indicated as a novel

107 feedstock, which have a high potential to convert biodiesel with high yield (Nautiyal et al.,

108 2014). Newly, algae-based lipid for biodiesel is in the beginning phases of research, however

109 it has posed enormous prospective to offer alternative new generation feedstock (Lam et al.,

110 2012). The advantage of the algae based feedstock is that the lipid productivity can be 15-300

5
Page 5 of 38
111 times (respect to the dry weight of biomass) larger than that derived from plant (Lam et al.,

112 2012; Lee at el., 2010). Nonetheless, the optimization of important parameters i.e. light, pH,

113 temperature, nutrients in their cultivation and also the extraction of oil for a high yield with

114 low cost is currently a challenge (Lam et al., 2012; Lee at el., 2010; Lam and Lee, 2012).

t
115 The use of different supported earth metal oxide, zirconia and zeolite based oxide

ip
116 catalysts have been reported in the literature (Xie and Zhao, 2014; Xie et al. 2011; Xie and

cr
117 Fan 2014; Xie and Zhao, 2013). Generally, most of these catalysts have been posed favorable

118 outcomes with high conversion of biodiesel. It may be mentioned that the calcium derived

us
119 based were still remain to be the most potential catalyst as they are abundance, inexpensive,

an
120 show low methanol solubility and also non-toxic among the catalysts studies recently.

121 A noteworthy amount of research work has been conducted on the kinetic reaction for
M
122 producing biodiesel. Uzun et al. (2012) reported the reaction kinetic of base catalyzed

123 transesterification using waste frying oil with excess methanol. Reaction kinetics of palm oil
d

124 based biodiesel has been reported recently by Shahbazi et al. (2012). The activation energies
te

125 of transesterification on different feedstocks for biodiesel production has been reported by

126 Asakuma et al. (2009) using the quantum computation chemistry. The kinetic approach of
p

127 algae (N. oculata) derived biodiesel using Ca(OCH3)2 nanocatalyst was reported in this study.
ce

128 The effect of reaction parameters on biodiesel yield and the rate constant for the Ca(OCH3)2

129 catalyzed transesterification of N. oculata were evaluated. Finally, the catalyst reusability in
Ac

130 the transesterification of microalgae N. oculata to biodiesel was also studied.

131

132

133

134

135

6
Page 6 of 38
136 2. Experimental

137 2.1. Materials

138 Pure microalgae N. oculata strain was obtained from AlgaeTech Sdn. Bhd., Malaysia.

139 Calcium oxide (CaO) >90 % was purchased from SigmaAldrich Chemical. All other reagents

t
140 used for this study were analytical grade.

ip
141
142 2.2. Microalgae biomass preparation

cr
143 500 ml of decontaminated microalgae N. oculata strain further propagation using

us
144 photoautotrophic cultivation method. The culture was flocculated, washed and triturated to

145 collect microalgae biomass. The lipid extraction was performed using modified bligh and

an
146 dyer method (Nautiyal et al., 2014). The microalgae were subjected to pulverization in mortar

147 for cell disruption follow by mesh sieving. The resultant powdered biomass was placed in the
M
148 500 ml conical flask along with methanol-chloroform (1:2 v/v) solvents. The contents in the

149 flask were stirred with the help of magnetic stirrer at a particular speed. In order to prevent
d

150 any losses of hexane and methanol due to evaporation the set up was completely covered
te

151 with aluminum fold. After the stipulated time, the stirring was stopped and the products were
p

152 filtered. The crude lipid was concentrated, and then stored in a dark, well close, glass bottle in
ce

153 a freezer for transesterification reaction.

154
155 2.3. Crude lipid analysis
Ac

156 The samples were analyzed by the method described elsewhere (Ariffin et al., 2009).

157 Conventional sodium hydroxide was used for converting fatty acid to methyl ester. After the

158 stipulated of the reaction, the products were allowed to settle into two layers. The upper layer

159 comprised solvent with biodiesel was transferred into a micro-centrifugal. The analysis of

160 biodiesel was performed by using GC-2010 series Gas chromatograph coupled to a mass

161 spectrometer (Shimadzu GCMS-QP2010).Moreover; the contents of moisture, free fatty acid

7
Page 7 of 38
162 and dried ash from crude lipid were determined using previous reported method

163 (Andrikopoulas et al., 2001).

164
165 2.4. Catalyst preparation

166 Ca(OCH3)2 catalyst was synthesized by refluxing calcium oxide (CaO) with dehydrated

t
ip
167 methanol at 65 °C in condition of N2 flow for 8 h. The following Eq. (1) states the reaction:
65 ○C, N2 atmosphere
168 CaO + 2CH3OH Ca(OCH3)2 + H2O (1)

cr
169 The experiment was performed in a reactor having a capacity of 500 ml. In a typical

us
170 preparation process, CaO (10 g) was suspended into 300 ml of methanol. The methanol was

171 then distilled until dryness using a rotary evaporator. The catalyst was powdered with mortar

an
172 and pestle and sieve with 250 μm to obtain micron sized catalyst for the reaction.

173
M
174 2.5. Catalyst characterization

175 The X-ray diffraction (XRD) of the catalyst was recorded by diffractometer (XRD-6000,

Shimadzu) using CuKα radiation at scanning rate of 2° min-1. The total surface area (SBET),
d

176
te

177 total pore volume (cm3 g-1) and average pore size (nm) of the catalyst were evaluated by

178 Sorptometer (1900 series, Thermo Finnigan Sorptomatic). Surface area was calculated using
p

179 Brunauer-Emmett-Teller (BET) equation form the adsorption/desorption branch of the


ce

180 isotherm in a relative pressure range of 0.07-0.3. Whereas, the pore size distribution was

181 determined from desorption branches derived from Barrett–Joyner–Halenda (BJH) plot at -
Ac

182 196 °C temperature. Thermogravimetric (TGA) and differential thermal analyses (DTA) were

183 performed using a Mettler Toledo thermogravimetric analyzer. The heating was carried out in

184 a normal air flow and maintained at 100 ml min-1. The heating rate was 10 °C min-1. The

185 analyses of the catalyst were determined using attenuated total reflection-Fourier transform-

186 infrared spectroscopy (ATR-FTIR). Infrared absorption spectra were obtained using a FTIR

187 spectrometer (PerkinElmer (PC) Spectrum 100 FTIR spectrometer) with He-Ne laser source

8
Page 8 of 38
188 at N2 temperature. The average spectrum was recorded form triplicate analysis ranging from

189 400 – 4000 cm-1 at a resolution of 4 cm-1. Scanning electron microscopy (SEM) was taken

190 with a JOEL JSM6700F Field Scanning Electron Microscope (FESEM). To protect the

191 induction of electric current, the catalyst was coated with Au (gold) by a sputter coater with

t
192 maximum operating voltage used was 25 kV. Hitachi H7100 Transmission Electron

ip
193 Microscopy (TEM) was used to measure the particle size of the catalyst with a voltage of 200

cr
194 kV.

195

us
196 2.6. Transesterification reaction

197 The experiments were performed using a glass-type instrument with water-cooled reflux

an
198 condenser and stirrer. A thermocouple was used to determine the temperature of the reaction. .

199 Transesterification was carried out with different methanol/oil ratio (10-60 molar ratios),
M
200 catalyst concentrations (3-15 % weight), temperatures ranging from 50-80 ◦C and reaction

201 times (30-240 min). To start each experiment, a typical amount of oil, catalyst and methanol
d

202 were added together into the reactor. Then, the mixture was heated in the silicon oil bath on a
te

203 hot plate with continuous stirring to get required temperatures. The product was settled into
p

204 three layers. The uppermost phase comprised of the solvent along with biodiesel was
ce

205 collected. Finally, evaporation was performed to remove excess solvent from the reaction

206 mixture.
Ac

207
208 2.7. Catalyst regeneration

209 Reusability and regeneration tests were carried out by reutilizing the spent catalyst after every

210 reaction. The mixture solvents (biodiesel and methanol) were removed from the reaction

211 system from the microreactor, then the required amount of fresh crude lipid and methanol

212 were put in to the microreactor. Lastly, the reaction was carried out under optimized

9
Page 9 of 38
213 conditions (MeOH : oil : catalyst wt% (30 : 1 : 3) at 80 ◦C for 3 h). The above procedures

214 were repeated until the significant drop on the FAME yield production.

215
216 2.8. FAME yield analysis

217 Gas chromatography (GC) (model 7890A, Agilent technologies) equipped with flame

t
ip
218 ionization detector (FID) was used to analyze the methyl ester yield. The silica capillary

cr
219 column (Agilent, HP-88 USA) with dimensions 300 m x 0.25 mm x 0.20 μm film thickness

220 was used in this experiment. The helium was used as a carrier gas during the analysis. Methyl

us
221 heptadecanoate and hexane were used as internal standard and solvent during the analysis of

222 biodiesel. The chromatogram of biodiesel was shown in Supp. 1. The FAME yield was

223

224 triplicate measures was reported in all experiment.an


calculated according to EN 14103 procedures (Liu et al., 2006). An average value of
M
225
226 2.8. Kinetic model
d

227 Transesterification of solid catalyst occurs at harsh reaction condition (higher temperature
te

228 and pressure) has been reported. This phenomenon was due to the fact that three-phase

229 catalytic reaction system slows down the transesterification reaction due to the mass transfer
p

230 limitations (Arifin et al., 2009; Liu et al., 2006; Vyas et al., 2009; Rathore and Madras, 2007;
ce

231 Zhang et al., 2010). Stoichiometrically, reaction requires 3 mol of methanol (M) and 1 mol of

232 triglyceride (TG) to give 3 mol of methyl ester (ME) and 1 mol of glycerol (GL), respectively,
Ac

233 and the overall equation is shown as following Eq. (2).

234 In general, transesterification reaction comprises three consecutive reversible

235 reactions (Eq. (3)), whereby 1 mol of ME is produced in each step and monoglycerides (MG)

236 (Eq. (5)) and diglycerides (DG) (Eq. (4)) are intermediate product (Islam et al., 2013b).

237
238 Overall reaction
catalyst
239 TG + 3R’OH 3R’COOR + GL (2)

10
Page 10 of 38
240 Stepwise reactions
k1
241 TG + R’OH DG + R’COOH (3)
k1’
242
k2
243 DG + R’OH MG + R’COOH (4)
k2’

t
244

ip
k3
245 MG + R’OH GL + R’COOH (5)
k3’

cr
246 Furthermore, the following assumption are used in this kinetic model i.e. (i) keq. must

247 be not depending on methanol concentration and the reaction is considered pseudo-1st order).

us
248 (ii) production of intermediate species is negligible and (iii) all chemical reaction only

an
249 occurred in the oil phase. Therefore, first assumption (i) can be written as Eq. (6) as follows:

250 (6)
M
251 Next, based on the second assumption (ii) (Eq. (8)),

252 (7)
d

253 (8)
te

254 (9)
p

255 And, according to the mass balance,


ce

256 (10)

257 (11)
Ac

258 (12)

259 Based on the equations above and experimental data, firstly, the concentration of ME at

260 various reaction times (based on the moles fraction) was obtained. Secondly, a graph with –ln

261 (1-XME) versus (T) was plotted using the Eq. (10) and the rate constant at each temperature

262 were measured. Therefore, activation energy and pre-exponential factor was calculated by the

263 Arrhenius equation (Eq. (13)):

11
Page 11 of 38
264 (13)

265 (14)

266 where k is the reaction constant, A is the frequency or pre-exponential factor, Ea is the

267 activation energy of the reaction, R is the gas constant and T is the absolute temperature. Plot

t
ip
268 of ln k versus 1/T are given from equation above (Eq. (14)), then Ea and pre-exponential

269 factor have been calculated.

cr
270
271 3. Results and discussion

us
272 3.1. Lipid characterization

273 The method reported by Bligh and Dyer was used to extract the highest lipid content from

274

275 an
microalgae (Nannochloropsis sp.). As shown in Supp. 1 and Table 1, the fatty acid

compositions of Nannochloropsis sp. contained mainly methyl esters of palmitic acid (35.43
M
276 %), palmitoleic acid (27.54 %), Oleic acid (8.62 %), EPA (8.29 %) and myristic acid (7.69

277 %), respectively. Interestingly, the data exhibited that methyl ester of approximately 70 %
d

278 saturated fatty acid and 30 % unsaturated fatty acid present in the biodiesel. The lower
te

279 percentage of unsaturated fatty acids in algae biodiesel makes it more stable as compared to
p

280 palm and tallow biodiesel (Sarin et al., 2007; Alcantara et al., 2000). Therefore, the calculated
ce

281 molecular weight (MWoil) of microalgae oil was 831.62 g mol-1 according to the following

282 Eq. (15).


Ac

283 MWoil = [(3 x MWfatty acid) + MWglycerol] - 3 x MWwater (15)

284 Furthermore, the crude lipid also included a small fraction of moisture (5.47 %), free fatty

285 acid (8.03 %) and dried ash (10.68 %) (Table 1(b)).

286
287 3.2. Catalyst Characterization

288 The TG-DTA profile of Ca(OCH3)2 catalyst under air flow environment is appeared in Fig. 1.

289 The TGA curve remained constant with the temperatures ranging from 35 °C to 370 °C with

12
Page 12 of 38
290 minor weight loss indicating loss of moisture or volatile matters. The exothermic peak

291 centered at 430 °C was attributed to the Ca(OCH3)2 decomposition reaction under oxygen

292 condition to form calcium carbonate (Eq. (16)):

293 Ca (OCH3)2 + 3O2 → CaCO3 + CO2 + 3H2O (16)

t
294 Likewise, the second stage of weight loss was in the temperature range of 600-800 °C

ip
295 from the TGA curve indicating the thermal decomposition of calcium carbonate. At 720 °C,

cr
296 the transformation of CaCO3 to stable CaO was caused, as confirmed from the DTA peak.

297 The thermogravimetry analysis indicated that the synthesized Ca(OCH3)2 catalyst is stable up

us
298 to 400 °C. A X-ray pattern of CaO and Ca(OCH3)2 catalysts is shown in Fig. 2. Two peaks at

2θ values of 32.1o and 37.2o (JCPDS File No. 00-037-1497) could be associated to the CaO

an
299

300 phase. The main peaks at 2θ values of 10.8o was attributed to the Ca(OCH3)2 catalyst (Teo et

al., 2015). In addition, three peaks at 2θ of 17.8o, 28.6o and 34.0o was corresponds to the
M
301

302 existence of Ca(OH)2 (JCPDS file No: 01-84-1264) which could be formed due to the
d

303 interaction of the catalyst with the water molecules.


te

304 The isotherms of catalyst, as shown in Fig. 3(a) resembles with Type IV according to

305 the IUPAC classification with H3 type hysteresis loop (Sing et al. 1985). The surface area,
p

306 average pore size and the pore volume of the catalyst (Fig. 3(b)) were 30 m2g-1, 32.97 nm and
ce

307 0.21 cm3g-1, respectively (Table 2). The surface of Ca(OCH3)2catalyst consist of mesopores

308 ranging from 2 to 5 nm which could be related with the activity of catalyst in the
Ac

309 transesyerification reaction (Sing et al., 1985).

310 The FTIR spectrum of Ca(OCH3)2 catalyst (Fig. 5(a)) presented the important

311 functional groups at 1070 cm−1 indicating C-O stretching vibration of primary alcohol. The

312 stretching frequency peaks at 3650 cm−1 and 2800–3000 cm−1 corresponding to the CH3 and

313 C-H stretching vibration of the catalyst respectively, and are usually strong. Besides, peaks at

314 3650 cm−1, could be due to the adsorption of moisture on the surface of the catalyst.

13
Page 13 of 38
315 Moreover, -OH groups ware also existed as isolated substance on the surface of the catalyst

316 (Ilgen and Akin, 2008). No organic functional group was observed on the CaO catalyst as

317 shown in Fig. 4(b).

318 The irregular shape of CaO catalyst might be due to agglomeration of the bulk

t
319 particles shown in (Figure 5(a)). A flower-like cluster structure was observed in Fig. 5(b)

ip
320 which is similar to the result reported by Kouzu et al. (2008). A rapid mass transfer into the

cr
321 interstices of the catalyst could be provided by flower-like cluster structure as reported by

322 Kouzu et al. (2008). TEM images of CaO (Fig. 5(c)) and Ca(OCH3)2 (Fig. 5(d)) catalysts

us
323 revealed to be cubic in shape. Both images are consistent with result obtained from FESEM

an
324 study, as shown in Fig. 5(a-b).

325
326 3.3. Process optimization of transesterification
M
327 The molar ratio of methanol to algae oil was varied from 10:1 to 60:1, as can be seen in Fig. 6.

328 The FAME yield derived from crude microalgae lipid was increased significantly from 6.9-
d

329 85.4 % with methanol molar ratio of 10:1 to 60:1. As reported by several researchers (Islam
te

330 et al., 2013) that high amount of methanol could be accelerated the transesterification
p

331 reaction rate. Fig. 7 demonstrated the effect of the Ca(OCH3)2 catalyst dosage on biodiesel
ce

332 yield. The concentration of catalyst during the reaction was varied from 0-15 wt. % for 3 h at

333 60 ○C. The maximum FAME yield of 92.0 % was obtained at 12 wt. % of catalyst dosage.
Ac

334 However, the yield was started to drop to 61.6 % with further increasing the catalyst loading

335 to 15 wt. %. This could be attributed to the poor diffusion between the methanol–oil–catalyst

336 systems when catalyst overloading in the reaction system. The high activity of the catalyst

337 could be due to the formation of active CH3O- species dissociated for methanol attached to

338 the triglyceride molecule under the optimum catalyst concentrations (Islam et al., 2013a;

339 Islam et al., 2013b).

14
Page 14 of 38
340 As can be seen that the triglyceride (TG) was converted into biodiesel by 20-30 min

341 of reaction time (Fig. 8). The results indicated that the reaction rate obtained from the

342 homogeneous NaOH catalyst was significantly faster reaction compared the heterogeneous

343 catalysts. On the other hand, the reaction rate of heterogeneous catalyst system (Ca(OCH3)2

t
344 and CaO) is very slow and it achieved high FAME yield after the reaction time of 3h.

ip
345 However, the conversion rate of 92 % FAME yield was observed for Ca(OCH3)2 catalyst

cr
346 compared to conversion 80% over the CaO after the same duration. In this case,

347 homogeneous system (NaOH) shows fast conversion rate and reach maximum yield at a short

us
348 reaction time, however, the formation of other intermediates and saponification are the main

an
349 drawbacks of the liquid catalyst (Nautiyal et al., 2014). The influence of reaction

350 temperature is very important for heterogeneously catalyzed. This is because high reaction
M
351 temperature requires more energy use in the process. Fig. 9 present the variation of

352 temperature on crude microalgae oil transesterification was studied at 3 wt.% catalyst loading
d

353 and methanol/ oil molar ratio of 30:1. The reaction was performed at four different
te

354 temperatures of 50, 60, 70 and 80 °C. As can be seen from Fig. 9 that the yield was reached

355 to 52 % at 50 °C after 180 min of reaction time whereas the yield increased to 92, 97 and 99
p

356 % at 60, 70 and 80 °C respectively after the same reaction time. Transesterification of oil
ce

357 with methanol in the presence of heterogeneous catalyst is a three-phase reaction system.

358 Therefore, raising the reaction temperature favoured the transesterification due to the
Ac

359 enhancement of miscibility at high temperature.

360
361 3.4. Kinetic of transesterification

362 Depending on the systems, pseudo-second-order (Birla et al., 2012), reversible

363 second-order (Hassan and Vinjamur, 2014), or pseudo-first-order (Nautiyal et al., 2014),

364 kinetic models have been reported for biodiesel production. In most cases, pseudo-first-order

365 kinetics has been observed in the presence of excess methanol (Sivasamy et al., 2009). As can

15
Page 15 of 38
366 be seen from Fig. 10, the fitted straight lines are not passing through origin for the initial

367 reaction kinetics, which means that the reaction obeys irreversible second-order kinetic

368 during the initial period (Hassan and Vinjamur, 2014). This irreversible kinetics in the initial

369 period is in agreement with sulfuric acid-catalyzed esterification of fatty acids with methanol

t
370 reported by Boucher et al. (2008). A similar conclusion has been reached by Liu et al. (2006)

ip
371 by comparing both the activity-based and concentration-based approaches in their studies.

cr
372 The values of rate constant calculated from Eq. (13) for both regimes obtained at

373 reaction temperatures of 50, 60, 70 and 80 °C, which are 0.003, 0.011, 0.014 and 0.019 min-1,

us
374 respectively. It has been shown by the researchers that the rate constant varies by the reaction

an
375 parameter i.e. fatty acid composition oil and process used for transesterification. The values

376 of reaction rate constant from Spirulina platensis algae biomass were reported to be

0.001 min−1 under optimized operating conditions (Nautiyal et al., 2014). Vyas et al.
M
377

378 (2009) reported the rate constant values varied from 0.0031-0.0363 min−1 for biodiesel
d

379 production through transesterification of Jatropha oil at temperatures ranges from 50-70 oC.
te

380 The rate constant of the transesterification reaction varies from 0.001 to 0.006 min−1, as

381 reported by several researchers (Rathore and Madras, 2007). Some authors (Kaur and Ali,
p

382 2014a) reported the higher rate constant (0.03 min−1) for the ethanolysis of waste cottonseed
ce

383 oil over lithium impregnated calcium oxide. Thus, the value of rate constant obtained in this

384 study was consistent with the values reported in the literature.
Ac

385 Furthermore, the Arrhenius energy of activation from a plot of the reaction rate

386 constant (k) vs. the reciprocal of absolute temperature (T) was determined from values of rate

387 constants at different temperatures (Fig. 11). The slope and intercept of the graph between

388 lnk vs. 1/T x 103 gives the values of activation energy and frequency factor. The observed Ea

389 value in present study (58.62 kJ mol-1) was found within the range of reported values (41-66

390 kJ mol-1) for transesterification reaction catalyzed by heterogeneous catalysts (Kaur and Ali,

16
Page 16 of 38
391 2014b; Kaur and Ali, 2015; Kaur and Ali, 2014c). Zhang et al. (2010) reported higher

392 activation energy for the transesterification catalyzed of waste frying oil using potassium

393 hydroxide (79.1 kJ mol-1) and snail shell derived catalysts (79 kJ mol-1), respectively. This

394 implies that, on increasing the temperature, the side reactions leading to soap formation are

t
395 more favored than the ones producing biodiesel (Eze et al. 2014).

ip
396 A conclusion has been reached by Nautiyal et al., (2014) that the higher amount of

cr
397 unsaturated fatty acid (52% of linoleic acid) presence in the waste cooking oil may increase

398 the value activation energy. Further, it has been shown by other workers (Lee et al., 2010)

us
399 that the activation energy for particular reactions is influenced by the concentration of

an
400 catalyst, reaction temperature, nature of reactant etc. Thus, the reason for the lower value of

401 activation energy as compared to values obtained for waste frying oil could be due to the
M
402 catalyst used in this study.

403
404 3.4. Durability of Catalyst
d

405 In order to evaluate the catalyst sustainable, the used Ca(OCH3)2 catalyst was washed
te

406 thoroughly with methanol and hexane, then, dried in an oven for overnight after each
p

407 experiment (Fig. 12). The catalyst was reused for transesterifying operation up to 5 times and
ce

408 sustained the activity to produce FAME yield at range of 92.7 – 96.0 %. However, the yield

409 of FAME was reduced to 67.2 %. After 5th runs of transesterification reaction, the well
Ac

410 defined flower-like cluster crystallites structure of Ca(OCH3)2 catalyst (Fig. 5(b)) was

411 reorganized in aggregates in a cakes-like sticky structure of Ca(OH)2 (Fig 13(a)). This

412 rephrase in the catalyst bulk morphology led to a deactivation of the catalyst by blocking the

413 contact between the active catalytic sites and the reactants. These results are consistent with

414 the obtained XRD profile on the existing Ca(OH)2 crystal phases (Fig. 13(b)). A conclusion

415 has been also reached by Teo et al. (2015) that the catalyst might be deactivated by the

416 impurities e.g., phospholipids and moisture presence in the reaction medium.

17
Page 17 of 38
417 4. Conclusions

418 The present study showed that nanoCa(OCH3)2 catalyst was effective solid catalyst for

419 biodiesel production rom Nannochloropsis microalgae. It is inexpensive and environment

420 friendly, has high catalytic activity. The high activity of the catalyst was described to the

t
421 methoxide species (CH3O-) that allowed more methanol to contact with the carbonyl group of

ip
422 triglyceride. The high specific surface area and large pore size are favorable for contact

cr
423 between catalyst and substrates, which effectively improved efficiency of transesterification.

424 The maximum conversion was achieved as 99.0% at 80 °C, 30 : 1 molar ratio of methanol to

us
425 oil and 3 wt.% of catalyst. This result indicates that the new route of synthesis has provided a

an
426 sustainable way with significant benefits for an environmentally benign procedure in produce

427 biodiesel.
M
428
429 References
d

430 Alcantara, R., Amores, J., Canoira, L. T., Fidalgo, E., Franco, M. J., Navarro, A., 2000.
431 Catalytic production of biodiesel from soy-bean oil used frying oil and tallow. Biomass
te

432 Bioenerg. 2000, 18(6), 515-527.

433 Andrikopoulas, N. K., Giannakis, I. G., Tzamtzis, V., Analysis of olive oil and seed oil
434 triglycerides by capillary gas chromatography as a tool for the detection of the
p

435 adulteration of olive oil. J. Chromatogr. Sci. 2001, 39, 137–145.


ce

436 Arifin, A. A., Bakar, J., Tan, C. P., Rahman, R. A., Krim R., Loi, C. C., 2009. Differential
437 scanning calorimetric analysis of edible oils: comparison of thermal properties
438 and chemical composition. Food Chem. 114, 561-564.
Ac

439 Asakuma, Y., Maeda, K., Kuramochi, H., Fukui, K., 2009. Theoretical study of the
440 transesterification of triglycerides to biodiesel fuel. Fuel 88(5), 786-791.

441 Birla, A., Singh, B., Upadhyay S. N., Sharma, Y. C., 2012. Kinetics studies of synthesis of
442 biodiesel from waste frying oil using a heterogeneous catalyst derived from snail
443 Bioresour technol. 106, 95-100.

444 Boucher, M. B., Unker, S. A., Hawley, K. R., Wilhite, B. A., Stuart, J. D., Parnas, R. S.,
445 2008. Variables affecting homogeneous acid catalyst recoverability and reuse after
446 esterification of concentrated omega-9 polyunsaturated fatty acids in vegetable oil
447 triglycerides. Green Chem.10, 1331-1336.

18
Page 18 of 38
448 Di Serio, D., Tesser, R., Pengmei L., Santacesaria, E., 2007. Heterogeneous catalysts for
449 biodiesel production. Energ. Fuel 22, 207-217.

450 Eze V. C., Phan, A. N., Harvey, A.P. 2014. A more robust model of the biodiesel reaction,
451 allowing identification of process conditions for significantly enhanced rate and water
452 tolerance. Bioresour. Technol., 156: 222-231.

453 Hassan, S. Z., Vinjamur, M., 2014. Concentration-independent rate constant for biodiesel

t
454 synthesis from homogeneous-catalytic esterification of free fatty acid. Chem. Eng. Sci.

ip
455 107, 290-301.

456 Ilgen, O., Akin, A. N., 2009. Transesterification of canola oil to biodiesel using MgO loaded

cr
457 with KOH as a heterogeneous catalyst. Energ. Fuels 23, 1786-1789.

458 Islam, A., Chan, E. S., Taufiq-Yap, Y. H., Mondal, M. A. H., Moniruzzaman, M., Mridha, M.,

us
459 2014. Energy security in Bangladesh perspective—An assessment and
460 implication. Renew. Sus. Energ. Rev. 32, 154-171.

461 Islam, A., Taufiq-Yap, Y. H., Chu, C. M., Chan E. S., Ravindra, P., 2013b. Studies on design

an
462 of heterogeneous catalysts for biodiesel production. Process Saf. Environ. 91, 131-144.

463 Islam, A., Taufiq-Yap, Y. H., Chu, C. M., Ravindra P., Chan, E. S., 2013a.
464 Transesterification of palm oil using KF and NaNO 3 catalysts supported on spherical
M
465 millimetric γ-Al2O3. Renew. Energ. 59, 23-29.

466 Kaur, M., Ali, A., 2014a. Potassium fluoride impregnated CaO/NiO: An efficient
467 heterogeneous catalyst for transesterification of waste cottonseed oil. Eur. J. Lipid Sci.
d

468 Technol. 116(1), 80-88.


te

469 Kaur, N., Ali, A., 2014b. Kinetics and reusability of Zr/CaO as heterogeneous catalyst for the
470 ethanolysis and methanolysis of Jatropha crucas oil. Fuel Process. Technol. 119, 173-
471 184.
p

472 Kaur, M., Ali, A. 2014c. Ethanolysis of waste cottonseed oil over lithium impregnated
ce

473 calcium oxide: Kinetics and reusability studies. Renew. Energ. 63, 272-279.

474 Kaur, N., Ali, A. 2015. Biodiesel production via ethanolysis of jatropha oil using
475 molybdenum impregnated calcium oxide as solid catalyst. RSC Advances 5(18),
Ac

476 13285-13295.

477 Kouzu, M., Kasuno, T., Tajika, M., Yamanaka, S., Hidaka, J., 2008. Active phase of calcium
478 oxide used as solid base catalyst for transesterification of soybean oil with refluxing
479 methanol. Appl. Catal. A: Gen. 334, 357-365.

480 Lam M. K., Lee, K. T., 2012. Potential of using organic fertilizer to cultivate Chlorella
481 vulgaris for biodiesel production. Appl. Energ. 94, 303-308.

482 Lam, M. K., Lee K. T., Mohamad, A. R., 2012. Microalgae biofuels: A critical review of
483 issues, problems and the way forward. Biotechnol. Adv. 30(3), 673-690.

484 Lee, J. Y., Yoo, C., Jun, S. Y., Ahn, C. Y., Oh, H.M., 2010. Comparison of several methods
485 for effective lipid extraction from microalgae. Bioresour Technol. 101, 75-77.

19
Page 19 of 38
486 Liu, Y., Lotero E., Goodwin Jr, J. G., 2006. Effect of water on sulfuric acid catalyzed
487 esterification. J. Mol. Catal. A: Chem. 245, 132-140.

488 Nautiyal, P., Subramanian, K. A.,. Dastidar, M. G., 2014. Kinetic and thermodynamic studies
489 on biodiesel production from Spirulina platensis algae biomass using single stage
490 extraction–transesterification process. Fuel 135, 228-234.

491 Palligarnai T. V., Michael, B., 2008. Biodiesel production—current state of the art and

t
492 challenges. J. Indian Microbiol. Biotechnol. 35, 421-430.

ip
493 Rathore, V., Madras, G., 2007. Synthesis of biodiesel from edible and non-edible oils in
494 supercritical alcohols and enzymatic synthesis in supercritical carbon dioxide. Fuel

cr
495 86(17), 2650-2659.

496 Sarin, R., Sharma, M., Sinharay, S., Malhotra, R. K., 2007. Jatropha- palm biodiesel blends:

us
497 an optimum mix for Asia. Fuel 86(10), 1365-1371.

498 Shahbazi, M. R., Khoshandam, B., Nasiri, M., Ghazvini, M., 2012. Biodiesel production via
499 alkali-catalyzed transesterification of Malaysian RBD palm oil–Characterization,

an
500 kinetics model. J. Taiwan Inst. Chem. Eng. 43(4), 504-510.

501 Sing, K. S. W., Everett, D. H., Haul, R. A. W., Moscou, L., Pierotti R. A., Rouquerol, J.,
502 1985. Reporting physisorption data for gas/solid systems with special reference to the
M
503 determination of surface area and porosity. Pure Appl. Chem. 57, 603-619.

504 Sivasamy, A., Cheah, K. Y., Fornasiero, P., Kemausuor, F., Zinoviev, S., Miertus, S., 2009.
505 Catalytic applications in the production of biodiesel from vegetable oils. Chem. Sust.
d

506 Chem.2(4), 278-300.


te

507 Teo, S. H., Islam, A., Ng, F. L., Taufiq-Yap, Y. H., 2015. Biodiesel synthesis from
508 photoautotrophic cultivated oleoginous microalgae using a sand dollar catalyst. RSC
509 Adv. 5(6), 4266-4276.
p

510 Uzun, B. B., Kılıç, M., Özbay, N., Pütün A. E., Pütün, E., 2012. Biodiesel production from
ce

511 waste frying oils: Optimization of reaction parameters and determination of fuel
512 properties Energ. 44(1), 347-351.

513 Vyas, A. P., Subrahmanyam N., Patel, P. A., 2009. Production of biodiesel through
Ac

514 transesterification of Jatropha oil using KNO3/Al2O3 solid catalyst. Fuel 88(4), 625-628.

515 Watkins, R. S., Lee A. F. Wilson, K., 2004. Li-CaO catalysed tri-glyceride transesterification
516 for biodiesel applications Green Chem. 6, 335-340.

517 Xie W. L., Fan M. L., 2014. Biodiesel production by transesterification using
518 tetraalkylammonium hydroxides immobilized onto SBA-15 as a solid catalyst. Chem.
519 Eng. J. 239, 60-67.

520 Xie, W., Wang, H., Li, H., 2011. Silica-supported tin oxides as heterogeneous acid catalysts
521 for transesterification of soybean oil with methanol. Ind. Eng. Chem. Res. 51(1), 225-
522 231.

20
Page 20 of 38
523 Xie, W., Zhao, L. L., 2013. Production of biodiesel by transesterification of soybean oil using
524 calcium supported tin oxides as heterogeneous catalysts. Energ. Convers. Manage. 76,
525 55-62.

526 Xie W. L., Zhao L. L., 2014. Heterogeneous CaO-MoO3-SBA-15 catalysts for biodiesel
527 production from soybean oil. Energ. Convers. Manage. 79, 34-42.

528 Zhang, L., Sheng, B., Xin, Z., Liu Q., Sun, S., 2010. Kinetics of transesterification of palm

t
529 oil and dimethyl carbonate for biodiesel production at the catalysis of heterogeneous

ip
530 base catalyst. Bioresour. Technol. 101, 8144-8150.
531

cr
us
an
M
d
p te
ce
Ac

21
Page 21 of 38
531 Table 1
532 Fatty acid content (a), moisture, free fatty acid and dried ash contents (b) of crude N. oculata
533 microalgae based methyl esters.
(a)
Substituent Fatty acid Common C Molecular Composition in Molecular mass
name mass (g mol- sample (%) contribution (g
1
) mol-1)
Decanoic acid Lauric acid 12:0 200.31 - -
Tetradecanoic acid Myristic acid 14:0 228.37 7.69 ± 0.52 17.572
Hexadecanoic acid Palmitic acid 16:0 256.42 35.43 ± 0.06 90.862

t
Saturated lipid
Octadecanoic acid Stearic acid 18:0 284.48 2.50 ± 0.26 7.115

ip
Eicosanoic acid Arachidic 20:0 312.54 - -
acid
Hexadecenoic acid Palmitoleic 16:1 254.41 27.54 ± 0.15 70.055
acid

cr
cis-9-octadecenoic Oleic acid 18:1 282.46 8.62 ± 0.02 24.335
Monounsaturated lipid
acid
trans-9-octadecenoic Elaidic acid 18:1 282.46 - -
acid

us
all cis-9,12- Linoleic acid 18:2 280.45 5.22 ± 0.23 14.643
octadecadienoic acid
all cis-5,8,11,14- Arachidonic 20:4 304.47 2.47± 0.05 7.529
eicosatetraenic acid acid
Polyunsaturated lipid
all cis-5,8,11,14,17- EPA 20:5 302.45 8.29 ± 0.08 25.058
eicosatetraenic acid

an
all cis-4,7,10,13,16,19- DHA 22:6 328.49 2.24 ± 0.16 7.355
docosahexenoic acid
Average molecular weight of 264.525
constituent fatty acid
(AMWfatty acid)
Unsaturated/ saturated lipid 1.19
M
ratio
(b)
Substituent Content (%)
Moiture 5.47 ± 0.04
d

Free fatty acid 8.03 ± 0.31


Dried ash 10.68 ± 0.32
te

534
535
536
p

537
538
ce

539
540
541
542
Ac

543
544
545
546
547
548
549
550
551
552
553
554
555

22
Page 22 of 38
556
557
558 Table 2
559 BET surface area, total pore volume and average pore diameter of Ca(OCH3)2 and CaO
560 catalysts.
a
Catalyst BET (m2 g-1) Total pore volume (cm3 g-1) Average pore diameter (nm)
Ca(OCH3)2 30.5 0.21 31.97
CaO 6.3 0.04 3.12

t
561

ip
562
563
564

cr
565
566
567

us
568
569
570

an
571
572
573
574
M
575
576
577
578
d

579
580
te

581
582
583
p

584
585
ce

586
587
588
589
Ac

590
591
592
593
594
595
596
597
598
599
600

23
Page 23 of 38
600 Figure of Captions:
601
602 Fig. 1 TG/DTA spectrum of Ca(OCH3)2 catalyst.
603
604 Fig. 2 X-ray diffraction patterns of Ca(OCH3)2 (a) and CaO (b) catalysts.∆, characteristic

t
605 peak of Ca(OCH3)2; ♦, characteristic peak of CaO; X, characteristic peak of Ca(OH)2.

ip
606
607 Fig. 3 Nitrogen adsorption-desorption isotherms (a) and pore size distribution (b) of

cr
608 Ca(OCH3)2 catalyst.
609

us
610 Fig. 4. FTIR spectra of Ca(OCH3)2 (a) and CaO (b) catalysts.
611

an
612 Fig. 5 SEM (a & b) and TEM (c & d) micrographs of CaO and Ca(OCH3)2 catalysts.
613
614 Fig. 6 Influence of methanol/oil molar ratio on the FAME yield of crude microalgae lipid by
M
615 Ca(OCH3)2 catalyst. Reaction condition: catalyst dosage = 3 wt. %, reaction time = 3 h,
616 reaction temperature = 60 °C.
617
d

618 Fig. 7 Influence of Ca(OCH3)2 catalyst loading on the FAME yield of crude microalgae lipid
te

619 by Ca(OCH3)2 catalyst. Reaction condition: n(methanol):n(lipid) = 30:1, reaction time = 3 h,


620 reaction temperature = 60 °C.
p

621
ce

622 Fig. 8 Influence of reaction time on the FAME yield of crude microalgae lipid by Ca(OCH3)2
623 (a), CaO (b) and NaOH (c) catalyst. Reaction condition: catalyst dosage = 12wt. % (a & b)
624 and 1 mol (c), n(methanol):n(oil) = 30:1, reaction temperature = 60 °C.
Ac

625
626 Fig. 9 Influence of reaction temperature on the FAME yield of crude microalgae lipid by
627 Ca(OCH3)2 catalyst. Reaction condition: n(methanol):n(lipid) = 30:1, reaction time = 0.5-4 h,
628 catalyst loading = 3 wt. %.
629
630 Fig. 10. Arrhenius plot ln k vs. 1/T x 103for transesterification reactionin the presence of
631 Ca(OCH3)2 catalyst.Reaction condition: n(methanol):n(lipid) = 30:1, reaction time = 0.5-4 h,
632 catalyst loading = 3 wt. %.
633
24
Page 24 of 38
634 Fig.11 -ln(1-XA) vs. time plot for transesterification reaction in the presence of
635 Ca(OCH3)2catalyst.Reaction condition: n(methanol):n(lipid) = 30:1, reaction time = 0.5-4 h,
636 catalyst loading = 3 wt. %.
637
638 Fig. 12 Stability and reusability test.

t
639

ip
640 Fig. 13 SEM micrograph (a) and XRD pattern (b) of deactivated Ca(OCH3)2 catalyst.
641

cr
642
643

us
644
645

an
646
647
648
M
649
650
651
d

652
te

653
654
p

655
ce

656
657
658
Ac

659
660
661
662
663

664

665

25
Page 25 of 38
100 2
o
430 C

95

90
o
720 C

DTA (uV/mg)

t
TGA (%)

ip
85 -2

cr
80

-4

us
75

70 -6
100 200 300 400 500 600 700 800 900 1000

an
o
Temperature ( C)
666
667 Fig. 1
M
668
669
670
d

671
te

672
673
p

674
ce

675
676
677
Ac

678
679
680
681
682
683
684
685
686

26
Page 26 of 38


(a)  

t
Intensity (a.u.)

ip
 Ca(OCH3)2

 CaO (JCPDS file: 00-037-1497)


cr
 Ca(OH)2 (JCPDS file: 1-84-1264)

(b)

us
an
5 10 15 20 25 30 35 40

2 (Degree)
687
M
688 Fig. 2
689
690
d

691
te

692
693
p

694
695
ce

696
697
Ac

698
699
700
701
702
703
704
705
706

27
Page 27 of 38
300

(a)

200
3 -1
Vads/cmg

t
ip
100

cr
0
0.0 0.2 0.4 0.6 0.8 1.0

707

us
Relative pressure (p/p0 )

708

(b)

an
M
d
te

709
p

710 Fig. 3
ce

711
712
713
Ac

714
715
716
717
718
719
720
721
722

28
Page 28 of 38
(b)

(a)
Transmittence (%)

t
ip
cr
us
723

an
Wavelength (cm-1)
724
725 Fig. 4
M
726
a)
727
728
d

729
te

730
731
p

732
733
ce

734
735
Ac

736
737
738
739
740
741
742
743
744

29
Page 29 of 38
a)
b)

t
ip
cr
745

us
c) d)

an
M
d
te

746
p

747 Fig. 5
ce

748
749
750
Ac

751
752
753
754
755

30
Page 30 of 38
t
ip
cr
us
756
757
Fig. 6

an
M
758
759
760
d

761
te

762
763
p

764
ce

765
766
767
Ac

768
769
770
771
772
773
774

31
Page 31 of 38
t
ip
cr
us
775

an
776 Fig. 7
777
778
M
779
780
781
d

782
te

783
784
p

785
ce

786
787
788
Ac

789
790
791
792
793
794
795
796

32
Page 32 of 38
t
ip
cr
us
797
CaO NaOH
anCa(OCH3)2
M
798 Fig. 8
799
800
d

801
te

802
803
p

804
805
ce

806
807
Ac

808
809
810
811
812
813

33
Page 33 of 38
t
ip

50 C

cr
60 ○C

70 C

us

80 C

814
815 Fig. 9
an
M
816
817
818
d

819
te

820
821
p

822
ce

823
824
825
Ac

826
827
828
829
830
831
832
833
834

34
Page 34 of 38

60 C

70 C

80 C

90 C

t
ip
cr
us
an
835
836 Fig. 10
837
M
838
839
840
d

841
te

842
843
p

844
ce

845
846
847
Ac

848
849
850
851
852
853
854
855

35
Page 35 of 38
t
ip
cr
us
an
856
857 Fig.11

858
M
859
860
d

861
862
te

863
864
p

865
ce

866
867
Ac

868
869
870
871
872
873
874
875
876

36
Page 36 of 38
t
ip
cr
us
877

an
878 Fig. 12

879
M
880
881
d

882
883
te

884
885
p

886
ce

887
888
Ac

889
890
891
892
893
894
895
896
897

37
Page 37 of 38
(a)
898
899
900
901
902

t
903

ip
904
905

cr
906
907

us
908 (b)
909

an
910
Ca(OH)2 (JCPDS file: 1-84-1264)
911 
M
Intensity (a.u.)

912

913 
d

914 
te

915

916
p

917 5 10 15 20 25 30 35 40
ce

2 (Degree)
918 Fig. 13

919
Ac

920
921
922
923
924
925
926
927
928

38
Page 38 of 38

You might also like