Subscriber access provided by University of Sydney Library
Article
In-situ transesterification of Chlorella vulgaris using carbon-dot functionalized strontium oxide as a heterogeneous catalyst under microwave irradiation Alex Tangy, Vijay Bhooshan Kumar, Indra Neel Pulidindi, Yael Kinel-Tahan, Yaron Yehoshua, and Aharon Gedanken Energy Fuels, Just Accepted Manuscript • DOI: 10.1021/acs.energyfuels.6b02519 • Publication Date (Web): 22 Nov 2016 Downloaded from http://pubs.acs.org on November 22, 2016
Just Accepted “Just Accepted” manuscripts have been peer-reviewed and accepted for publication. They are posted online prior to technical editing, formatting for publication and author proofing. The American Chemical Society provides “Just Accepted” as a free service to the research community to expedite the dissemination of scientific material as soon as possible after acceptance. “Just Accepted” manuscripts appear in full in PDF format accompanied by an HTML abstract. “Just Accepted” manuscripts have been fully peer reviewed, but should not be considered the official version of record. They are accessible to all readers and citable by the Digital Object Identifier (DOI®). “Just Accepted” is an optional service offered to authors. Therefore, the “Just Accepted” Web site may not include all articles that will be published in the journal. After a manuscript is technically edited and formatted, it will be removed from the “Just Accepted” Web site and published as an ASAP article. Note that technical editing may introduce minor changes to the manuscript text and/or graphics which could affect content, and all legal disclaimers and ethical guidelines that apply to the journal pertain. ACS cannot be held responsible for errors or consequences arising from the use of information contained in these “Just Accepted” manuscripts.
Energy & Fuels is published by the American Chemical Society. 1155 Sixteenth Street N.W., Washington, DC 20036 Published by American Chemical Society. Copyright © American Chemical Society. However, no copyright claim is made to original U.S. Government works, or works produced by employees of any Commonwealth realm Crown government in the course of their duties.
Page 1 of 31
1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
Energy & Fuels
1
In-situ transesterification of Chlorella vulgaris
2
using carbon-dot functionalized strontium
3
oxide as a heterogeneous catalyst under
4
microwave irradiation
5 6
Alex Tangy,#† Vijay Bhooshan Kumar,#† Indra Neel Pulidindi,† Yael Kinel-Tahan,‡
7
Yaron Yehoshua‡ and Aharon Gedanken†*
8
† Department of Chemistry, Bar-Ilan University, Ramat-Gan 52900, Israel
9
‡ The Mina and Everard Goodman Faculty of Life Sciences, Bar-Ilan University,
10
Ramat-Gan 5290002, Israel
11 12 13 14
15
16
17
18
19
20
21
1 ACS Paragon Plus Environment
Energy & Fuels
1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
22
ABSTRACT:
23
The main goal of this study is to functionalize SrO with carbon dots (C-dots) and to
24
explore the composite as a catalyst for fatty acid methyl esters (FAME) production
25
using Chlorella vulgaris as feedstock. C-dots are synthesized by sonicating
26
polyethylene glycol followed by sonochemical modification of Sr(NO3)2 (precursor
27
for SrO) with C-dots. Sonication facilitates the adhesion of C-dots to the surface of
28
Sr(NO3)2. The resulting material is calcined in an inert environment to form SrO-C-
29
dot composite. The effect of functionalizing SrO with C-dots on the transesterification
30
of the lipids in the alga with methanol is studied. The optimization of a one-stage
31
process of conversion of the lipid fraction of microalga Chlorella vulgaris into FAME
32
using direct transesterification under microwave irradiation is illustrated. A lipid
33
conversion value of 45.5 wt. % is achieved using the SrO-C-dot catalyst after 2.5
34
minutes of microwave (MW) irradiation. The catalyst displayed better activity than
35
commercial SrO. Microwave irradiation accelerates the disruption of the microalgal
36
cells and facilitates the release of lipid content into the reaction medium. The catalyst
37
is characterized by a variety of physico-chemical techniques. The FAME product
38
obtained from the alga is quantified using 1H NMR spectroscopy. The new catalyst,
39
namely, SrO-C-dot nanoparticles (NPs), yielded 97 wt.% FAME from Chlorella
40
vulgaris in 2.5 minutes of MW irradiation.
41
KEYWORDS: Microalga, Chlorella vulgaris, Lipid, FAME, SrO-C dot
42
43
44
2 ACS Paragon Plus Environment
Page 2 of 31
Page 3 of 31
1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
Energy & Fuels
45
1. INTRODUCTION
46
Depletion of fossil fuels, environmental concerns, and growing energy requirements
47
prompt the replacement of at least part of the petroleum-based fuels used for
48
transportation by renewable energy sources. The transportation sector, representing 30
49
% of the world’s total energy consumption, is the largest energy-consuming sector
50
after the industrial sector. Almost 60 % of world’s oil demand is for the transportation
51
sector, which will probably be the most energy-demanding sector in the future.1
52
Biodiesel produced from renewable biological sources has gained importance as an
53
alternative energy source to fossil-based transportation fuels such as petrol and
54
diesel.2,3 Biodiesel is generally produced using vegetable oils as feedstock. The major
55
drawback of using edible biomass as feedstock is that arable land is used to farm the
56
oil seed crops, and this can lead to a food-Vs-fuel-Vs-land conflict. Moreover, it has
57
been reported that a number of feedstock like algal oils, jatropha and waste cooking
58
oil, can reduce 60-90 % of the biodiesel production cost.4
59
Microalgae are photosynthetic microorganisms capable of converting sunlight,
60
atmospheric CO2 and wastewater into a variety of high-energy molecules, including
61
fatty acids (FA) and triacylglycerols (TAGs), the major feedstock used for biodiesel
62
or FAME production.5,6 Microalgae, also called third-generation fuel, are considered
63
one of the most promising feedstock for biofuels.7 The productivity of these
64
photosynthetic microorganisms in converting CO2 into carbon-rich lipids greatly
65
exceeds that of agricultural oleaginous crops, without competing for arable land and
66
food crops.8 Microalgae can be grown on fresh or salt water, and even in a wastewater
67
environment.6 Recent advances present opportunities to develop microalgae as
68
feedstock for biodiesel production in a sustainable and economical way within the
69
next 10 to 15 years.9,10 Microalgae can be cultivated from several aqueous systems 3 ACS Paragon Plus Environment
Energy & Fuels
1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
70
such as ponds and photo-bioreactors. Microalgae have the highest growth rates among
71
all the photosynthetic organisms.11 Approximately 46 tons of oil/hectare/year can be
72
produced from diatom algae.12 These characteristics make them an attractive
73
feedstock for biodiesel production.
74
Algal biomass contains macromolecular proteins, polysaccharides and lipids,
75
along with inorganic components.13 Only the lipid content is generally targeted for
76
FAME production. FAME production from microalgae can be divided into three main
77
stages: a) cultivation, b) harvesting, and c) lipid extraction and transesterification for
78
FAME production. The lipid content of microalgae is dependent on the microalgae
79
species and cultivation conditions. Microalgae can accumulate approximately between
80
20 and 70% of lipid content from the dry biomass.14 The usual method for extracting
81
lipids from the whole algae biomass consists of mixing the microalgae with an
82
organic solvent such as chloroform or hexane as co-solvent, or adopting other
83
methods such as using a functional membrane coated with a cationic polymer15 or
84
using ionic liquids.16 The extracted micro algal lipids are then converted to FAME by
85
transesterification using alcohol and a catalyst. Recently, a combination of lipid
86
extraction and FAME conversion in a one-step process, called direct (in-situ)
87
transesterification, has been reported.17 This methodology simplifies the production
88
process and improves the FAME yield compared to conventional lipid extraction step
89
that involves lipid loss. Moreover, the number of stages in the biodiesel production
90
process are also reduced in the in-situ extraction based processes. Direct
91
transesterification has several advantages, such as cell disruption, lipid extraction and
92
transesterification in a single step, and has been proved to be more useful than
93
conventional methods.18 The conventional two-step preparation of biodiesel from
94
microalgae has several disadvantages compared to in situ transesterification, including 4 ACS Paragon Plus Environment
Page 4 of 31
Page 5 of 31
1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
Energy & Fuels
95
operation complexity, a longer process time, high energy consumption and relatively
96
higher cost, which restricts the commercial production of biodiesel from microalgae.
97
Higher FAME yield is obtained by the direct transesterification than in a two-stage
98
process.17,20 Moreover, direct transesterification provides energy-efficient and
99
economical routes for biodiesel production.20
100
Despite the high conversion yields achieved using homogeneous catalysts such
101
as NaOH, KOH, H2SO4, etc., catalysts may remain in the FAME phase and, therefore,
102
a FAME refining or washing step is needed. In this sense, the use of heterogeneous
103
catalysts is a suitable alternative, allowing easy separation of the final product and
104
reusability. Metal-oxide materials play an important role in the catalytic reaction
105
because of their obvious advantages, which include easy processing, low cost and
106
easy modification into composites. Chang et al. have reviewed different types of
107
catalytic systems mediated by solid mixed metal-oxides for biodiesel production.21
108
Among the metal-oxide catalysts, alkaline earth metal oxides possess excellent
109
basicity and exhibit significant activity and selectivity in the transesterification
110
reaction.22 Specifically, the use of SrO, despite its lower surface area23 and the partial
111
solubility of the metal ion in the reaction medium,24 results in almost complete
112
conversion of waste cooking oil into FAME in 10 s under microwave irradiation.51,52
113
Moreover, all the lipids in dry Nannochloropsis microalgae are fully converted to
114
FAME via direct transesterification of the microalgae in 5 min of microwave
115
irradiation.17
116
The existence of both microalgae, catalyst dispersed in methanol and FAME at
117
the end of the transesterification reaction when a heterogeneous catalyst is used, leads
118
to diffusion constraints of various components present in the medium relative to the
119
case when a homogeneous catalyst is used. This leads to lowering of the reaction rate. 5 ACS Paragon Plus Environment
Energy & Fuels
1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
120
The reaction takes place in the interfacial region between the catalyst, lipids, and
121
methanol. Thus, either a highly active catalyst or vigorous mixing is required to
122
increase the area of interaction/contact between the two phases.
123
The use of metal oxide as a host matrix for C-dots has opened a new
124
technological gateway for the development of catalysts due to the novel inherent
125
characteristics of carbon nanostructures as well as nano-metal oxides. It is expected
126
that replacing the commercial micron-sized catalytic particles by nanosized particles
127
of SrO and C-dots can lead to further improvement in the transesterification rate of
128
the conversion of algal biomass. C-dots are a nanostructured material with a
129
substantial fraction of carbon, oxygen and hydrogen. Carbon dots are a kind of quasi-
130
zero-dimensional nanomaterial. The three dimensions are all 10 nm or less in size, so
131
that the mobility of the internal electrons is restricted within the nanoscale dimensions
132
in all directions.27 C-dots may be a promising catalytic material because of their high
133
chemical stability and multifunctionality.28
134
The efficiency of FAME production from microalgae can be improved by the
135
use of microwaves.17 Microwave irradiation enhances the mass transfer rate between
136
immiscible phases and also diminishes the reaction time.18 Moreover, the FAME
137
production rate and yield from wet microalgae biomass obtained through the one-step
138
process using microwave irradiation are 6-fold and 1.3-fold higher than those
139
obtained through the two-step process using conventional heating.29
140
Our main objective in using microwave irradiation for the conversion of algae
141
to FAME is to reduce production costs and energy requirements while maximizing
142
lipid productivity, increasing the utilization efficiency of the biomass by making use
143
of all algal biomass components and reducing the process time. The current report
6 ACS Paragon Plus Environment
Page 6 of 31
Page 7 of 31
1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
Energy & Fuels
144
focuses on the design of an innovative catalytic process for a single-step conversion of
145
Chlorella vulgaris into FAME using microwave irradiation. The goal of our
146
investigation is to develop a new active catalytic material for FAME production,
147
namely, C-dot-modified by SrO obtained from Sr(NO3)2, and to study its physical,
148
chemical and catalytic properties. To the best of our knowledge, there are no reports
149
on the use of the SrO-C-dot composite as a catalyst for FAME production from
150
microalga. This study is a continuation of our recent work30 on the synthesis of C-dots
151
and the elucidation of their physicochemical properties.
152
2. EXPERIMENTAL SECTION
153
2.1. Materials and Methods
154
2.1. Chemical reagents
155
Polyethylene glycol-400 (PEG-400, 99.998%) and Sr(NO3)2 (≥ 99.0%) used as
156
precursor of SrO are purchased from Sigma-Aldrich, Israel. Strontium oxide (SrO)
157
(99.5%) is purchased from Alfa Aesar. Isopropyl alcohol (99.7%), acetone, methanol,
158
and chloroform are purchased from Bio Lab and are used as received.
159
2.2. Preparation of SrO-C-dots
160
Recently, we have reported the formation of ultrafine C-dots30 by ultrasonication of
161
PEG-400. The current synthesis (Figure 1) was performed as follows: Step 1) 20 mL
162
of polyethylene glycol (PEG-400) was transferred into a quartz test tube that had been
163
dipped in a water/Si-oil bath at 70 °C. The tip of an ultrasonic transducer (Sonics and
164
Materials Inc., USA, model VCX 750, frequency 20 kHz, AC voltage 230 V) was
165
dipped in the solution, about 2 cm above the bottom of the test tube. Ultrasonic
166
irradiation was applied with a sonication amplitude of 70 % for 2.5 h. These
7 ACS Paragon Plus Environment
Energy & Fuels
1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
167
conditions are found to be optimal for the formation of C-dots.30 Step 2) 10 g of
168
Sr(NO3)2 is added to the test tube and the sonication was continued for another 30 min
169
in a hot water bath at 50 °C. Subsequently, the product mixture was centrifuged. The
170
Sr(NO3)2-C-dot nanoparticles were washed with double-distilled water and acetone
171
(1:1) and then dried in vacuum at room temperature. The dried product was subjected
172
to calcination for varying periods of time to determine the optimal conditions for the
173
decomposition of Sr(NO3)2. The optimal condition was found by thermogravimetric
174
analysis (TGA) measurements in the presence of N2 gas. The calcination was
175
performed in an inert atmosphere to prevent the decomposition of the C-dots. This
176
methodology aided the in-situ deposition of C-dots on the surface of Sr(NO3)2 via
177
sonication and subsequent calcination of the material in an inert atmosphere resulted
178
in the formation of SrO-C-dot composite. The amount of carbon dots in the reaction
179
medium is much lower than that of Sr(NO3)2 (10 g). In the calcined product, the
180
relative amount of SrO (93 wt. %) is much higher compared to carbon dots (7 wt. %).
181
Moreover, the particle size of SrO (20-40 nm) is greater than that of the carbon dots (7
182
nm). These factors confirm the possibility that SrO is a support upon which the carbon
183
dots are functionalized in the composite material, SrO-C-dots.
184
2.3 Characterization of SrO-C-dots
185
The precise decomposition temperature of Sr(NO3)2-C-dot to SrO-C-dot was deduced
186
from the TGA. The TGA curves were recorded using TGA-GC-MS (EI/CI) Clarus
187
680/Clarus SQ 8C by Perkin Elmer over the temperature range of 25-900 °C in both
188
nitrogen and air atmospheres at a heating rate of 10 °C/min. Powder X-ray diffraction
189
(PXRD) study is conducted to probe the crystallographic nature of the product (SrO-
190
C-dot) and to verify the transformation of Sr(NO3)2 to SrO. XRD patterns were
191
collected using a Bruker AXS Advance powder X-ray diffractometer (CuKα 8 ACS Paragon Plus Environment
Page 8 of 31
Page 9 of 31
1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
Energy & Fuels
192
radiation; λ = 0.15417 nm) operating at 40 kV/30 mA with a 0.0019 step size per 0.5 s
193
over the range of 10-80° (2θ). The phases were identified using the power diffraction
194
file (PDF) database (JCPDS, International Centre for Diffraction Data). Fourier
195
Transform Infra-Red (FT-IR) spectra are recorded in KBr pallet mode on a Nicolet
196
(Impact 410) FT-IR spectrophotometer under atmospheric conditions. The samples
197
are scanned over the range of 400-4,000 cm−1. The morphology of the C-dots-SrO
198
composite and the respective elemental mapping is evaluated using high-resolution
199
scanning electron microscopy (HR-SEM) with a JEOL-JSM 700F instrument and an
200
LEO Gemini 982 field-emission gun SEM (FEG-SEM). Due to the low conductivity
201
of SrO, the sample was coated with Iridium (Ir). Elemental analysis was carried out
202
using energy-dispersive x-ray spectroscopy (EDS) in conjunction with the HR-SEM
203
instrument. Transmission electron microscope (TEM) images of the SrO-C-dot
204
particles are obtained using a Tecnai G2 (FEI, Oregon USA), a high-contrast/cryo
205
TEM equipped with a bottom CCD 1 Kx1 K camera to visualize and evaluate the
206
shape, size and surface morphology of the SrO-C-dots. Samples for the TEM images
207
were prepared by making a suspension of the particles in isopropyl alcohol, using
208
water-bath sonication. Two small droplets of the sample containing SrO-C-dot
209
particles are then applied on a TEM copper grid coated with a carbon film, and dried
210
on a covered Petri dish in vacuum. The particle-size distribution of the SrO-C-dot
211
nanoparticles is measured by the dynamic light-scattering (DLS) particle-size analysis
212
model using a Zetasizer Nano SZ (Malvern Instruments Ltd., Worcestershire, UK).
213
Raman spectra of the SrO-C-dots are recorded on a Renishaw inVia Raman
214
microscope equipped with RL785 and RL830 Class 3B wavelength stabilized diode
215
lasers and a Leica DM2500 M (Leica Microsystems) materials analysis microscope. A
216
powder sample was sprinkled on the glass slide and the Raman spectra are recorded.
9 ACS Paragon Plus Environment
Energy & Fuels
1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
217
The fluorescence properties of the SrO-C-dots are studied by a Varian Cary Eclipse
218
Fluorescence spectrophotometer. The specific surface area of the particles was
219
measured by the Brunauer–Emmett–Teller (BET) method at 77 K under liquid
220
nitrogen on a Micromeritics instrument (Gemini 2375) after the samples had been
221
evacuated at 25 °C for 12 h by an evacuator (Micromeritics, Flow Prep 060).
222
2.4. Cultivation and harvesting of Chlorella vulgaris
223
Freshwater microalga C. vulgaris cultures are grown in the laboratory in a 3 L sterile
224
flask using a Bristol medium (https://utex.org/products/bristol-medium), with air bub-
225
bling, under 70 µmol photons m-2s-1 and continuous LED light, at 22 °C with three
226
replicates. After a week, the algae are transferred outdoors to a 30 L flat-panel photo-
227
bioreactor (PBR). The light, temperature, oxygen and pH were monitored by a
228
computerized system for 12 days. The light is monitored by a Li-COR QUANTUM
229
sensor; the temperature and dissolved oxygen is monitored by LDO sensors (HACH
230
LANGE); the pH was monitored by pHD sc Digital Differential pH/ORP sensors
231
(HACH LANGE). The daylight (12 h) reached up to 2,200-2,500 µmol photons m-2s-
232
1
233
range was 6.1-7.8. During the first 7 days of the experiment, the initial inoculum
234
reached maximum culture density. Then nutrient deficiency diverted the biosynthesis
235
toward lipid accumulation in the algae cells. The algal culture was harvested by
236
centrifugation and freeze-dried in a lyophilizer.
237
2.5. In-situ transesterification of Chlorella vulgaris
238
The transesterification reaction was conducted under domestic microwave oven
239
(DMWO) irradiation (Figure S1 in the Supplementary Information). The DMWO was
240
operated at 2.45 GHz in batch mode. The output of the domestic microwave reactor
. The temperature reached up to 29-42 °C at noon and 16-24 °C at night. The pH
10 ACS Paragon Plus Environment
Page 10 of 31
Page 11 of 31
1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
Energy & Fuels
241
was 1,100 W .The microwave oven was operated at 70 % power (cycle mode of 21 s
242
on and 9 s off), a cycle mode function provided by the DMWO manufacturer. The
243
microwave oven was modified to allow the distillation column to pass through it (for
244
enhanced safety of operation) and a stirring facility during the reaction. The
245
modification was performed by replacing the bottom part of the oven by a rounded
246
aluminum plate. The plate was carefully attached to the framework in a manner that
247
enables magnetic stirring. The reaction temperature obtained as a result of the
248
microwave irradiation was measured using a Pyrometer (Fluke, 65 Infrared
249
Thermometer) after the irradiation was completed. The reaction temperature was
250
found to be 333 K immediately after completion of the irradiation.
251
The transesterification reactions are performed by a DMWO equipped with a
252
condenser and carried out in a 50 mL round-bottom flask. A typical batch process of
253
the transesterification reaction consists of taking 1 g of dried microalgae mixed with
254
10 mL of chloroform, 5 mL of MeOH (1:2 v/v) (based on the procedure of Folch et
255
al.31) and 0.3 g of catalyst, SrO-C-dots or commercial SrO, and irradiating the content
256
in a microwave oven for a certain time. First, the catalyst was dispersed in MeOH by
257
high magnetic stirring to ensure good dispersion of the catalyst into the MeOH. The
258
microalgae and chloroform are subsequently added and the mixture was irradiated. At
259
the end of the reaction, the temperature of the mixture is measured by a pyrometer and
260
was found to be 333 K, owing to the presence of the condenser which allowed the
261
cooling of the reaction. After the completion of the reaction, the samples are
262
centrifuged and filtered under vacuum to separate the methanol-chloroform phase that
263
contained the FAME from the microalgae, the glycerol and the catalyst. The solution
264
of methanol-chloroform was evaporated in a rotary evaporator, and the FAME content
11 ACS Paragon Plus Environment
Energy & Fuels
1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
Page 12 of 31
265
was determined gravimetrically.17 The oil content was expressed by its weight relative
266
to the weight of the microalgae biomass.
267
The FAME product is analyzed by 1H NMR spectroscopy (Bruker Avance 300
268
spectrometer). Chloroform (CDCl3) was used as a solvent for 1H NMR sample
269
preparation. The conversion is calculated directly from the integrated areas of the
270
methoxy group in the FAME at 3.65 ppm (singlet) and from the integrated areas of
271
the α-carbonyl methylene protons present in the triglyceride derivatives at 2.26 ppm
272
(triplet). Equation 1 was used to estimate the conversion of the lipid content to
273
FAME:
274
275
where
is the integrated value of the protons of the methyl esters and
276
integrated value of the methylene protons. The conversion ratio of the lipids in the
277
microalgae to the resultant FAME was obtained by dividing IMe by ICH2. The factors 2
278
and 3 are derived from the fact that methylene carbon possesses two protons and
279
methyl carbon has three attached protons.52
280
3. RESULTS AND DISCUSSION
281
3.1. Characterization of SrO-C-dot catalyst
282
3.1.1. Thermal analysis of Sr(NO3)2-C-dots and SrO-C-dot composite
283
The TGA analysis of the Sr(NO3)2@C-dots before and after calcination for 4 h at 650
284
°C is shown in Figure 2. The TGA of the Sr(NO3)2-C-dots before calcination was
285
performed in a nitrogen environment. The resulting material, SrO-C-dots, was further
286
analyzed by TGA in an air atmosphere to evaluate the temperature decomposition of 12 ACS Paragon Plus Environment
is the
Page 13 of 31
1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
Energy & Fuels
287
the C-dots. As indicated in Figure 2, in the case of the Sr(NO3)2-C-dots, the first
288
weight loss (~4 wt.%) occurs between 200-325 °C due to dehydration and loss of
289
water. The second and major wt. loss (48.5 wt. %) occurs between 650-660 °C,
290
corresponding to the decomposition of the Sr(NO3)2 to SrO. 32 Pure Sr(NO3)2 melts at
291
571 °C and decomposes at 658 °C.33 Our observation of the conversion of Sr(NO3)2 to
292
SrO is consistent with the literature.68 Moreover, the calcination temperature of 650
293
°C is reported as being the most suitable temperature for preparing solid-base
294
catalysts from Sr(NO3)2, as a result of its thermal property.32,34 The temperature must
295
be high enough to decompose Sr(NO3)2 to the active oxide species, yet not high
296
enough to hinder the sintering of the catalyst particles. Therefore, a calcination
297
temperature of 650 °C was subsequently used for the decomposition of the Sr(NO3)2-
298
C-dots to produce the SrO-C-dot catalyst. The TGA of the SrO-C-dot composite in air
299
indicates a weight loss of 7% at ~ 450 °C. The weight loss is attributed to the
300
decomposition of the C-dots or the loss of OH, CO and CO2 from the surface
301
functional groups of the C-dots. TGA analysis also confirms that the remaining 93 wt.
302
% is SrO and 7% is C-dots. Moreover, the plateau achieved at a temperature higher
303
than 500 °C indicates the thermal stability of the SrO material. According to the
304
decomposition formula of Sr(NO3)2 to SrO, the expected weight loss is 48.96 wt. %,
305
which is very close to the observed weight loss (48.5 wt. %).
306
3.1.2. XRD analysis of SrO-C-dot composition and determination of calcination
307
time
308
The crystal structure of the SrO-C-dot composite was analyzed using XRD. A
309
calcination time of 2-3 h at 650 °C in a N2 environment cannot completely convert all
310
the Sr(NO3)2-C-dots to SrO-C-dots (Figure 3). Intense signals of a strontianite
311
(SrCO3) (JCPDF file no. 84-1778) with orthorhombic phase belonging to a Pmcn 13 ACS Paragon Plus Environment
Energy & Fuels
1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
312
space group were identified (2θ = 25.3°, 26.0°, 37.3°, 44.5°, 46.8° and 48.0°) in
313
addition to the SrO phase. The formation of strontianite is probably due to the reaction
314
of the Sr(NO3)2 with the C-dots at elevated temperatures, even in an inert atmosphere,
315
due to the high reactivity of the carbon dots. The SrCO3 peaks are observed in the
316
samples that were calcined for a period of over 3 h. However, the intensity of the
317
peaks typical of Sr(CO3)2 decrease with calcination time. At the same time, the
318
intensity of the signals corresponding to oxide phase increased. A pure SrO phase is
319
observed in the material after calcination for 4 h at 650 °C (JCPDF file no. 06-0520;
320
cubic structure with a lattice constant value of 5.16 Å). The diffraction peaks at the 2θ
321
values of 30.0°, 34.8°, 50.2°, 59.4° and 62.4° respectively, correspond to the (111),
322
(200), (220), (311) and (222) planes of SrO. Signals typical of carbon structure were
323
not observed at ~25 and ~45° due to the much smaller amount of carbon dots in the
324
SrO-C-dot composite. Although, from the TGA analysis, the amount of carbon dots is
325
~7 wt. %, the carbon dots comprise of elemental carbon, oxygen and hydrogen
326
existing as surface functional groups like the carboxyl and phenolic groups.
327
Therefore, the actual amount of elemental carbon is even less than 7 wt. %. This may
328
be the reason for the absence of carbon signals in the XRD analysis (Figure 3). The
329
FT-IR spectrum of the SrO-C-dots (Figure S2) shows a broad and strong absorption
330
band at 592 cm-1, with a shoulder at 620 cm-1 attributed to the Sr–O stretching
331
vibration. The observed peak at 592 cm-1 attests to the formation of metal-oxide (Sr–
332
O) bonds.52,70,35 For comparison, commercial SrO was analyzed using FT-IR
333
spectroscope and a significant attenuation of typical absorption bands is observed,
334
probably due to the micron-size particles of commercial SrO that absorbs far less than
335
the nanometric SrO powder in the SrO-C-dot composite.
336
3.1.3. Surface morphology and composition of SrO-C-dot composite 14 ACS Paragon Plus Environment
Page 14 of 31
Page 15 of 31
1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
Energy & Fuels
337
The surface morphology of the SrO-C composite was deciphered from HR-SEM
338
analysis and the chemical composition of the composite was determined from EDS
339
analysis. The SEM images, the EDS spectrum and the elemental mapping recorded on
340
the SrO-C-dot composite are shown in Figure 4. Functionalization of SrO with C-
341
dots generates a high density of irregular small particles with a rough surface. In
342
contrast, commercial SrO comprises large dense particles, exposing smooth planes or
343
regularly facetted steps with micron-sized particles.36 The results of the SEM analysis
344
pertaining to the particle size and morphology of the SrO-C-dot composite is
345
consistent with the BET sorptometry results. The specific surface area of the SrO-C-
346
dot composite is 3.5 m2g-1, whereas that of commercial SrO is 0.5 m2g-1. BET specific
347
surface area value of the SrO-C-dot composite is 7 times greater than that of
348
commercial SrO.49 The particle size distribution of the catalyst was also evaluated by
349
HRSEM and the average particle size is 20-30 nm. Elemental mapping of the SrO-C-
350
dot composite shows uniform distribution of Sr, O and elemental carbon on the
351
surface. This indicates the complete decomposition of the Sr(NO)3-C-dot to SrO-C-
352
dot composite as well as the homogeneous surface functionalization of SrO with
353
carbon dots. The elemental composition shown by the EDS analysis is considered
354
only for qualitative purpose and not as a quantitative estimate of carbon owing to the
355
fact that the analysis is carried out by spreading the sample on a carbon tape placed on
356
a Cu stub.
357
3.1.4. TEM and DLS analysis of SrO-C-dots
358
The TEM image of SrO NPs surrounded by carbon dots displays a porous (Figure 5).
359
The average particle size of carbon dots on the SrO surface is ~ 7 nm. The average
360
particle size of the SrO is also measured in an isopropyl alcohol suspension by DLS
361
(Figure S3). The suspension was formed using a bath sonicator. The results of the 15 ACS Paragon Plus Environment
Energy & Fuels
1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
Page 16 of 31
362
DLS analysis show that the size of the SrO particles in the composite is ~ 200 nm.
363
The particle size measured by DLS is about 10 times greater than that measured by
364
HRSEM. The larger particle size values obtained by DLS measurement are due to
365
particle aggregation.
366
3.1.5. Raman analysis
367
To confirm the presence of C-dots on the surface of SrO in the SrO-C-dot composite,
368
Raman analysis was performed. Figure S4 shows the appearance of two prominent
369
peaks at 1,380 and 1,549 cm-1, corresponding to the D- and G-bands respectively. The
370
peak pattern is characteristic of the formation of carbon dots.37 Moreover, in the case
371
of the SrO-C-dot composite, the signal intensity of the graphic band increases
372
compared to the D-band, which is attributed to the increase in the crystallinity of the
373
carbon structure formed at a high temperature calcination (650 °C for 4 h in an N2
374
atmosphere), during which a variety of reactions such as dehydration, cyclization and
375
condensation of the surface functional groups of carbon dots can take place.
376
3.2.
377
transesterification of lipids of Chlorella vulgaris
378
The catalytic activity of the SrO-C-dot composite during the transesterification
379
reaction is tested. A one-step direct transesterification reaction is carried out by
380
irradiating the mixture of methanol and chloroform (1:2 v/v) and Chlorella vulgaris in
381
the presence of SrO-C-dot catalyst in a domestic microwave oven at various
382
irradiation times (2 to 5 min). The effect of the microwave irradiation time on the
383
conversion of the lipids and the yield of FAME is depicted in Figure 6. A lipid
384
conversion value of 97 wt. % and a corresponding FAME yield of 45.5 wt. %,
385
calculated gravimetrically, were achieved in a short reaction time of 2.5 min. The
Evaluation
of
the
catalytic
activity
of
16 ACS Paragon Plus Environment
SrO-C-dots
for
the
Page 17 of 31
1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
Energy & Fuels
386
values of conversion of lipids and yield of FAME are achieved with a modest amount
387
of catalyst (0.3 g) comprising 93 wt. % SrO (0.279 g) and 7 wt. % (0.021 g) C-dots.
388
For comparison, commercial micrometer-sized SrO particles were also used as
389
the catalyst for the direct transesterification of Chlorella vulgaris under identical
390
reaction conditions. A plot of the variation in the conversion of the lipid and yield of
391
FAME using a commercial SrO catalyst is depicted in Figure 7. With the commercial
392
SrO catalyst, an irradiation time of 6 min is required to reach similar conversion and
393
yield values obtained using the SrO-C-dot composite catalyst in 2.5 minutes. Thus,
394
the transesterification is accelerated by approximately 2.4 times using the SrO-C-dot
395
catalyst, indicating its improved performance over the conventional SrO catalyst. The
396
improved activity of the SrO-C-dot composite for the acceleration of the
397
transesterification of lipid content is attributed to the unique structural features of the
398
C-dots. C-dots are extremely hydrophilic owing to their extensive surface
399
functionalization. Surface functionalities such as carboxylic and phenolic groups are
400
acidic in nature. These moieties play a crucial role in the esterification of free fatty
401
acids (FFA) in the algal biomass. Thus, the synergistic role of the acidic sites of the
402
C-dots as well as the highly basic sites of SrO result in a 2.5-fold acceleration in the
403
transesterification reaction compared to pure SrO. Another factor that is responsible
404
for the enhanced activity of the SrO-C-dot composite is the surface area value, which
405
is 7 times higher than that of commercial SrO.
406
A typical 1H NMR spectrum of the FAME product obtained from Chlorella
407
vulgaris using the SrO-C composite upon irradiation for 2.5 min is shown in Figure
408
8. The yield of FAME is estimated gravimetrically by calculating its weight relative to
409
the weight of the dry microalgae biomass. The conversion percentage of oil to FAME
410
is calculated by integrating the 1H NMR peaks as shown in Figure 8 using Eq. 1. 17 ACS Paragon Plus Environment
Energy & Fuels
1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
411
When SrO-C-dot composite is used as catalyst optimal conversion is achieved after
412
irradiation for 2.5 min resulting in an oil weight percentage of 45.5%, which
413
corresponds to the lipid conversion value of 97 wt. %, as deduced from the 1H NMR
414
spectrum (Figure 8). Conversion values greater than 97 wt. % could not be achieved
415
even after irradiation for a longer time (Figure 6). This is probably due to the
416
presence of FFA in the microalgae.38
417
Besides the characteristic peak at 3.6 ppm, which arises from the methyl-ester,
418
and the triplet at 2.3 ppm related to α-carbonyl methylene, which are characteristic
419
peaks confirming the presence of methyl esters in FAME, the 1H NMR spectrum
420
shows a strong peak at 1.226 ppm arising from methylene protons of the carbon chain
421
of esters and a triplet at 0.851 ppm is related to terminal methyl-hydrogen protons.
422
The other peak, which appears at 1.576 ppm, is attributed to β-carbonyl methylene
423
protons.39 The absence of a signal between 4.05-4.37 ppm in the 1H NMR spectrum is
424
attributed to the absence of CH2 groups (–CH2–O–COR) of triglycerides, indicating
425
high yield of FAME and high catalytic activity of the SrO-C-dots.25 Thus, a novel
426
composite of SrO and C-dots has been designed and tested for the accelerated
427
production of FAME in a single step from Chlorella vulgaris.
428
Designing green and sustainable methods for extraction is a vital research
429
topic in multidisciplinary areas. Recently, Crampon et al., reported effective
430
extraction of neutral lipids from the microalgae Spirulina platensis using super critical
431
CO2.40 Chemat et al., introduced the concept of six principles of green-extraction that
432
could be applied to industrial process for intensification and energy saving.41 From
433
the perspective of reduced energy consumption,26 safe and high quality extraction of
434
lipids from the algae using microwave irradiation as well as ultrasonic waves. Even
435
though, microalgae is a versatile feedstock for biofuel production due to the tenability 18 ACS Paragon Plus Environment
Page 18 of 31
Page 19 of 31
1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
Energy & Fuels
436
of chemical composition (for instance obtaining high lipid content algae), higher
437
productivity and photosynthetic efficiency,42-45 for the biodiesel production process to
438
be sustainable and competitive several challenges like resource constraints (CO2,
439
water, nutrients), development of innovative large scale production technology require
440
immense research efforts and remains a challenge.46-50
441
4. CONCLUSION
442
A novel SrO-C-dot catalyst is prepared from Sr(NO3)2 and PEG400 as precursors for
443
Sr and C, respectively, under sonochemical conditions. Homogeneous surface
444
functionalization of SrO nanoparticles with carbon dots is achieved using the
445
synthetic strategy developed based on sonication and subsequent calcination in an
446
inert atmosphere at 650 °C for 4 h. The developed catalyst is used for the single-step
447
conversion of Chlorella vulgaris to FAME in a microwave oven. A modest lipid
448
conversion value of 97 wt. % and a FAME yield of 45.5 wt. % are achieved using the
449
SrO-C-dot composite catalyst in a short reaction time of 2.5 min. The SrO-C-dot
450
catalyst shows improved performance for the transesterification of lipids compared to
451
the conventional SrO catalyst. The enhanced catalytic activity (2.4-fold) of the SrO-C-
452
dot composite is attributed to the unique structural features of carbon-dots with a
453
variety of oxygen, carboxylic and phenolic functionalities that promote the
454
esterification of FFAs in the algal biomass. Moreover, the higher surface area of SrO-
455
C-dots compared to commercial SrO is another factor that enhances the catalytic
456
activity with an increased number of accessible active reactant sites (lipids and
457
MeOH). Thus, a SrO-C-dot composite catalyst is designed in an innovative way with
458
very high potential for the fast and enhanced production of FAME from Chlorella
459
vulgaris. This new methodology can be used for developing a CaO/BaO/MgO-C-dot
460
composite exclusively for biodiesel production. 19 ACS Paragon Plus Environment
Energy & Fuels
1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
461
ASSOCIATED CONTENT
462
Supporting Information.
463
Image of the modified microwave oven (Figure S1), FT-IR spectra of SrO-C-dots and
464
commercial SrO catalysts (Figure S2), dynamic light scattering (DLS) trace of SrO-C
465
dot composite (Figure S3) and Raman spectra of C-dots and SrO-C-dot composite
466
(Figure S4) are provided in the supporting information. This material is available free
467
of charge via the internet at http://pubs.acs.org.
468
AUTHOR INFORMATION
469
Corresponding Author
470
*Phone: 972-3-5318315; Fax: 972-3-7384053.
471
E-mail:
[email protected].
472
Notes
473
The authors declare no competing financial interest.
474
Author Contributions
475
# Alex Tangy and Vijay Bhooshan Kumar have contributed equally for this work.
476
ACKNOWLEDGEMENTS
477
Gedanken thanks the Israeli Ministry of Science, Technology and Space for the
478
research grant (206712) supporting this work. Grateful thanks are due for the India-
479
Israel cooperative scientific research grant (203768) supporting this research.
480
REFERENCES
481
(1)
Atabani, A. E.; Silitonga, A. S.; Badruddin, I. A.; Mahlia, T. M. I.; Masjuki, H. 20 ACS Paragon Plus Environment
Page 20 of 31
Page 21 of 31
1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
Energy & Fuels
H.; Mekhilef, S. A Comprehensive Review on Biodiesel as an Alternative Energy Resource and Its Characteristics. Renew. Sustain. Energy Rev. 2012, 16 (4), 2070–2093.
482 483 484 485 486 487
(2)
Aransiola, E. F.; Ojumu, T. V.; Oyekola, O. O.; Madzimbamuto, T. F.; IkhuOmoregbe, D. I. O. A Review of Current Technology for Biodiesel Production: State of the Art. Biomass and Bioenergy 2014, 61, 276–297.
488 489 490
(3)
Luque, R.; Lovett, J. C.; Datta, B.; Clancy, J.; Campelo, J. M.; Romer, A. A. Biodiesel as Feasible Petrol Fuel Replacement: A Multidisciplinary Overview. Energy Environ. Sci. 2010, 3, 1706–1721.
491 492
(4)
Baskar, G.; Aiswarya, R. Trends in Catalytic Production of Biodiesel from Various Feedstocks. Renew. Sustain. Energy Rev. 2016, 57, 496–504.
493 494 495
(5)
Fu, C. C.; Hung, T. C.; Chen, J. Y.; Su, C. H.; Wu, W. T. Hydrolysis of Microalgae Cell Walls for Production of Reducing Sugar and Lipid Extraction. Bioresour. Technol. 2010, 101 (22), 8750–8754.
496 497
(6)
Leite, G. B.; Abdelaziz, A. E. M.; Hallenbeck, P. C. Algal Biofuels: Challenges and Opportunities. Bioresour. Technol. 2013, 145, 134–141.
498 499
(7)
Foley, P. M.; Beach, E. S.; Zimmerman, J. B. Algae as a Source of Renewable Chemicals: Opportunities and Challenges. Green Chem. 2011, 13 (6), 1399.
500 501 502
(8)
Singh, B.; Guldhe, A.; Rawat, I.; Bux, F. Towards a Sustainable Approach for Development of Biodiesel from Plant and Microalgae. Renew. Sustain. Energy Rev. 2014, 29, 216–245.
503 504
(9)
Wijffels, R. H.; Barbosa, M. J. An Outlook on Microalgal Biofuels. Science (80-. ). 2010, 329, 796–799.
505 506 507
(10)
Yan, X.; Tan, D. K. Y.; Inderwildi, O. R.; Smith, J. a. C.; King, D. a. Life Cycle Energy and Greenhouse Gas Analysis for Agave-Derived Bioethanol. Energy Environ. Sci. 2011, 4 (9), 3110.
508 509 510 511
(11)
Sharma, Y. C.; Singh, B.; Korstad, J. A Critical Review on Recent Methods Used for Economically Viable and Eco-Friendly Development of Microalgae as a Potential Feedstock for Synthesis of Biodiesel. Green Chem. 2011, 13 (11), 2993–3006.
512 513 514
(12)
Lin, L.; Cunshan, Z.; Vittayapadung, S.; Xiangqian, S.; Mingdong, D. Opportunities and Challenges for Biodiesel Fuel. Appl. Energy 2011, 88 (4), 1020–1031.
515 516
(13)
Savage, P. E. Algae Under Pressure and in Hot Water. Science (80-. ). 2012, 338 (6110), 1039–1040.
517 518 519
(14)
Milano, J.; Ong, H. C.; Masjuki, H. H.; Chong, W. T.; Lam, M. K.; Loh, P. K.; Vellayan, V. Microalgae Biofuels as an Alternative to Fossil Fuel for Power Generation. Renew. Sustain. Energy Rev. 2016, 58, 180–197.
520 521
(15)
Yoo, G.; Yoo, Y.; Kwon, J.-H.; Darpito, C.; Mishra, S. K.; Pak, K.; Park, M. S.; Im, S. G.; Yang, J.-W. An Effective, Cost-Efficient Extraction Method of 21 ACS Paragon Plus Environment
Energy & Fuels
1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
Biomass from Wet Microalgae with a Functional Polymeric Membrane. Green Chem. 2014, 16 (1), 312.
522 523 524 525 526
(16)
Passos, H.; Freire, M. G.; Coutinho, J. A. P. Ionic Liquid Solutions as Extractive Solvents for Value-Added Compounds from Biomass. Green Chem. 2014, 16 (12), 4786–4815.
527 528 529
(17)
Koberg, M.; Cohen, M.; Ben-Amotz, A.; Gedanken, A. Bio-Diesel Production Directly from the Microalgae Biomass of Nannochloropsis by Microwave and Ultrasound Radiation. Bioresour. Technol. 2011, 102 (5), 4265–4269.
530 531 532
(18)
Hidalgo, P.; Toro, C.; Ciudad, G.; Navia, R. Advances in Direct Transesterification of Microalgal Biomass for Biodiesel Production. Rev. Environ. Sci. Biotechnol. 2013, 12 (2), 179–199.
533 534 535
(19)
Nautiyal, P.; Subramanian, K. A.; Dastidar, M. G. Production and Characterization of Biodiesel from Algae. Fuel Process. Technol. 2014, 120, 79–88.
536 537 538 539
(20)
Patil, P. D.; Reddy, H.; Muppaneni, T., Mannarswamy, A., Schuab, T., Holguin, F. O., Lammers, P., Nirmalakhandan, N., Cooke, P., Deng, S. Power Dissipation in Microwave-Enhanced in Situ Transesterification of Algal Biomass to Biodiesel. Green Chem. 2012, 14 (3), 809–818.
540 541 542
(21)
Chang, F.; Zhou, Q.; Pan, H.; Liu, X.; Zhang, H.; Xue, W.; Yang, S. Solid Mixed-Metal-Oxide Catalysts for Biodiesel Production: A Review. Energy Technol. 2014, 2 (11), 865–873.
543 544 545
(22)
Yan, S.; Dimaggio, C.; Mohan, S.; Kim, M.; Salley, S. O.; Ng, K. Y. S. Advancements in Heterogeneous Catalysis for Biodiesel Synthesis. Top. Catal. 2010, 53 (11–12), 721–736.
546 547 548
(23)
Faungnawakij, K.; Yoosuk, B.; Namuangruk, S.; Krasae, P.; Viriya-empikul, N.; Puttasawat, B. Sr-Mg Mixed Oxides as Biodiesel Production Catalysts. ChemCatChem 2012, 4 (2), 209–216.
549 550 551
(24)
Nasreen, S.; Liu, H.; Khan, R.; Zhu, X. C.; Skala, D. Transesterification of Soybean Oil Catalyzed by Sr-Doped Cinder. Energy Convers. Manag. 2015, 95, 272–280.
552 553 554 555
(25)
Koberg, M.; Abu-Much, R.; Gedanken, A. Optimization of Bio-Diesel Production from Soybean and Wastes of Cooked Oil: Combining Dielectric Microwave Irradiation and a SrO Catalyst. Bioresour. Technol. 2011, 102 (2), 1073–1078.
556 557 558
(26)
Tangy, A.; Pulidindi, I. N.; Gedanken, A. SiO 2 Beads Decorated with SrO Nanoparticles for Biodiesel Production from Waste Cooking Oil Using Microwave Irradiation. Energy & Fuels 2016, 30 (4), 3151–3160.
559 560
(27)
Li, H.; Kang, Z.; Liu, Y.; Lee, S.-T. Carbon Nanodots: Synthesis, Properties and Applications. J. Mater. Chem. 2012, 22 (46), 24230.
561 562
(28)
Zheng, X. T.; Ananthanarayanan, A.; Luo, K. Q.; Chen, P. Glowing Graphene Quantum Dots and Carbon Dots: Properties, Syntheses, and Biological 22 ACS Paragon Plus Environment
Page 22 of 31
Page 23 of 31
1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
Energy & Fuels
Applications. Small 2015, 11 (14), 1620–1636.
563 564 565 566
(29)
Cheng, J.; Yu, T.; Li, T.; Zhou, J.; Cen, K. Using Wet Microalgae for Direct Biodiesel Production via Microwave Irradiation. Bioresour. Technol. 2013, 131, 531–535.
567 568 569
(30)
Kumar, V. B.; Porat, Z.; Gedanken, A. Facile One-Step Sonochemical Synthesis of Ultrafine and Stable Fluorescent C-Dots. Ultrason. Sonochem. 2016, 28, 367–375.
570 571 572 573
(31)
Gossert, A. D.; Hinniger, A.; Gutmann, S.; Jahnke, W.; Strauss, A.; Fernandez, C. A Simple Protocol for Amino Acid Type Selective Isotope Labeling in Insect Cells with Improved Yields and High Reproducibility. Journal of Biomolecular NMR. 2011, pp 449–456.
574 575
(32)
Tantirungrotechai, J.; Thepwatee, S.; Yoosuk, B. Biodiesel Synthesis over Sr/MgO Solid Base Catalyst. Fuel 2013, 106, 279–284.
576 577 578
(33)
Hosseini, S. G.; Eslami, A. Thermoanalytical Investigation of Relative Reactivity of Some Nitrate Oxidants in Tin-Fueled Pyrotechnic Systems. J. Therm. Anal. Calorim. 2010, 101 (3), 1111–1119.
579 580 581
(34)
Chen, W.; Huang, Z.; Liu, Y.; He, Q. Preparation and Characterization of a Novel Solid Base Catalyst Hydroxyapatite Loaded with Strontium. Catal. Commun. 2008, 9 (4), 516–521.
582 583 584
(35)
Rahman, M. M.; Hussain, M. M.; Asiri, A. M. A Novel Approach towards Hydrazine Sensor Development Using SrO·CNT Nanocomposites. RSC Adv. 2016, 6 (70), 65338–65348.
585 586 587
(36)
Yoosuk, B.; Krasae, P.; Puttasawat, B.; Udomsap, P.; Viriya-empikul, N.; Faungnawakij, K. Magnesia Modified with Strontium as a Solid Base Catalyst for Transesterification of Palm Olein. Chem. Eng. J. 2010, 162 (1), 58–66.
588 589 590
(37)
Wang, Q.; Liu, X.; Zhang, L.; Lv, Y. Microwave-Assisted Synthesis of Carbon Nanodots through an Eggshell Membrane and Their Fluorescent Application. Analyst 2012, 137 (22), 5392–5397.
591 592 593
(38)
Dong, T.; Wang, J.; Miao, C.; Zheng, Y.; Chen, S. Two-Step in Situ Biodiesel Production from Microalgae with High Free Fatty Acid Content. Bioresour. Technol. 2013, 136, 8–15.
594 595 596
(39)
Monteiro, M. R.; Ambrozin, A. R. P.; Liao, L. M.; Ferreira, A. G. Determination of Biodiesel Blend Levels in Different Diesel Samples by 1H NMR. Fuel 2009, 88 (4), 691–696.
597 598 599 600
(40) Crampon, C.; Nikitine, C.; Zaier, M.; Lepine, O.; Tanzi, C. D.; Vian, M. A.; Chemat, F.; Badens, E. Oil extraction from enriched Spirulina platensis microalgae using supercritical carbon dioxide. J. of supercritical Fluids, 2017, 119, 289-296.
601 602 603
(41) Chemat, F.; Vian, M. A.; Cravotto, G.; Green extraction of natural products: Concept and principles. Int. J. Mol. Sci., 2012, 13, 8615-8627.
23 ACS Paragon Plus Environment
Energy & Fuels
1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
604 605
[42] R. Rajkumar, Z. Yaakob, M. S. Takriff, Potential of the micro and macro algae for biofuel production: A brief Review. BioResources, 9 (1), 1606-1633, 2014.
606 607 608
[43] A. P. Dean, D. C. Sigee, B. Estrada, J. K. Pittman, Andrew P. Dean. David C. Sigee, Beatriz Estrada, Jon K. Pittman, Bioresource Technology 101, 4499–4507, 2010.
609 610
[44] M. Stockenreiter, A. K. Graber, F. Haupt, H. Stibor, The effect of species diversity on lipid production
611
by micro-algal communities, J Appl Phycol DOI 10.1007/s10811-010-9644-1
612 613
[45] U. S. Beevi, R. K. Sukumaran, Cultivation of microalgae in dairy effluent for oil production and removal
614
of organic pollution load, Bioresource Technology 165, 295–301, 2014.
615 616 617 618 619
[46] E. Stephens, R. de Nys, I. L. Ross and B. Hankamer, J Pet Environ Biotechnol 2013, 4:4 http://dx.doi.org/10.4172/2157-7463.1000148. [47] I. Hamawand, T. Yusaf, S. Hamawand, Growing algae using water from coal seam gas industry and harvesting using an innovative technique: A review and a potential. Fuel 117, 422–430, 2014.
620 621 622
[48] C. C. S. Steve, N. S. Mufti, M. D. Chamberlain, J. Kim, A. R. Wheeler, A droplet-based screen for wavelength-dependent lipid production in algae, Energy Environ. Sci., 7, 2366-2375, 2014.
623 624 625
[49] Z. Yue, D. Ma, J. Wang, J. Tan, S. C. Peng, T.H. Chen, Goethite promoted anaerobic digestion of algal biomass in continuous stirring-tank reactors, Fuel, 159, 883–886, 2015.
626 627 628
[50] A. S. Pedro, C.V. González-López, F.G. Acién, E. Molina-Grima, Outdoor pilot production of Nannochloropsis gaditana: Influence of culture parameters and lipid production rates in flat-panel photobioreactors. Algal Research 18, 156–165, 2016.
629 630 631 632 633 634 635 636 637 638
LIST OF FIGURES:
639 24 ACS Paragon Plus Environment
Page 24 of 31
Page 25 of 31
1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
Energy & Fuels
640 641 642
Figure 1. Schematic overview of the sonochemical synthesis of the Sr(NO3)2-C-dot composite.
643 644 645
646 25 ACS Paragon Plus Environment
Energy & Fuels
1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
647 648 649
Figure 2. TGA curve of Sr(NO3)2-C-dots in N2 atmosphere and of the SrO-C-dot composite in air after calcination of Sr(NO3)2-C-dots at 650 °C for 4 h in an N2 environment at a constant heating rate of 10 °C/min.
650 651 652 653 654 655
656 657 658
Figure 3. Effect of calcination time on the conversion of Sr(NO3)2-C-dots to SrO-Cdots (T = 650 °C; t = 1 – 4 h).
659 660 661 662
26 ACS Paragon Plus Environment
Page 26 of 31
Page 27 of 31
1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
Energy & Fuels
a
663 664 665 666
Figure 4. HR-SEM images (a and b), EDS spectrum (c) and elemental mapping (d) of the SrO-C-dot composite.
667 668 669 670 671 672 673 674 675 676 677 678 679 680 681 682 683 684 685 686 687 688 689 690 691 692 27 ACS Paragon Plus Environment
Energy & Fuels
1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
693 694 695 696 697 698 699 700 701 702 703 704 705 706 707 708 709 710 711 712 713 714 715 716 717 718 719 720 721
Figure 5. TEM of SrO-C-dot composite
28 ACS Paragon Plus Environment
Page 28 of 31
Page 29 of 31
1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
Energy & Fuels
722
723 724 725 726 727 728 729 730 731
Figure 6. Effect of microwave irradiation time on FAME yield and conversion of lipids of Chlorella vulgaris to FAME using SrO-C-dot composite as catalyst.
29 ACS Paragon Plus Environment
Energy & Fuels
1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
732 733 734 735 736 737 738 739 740 741 742 743 744 745 746 747 748 749 750 751 752 753 754 755 756 757 758
Figure 7. Effect of microwave irradiation time on FAME yield and conversion of lipids of Chlorella vulgaris to FAME using a commercial SrO catalyst.
30 ACS Paragon Plus Environment
Page 30 of 31
Page 31 of 31
1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
Energy & Fuels
759 760 761 762 763 764 765 766 767 768 769 770 771 772 773 774 775 776 777 778 779 780 781
Figure 8. 1H NMR spectrum of FAME obtained from the direct transesterification of Chlorella vulgaris by microwave irradiation for 2.5 min using the SrO-C-dot composite catalyst.
31 ACS Paragon Plus Environment