You are on page 1of 8

Fuel 153 (2015) 4855

Contents lists available at ScienceDirect

Fuel
journal homepage: www.elsevier.com/locate/fuel

A biomimetic silicication approach to synthesize CaOSiO2 catalyst for


the transesterication of palm oil into biodiesel
Guanyi Chen a,b,c, Rui Shan a, Shangyao Li a, Jiafu Shi a,
a

School of Environmental Science and Engineering, Tianjin University, Tianjin 300072, China
State Key Laboratory of Internal Combustion Engines, Tianjin University, Tianjin 300072, China
c
Key Laboratory of Efcient Utilization of Low and Medium Grade Energy (Tianjin University), Ministry of Education, Tianjin 300072, China
b

h i g h l i g h t s

g r a p h i c a l a b s t r a c t

 A biomimetic silicication approach

was developed to prepare CaOSiO2


catalysts.
 The catalysts were synthesized by
using eggshell and Na2SiO3 as raw
materials.
 With increase of Si compound, the
reusability of catalysts enhanced
signicantly.

a r t i c l e

i n f o

Article history:
Received 23 November 2014
Received in revised form 4 February 2015
Accepted 24 February 2015
Available online 7 March 2015
Keywords:
Biomimetic silicication
Biodiesel
CaOSiO2 catalyst
Eggshell
Transesterication

a b s t r a c t
In the present study, CaOSiO2 catalysts were successfully synthesized through a biomimetic silicication
approach by using eggshell and Na2SiO3 as raw materials. More specically, the powdered egg shells,
where lysozyme (the inducer) was located, were dispersed into Na2SiO3 aqueous solution to implement
the biomimetic silicication under ambient conditions. The as-obtained egg shell-SiO2 composites were
then calcined at 800 C under oxygen atmosphere, thus acquiring the CaOSiO2 catalysts. These catalysts
were detailedly characterized by SEM, EDS, 29Si NMR, FTIR, XRD, BET analysis, and Hammett indicator.
When utilized for catalytic transesterication of palm oil, the CaOSiO2 catalysts exhibited decreased
catalytic activity and increased reusability as the amount of Si compounds increased. Particularly, when
the Na2SiO3 concentration was below 0.4 M, a slight decrease in the catalytic activity could be observed
(0Si5Ca, 1Si5Ca and 2Si5Ca showed good performance with biodiesel yields of 90.2%, 87.7% and 80.1%, respectively), whereas the reusability of the catalyst was signicantly improved (little deactivation was found
after 12 cycles on the 2Si5Ca catalyst during the transesterication reaction). Hopefully, this biomimetic
silicication approach can be applied for the synthesis of a wide range of efcient and stable solid base
catalysts for transesterication/esterication reactions.
2015 Elsevier Ltd. All rights reserved.

1. Introduction
Corresponding author. Tel./fax: +86 22 27890566.
E-mail address: shijiafu@tju.edu.cn (J. Shi).
http://dx.doi.org/10.1016/j.fuel.2015.02.109
0016-2361/ 2015 Elsevier Ltd. All rights reserved.

Recently, due to the excessive consumption of fossil fuel and


increased demand of energy, exploiting alternative energy sources

G. Chen et al. / Fuel 153 (2015) 4855

have gained great attentions. Biodiesel derived from vegetable oils


or animal fats is one of the promising sustainable fuels mainly
owing to its renewability, sustainability and biodegradability
[1,2]. Generally, biodiesel can be synthesized via transesterication
of triglycerides (such as vegetable oils or animal fats) with alcohols
catalyzed by homogeneous base catalysts [3,4]. Although the
homogeneous base catalysts usually exhibit high catalytic activity,
they possess several disadvantages, such as occurrence of saponication and reactor corrosion, difculty in catalyst reuse, and the
generation of large amount of waste water [5,6]. To circumvent
these issues, heterogeneous base catalysts are exploited and adopted to catalytically produce biodiesel, which commonly possess the
following outstanding advantages, e.g., ease of separation, none
toxicity, none corrosion, or less environmental pollution [7].
Presently, various solid catalysts have been synthesized, such as
metal oxides, composite metal oxides, zeolites, hydrotalcites, anion
exchange resins and lipase immobilized on/in various supports
[3,813].
As a typical solid base catalyst, CaO has provided great potentials in the biodiesel production due to its high catalytic activity
and inexpensive preparation process [14,15]. To further reduce
the synthesis cost of biodiesel, a series of eggshell/mollusk shellderived CaO have been widely used as solid base catalysts for
biodiesel production [1618]. These shell-derived CaO catalysts
could not only eliminate the waste but display desirable catalytic
performance in the transesterication reaction. However, CaO
would be rapidly contaminated by water or CO2 (through hydration or carboxylation) under ambient conditions, thus resulting
in a decreased catalytic activity [17]. Besides, the partial dissolution of Ca2+ ions from the CaO surface often occurs during the
transesterication reaction. The removal of Ca2+ ions from the biodiesel product commonly requires additional separation/purication processes, which would generate a large amount of waste
water to the environment [7,19]. For example, Granados et al.
[20] found that pure CaO was partially soluble in methanol and a
larger amount of leached Ca2+ ions was observed in the glycerol
phase. In order to address these problems, some composite catalysts have been synthesized to enhance the stability of CaO, including MgOCaO mixed oxides [21], CaOZnO mixed oxides [22,23],
CaOCeO2 mixed oxides, and so on [7]. These composite catalysts
can reduce the leaching amount of Ca2+ ions through generating
some interactions (such as electrostatic interaction) between CaO
and other inorganic oxides [24].
Particularly, a variety of CaOSiO2 composite catalysts have
been developed against the Ca2+ ions leaching problem [19,25,26].
For instance, Xie and Zhao [25] prepared CaOMoO3SBA-15 catalysts through an incipient impregnation method for catalytically
producing biodiesel from soybean. The obtained catalysts could
be easily recovered and reused without signicant loss of activity.
Sun et al. [26] synthesized mesoporous CaO/SBA-15 catalysts via
a hydrothermal method for transesterication of sunower oil with
methanol. The acquired CaO/SBA-15 catalysts showed a much better stability than those prepared by conventional impregnation
method. The enhancement of the stability for the SBA-15 supported
catalysts was mainly owing to the following two aspects: (1) the
better distribution of Ca2+ ions on the SBA-15 support surface;
and (2) the formation of CaOSi bond through the adsorption of
the leached Ca2+ ions by SiO2 compound in the SBA-15 support.
However, the preparation procedure of those catalysts was often
complicated and time-consuming. Besides, the raw materials were
often expensive. All the above restricts made these catalysts difculty for large-scale production. Therefore, developing a simple
and efcient method to synthesize composite solid base catalysts
is urgently required and faces great challenges.
In the last decades, biosilicication/biomimetic silicication, a
process of utilizing cationic proteins/polymers to induce the

49

generation of SiO2 from aqueous precursors, has drawn great


attentions and showed potentials in some application areas (e.g.,
catalysis, drug delivery, sensoring, etc.) [2734]. Till now, several
proteins/polymers, including silicateins, silafns, lysozyme, protamine, polyethylenimine, etc., have been applied for fabricating
SiO2 upon the biosilicication/biomimetic silicication approach
[2734]. Amongst, lysozyme that is widely existed within the eggshell matrix, shows desirable silicication-inducing capability.
Besides, some researchers have extracted lysozyme from eggshell
for the synthesis of SiO2 and its derivatives [31,3335]. Suppose
that directly utilization of eggshell to implement the silicication
process may acquire eggshell-SiO2 composites, which could offer
a promising precursor for synthesizing CaOSiO2 catalysts.
In the present study, we utilize the eggshell (mainly composed
of lysozyme and CaCO3) to induce the silicicaiton of Na2SiO3 (a
SiO2 precursor), thus acquiring the CaCO3SiO2 composites. After
calcination, the as-obtained CaCO3SiO2 composites could be converted into the CaOSiO2 catalysts, where SiO2 is primarily
attached on the surface of CaO. Afterwards, the physicochemical
properties of the prepared CaOSiO2 catalysts are characterized
by scanning electron microscopy (SEM), energy dispersive X-ray
spectrometer (EDS), 29Si NMR spectrometer, Fourier transform
infrared (FTIR) spectrometer, X-ray diffraction (XRD), Brunauer
EmmettTeller (BET) analysis and the Hammett indicator method.
The effects of reaction conditions (such as reaction time, reaction
temperature, methanol-to-oil molar ratio and amount of catalyst
loading) and the catalyst reusability are also thoroughly studied.
2. Materials and methods
2.1. Materials
Rened palm oil was purchased from local market. Chicken
eggshells were collected from local restaurants. Methanol, sodium
silicate (Na2SiO3) and tetrahydrofuran (analytical grade) were purchased from SigmaAldrich Chemical Co. Ltd. All other reagents
were of analytical grade and used without further purication.
Deionized water was used throughout all the experiments. The
pH values of solutions were measured with a PHS-3C pH-meter
(REX Instruments, PHS-3C, Shanghai, China) and adjusted by addition of HCl solution (100 mM) or NaOH solution (100 mM).
2.2. Catalyst preparation and characterizations
The chicken eggshells were washed by deionized water to
remove dust and impurities, and then dried overnight in an oven
at 105 C. The as-acquired chicken eggshells were crushed in a
grinder for further use.
The CaOSiO2 catalysts were prepared as follows: 5 g of chicken
eggshell powders were added to 50 mL aqueous solution of
Na2SiO3 with different concentrations (0, 0.2, 0.4, 0.6 and 1 M) at
pH 8.0. The mixture solution was continuously stirred at 500 rpm
for 4 h, followed by drying overnight at 100 C and calcination at
800 C (calcination rate: 10 C min1; calcination time: 4 h). After
calcination, the catalysts were kept in the vacuum oven for further
use. These catalysts synthesized with different Na2SiO3 concentrations (0, 0.2, 0.4, 0.6 and 1 M) were denoted as the 0Si5Ca, 1Si5Ca,
2Si5Ca, 3Si5Ca and 5Si5Ca, respectively.
The X-ray diffractometer (XRD-Bruker; D8 Advance) patterns of
the several catalysts were analyzed using a powder coupled with
Cu Ka radiation. The FTIR spectra were recorded on a Shimadzu
IR-Prestige-21 spectrometer in the range of 5502000 cm1. The
KBr pellet method was used for the sample preparation. The BET
surface area was measured by N2 adsorptiondesorption isotherm
apparatus (BEL; BELSORP-max). The SEM images and surface

50

G. Chen et al. / Fuel 153 (2015) 4855

Fig. 1. SEM images of the (a) 0Si5Ca and (b) 2Si5Ca catalysts.

Table 1
EDS analysis of the surface of the 0Si5Ca and 2Si5Ca catalysts.
Image

Text point

Fig. 1a
Fig. 1b

A
A
B
C

Elements content (wt.%)

Others

Ca

Si

36.58
32.42
26.88
28.26

0
5.48
11.21
9.58

57.64
58.28
59.23
58.77

2.24
1.45
1.24
1.60

3.54
2.37
1.24
1.79

elemental analysis were recorded on a Quanta 200 SEM system


equipped with EDS detector (FEI company, Netherlands). Solidstate 29Si NMR spectrum of the catalyst sample was recorded on
an Innity Plus-300 MHz spectrometer. The basic property of the
catalysts was evaluated by the Hammett indicator method and
expressed by an acidity function (H_). The used indicators were
as follows: neutral red (H_ = 6.8), bromothymol blue (H_ = 7.2),
phenolphthalein (H_ = 9.8), 2,4-dinitroaniline (H_ = 15.0), and
4-nitroaniline (H_ = 18.4). Basicity of the catalysts was measured
by the method of Hammett indicator-benzene carboxylic acid
(0.02 M anhydrous methanol solution) titration.
2.3. Transesterication reaction
The transesterication reaction was carried out in a three-neck
glass batch reactor equipped with a condenser (with a glass stopper) and a thermocouple thermometer. The reactor (stirring rate
at 800 rpm) was initially lled with 30 g of palm oil, different
amounts of the catalysts and different volumes of anhydrous
methanol. All of the experiments were performed at atmospheric
pressure.

After the transesterication reaction, the catalysts were


removed from the mixture by centrifugation. The ltrate was kept
in the funnel separator for 24 h. The upper layer was subjected to
rotary evaporation to remove excess methanol and the water of the
product obtained was removed by sodium sulfate before Gas chromatography analysis.
The amount of fatty acids methyl ester (FAME) obtained
through transesterication were analyzed by gas chromatography
(Agilent 7890A), equipped with a ame ionization detector and a
capillary column (DB-WAX, 30 m  0.25 mm  0.25 lm). Methyl
heptadecanoate was used as an internal standard to quantify the
content of biodiesel. The biodiesel yield was calculated by the following expression:

Yield weight of biodiesel  % FAME=weight of oil  100%


where % FAME is the concentration of FAME obtained by GC
analysis.

2.4. Catalyst reusability


After completing the reaction, the catalysts were ltered and
washed thoroughly with tetrahydrofuran. Subsequently, the
washed catalysts were dried in an oven at 110 C for 12 h. The
recovered catalysts were used for transesterication of palm oil
under similar reaction conditions to measure its activity in a
number of cycles.

871

1470

(a)

Transmittance (%)

(b)
991

(c)
(d)
(e)
1050

20%

2000

1800

1600

1400

1200

1000

800

600

Wavenumber (cm-1)

Fig. 2.

29

SiNMR spectrum of the 2Si5Ca catalyst.

Fig. 3. FTIR spectra of the (a) 0Si5Ca, (b) 1Si5Ca, (c) 2Si5Ca, (d) 3Si5Ca and (e)
5Si5Ca catalysts.

51

G. Chen et al. / Fuel 153 (2015) 4855

Intensity (a.u.)

5Si5Ca

3Si5Ca

2Si5Ca

1Si5Ca

0Si5Ca

20

30

40

50

60

70

80

2 ()
Fig. 4. XRD patterns of the 0Si5Ca, 1Si5Ca, 2Si5Ca, 3Si5Ca and 5Si5Ca catalysts.

3. Results and discussion


3.1. Characterization of catalysts
In this study, ve kinds of catalysts were synthesized with different Na2SiO3 concentrations (0, 0.2, 0.4, 0.6 and 1 M), which were
denoted as the 0Si5Ca, 1Si5Ca, 2Si5Ca, 3Si5Ca and 5Si5Ca, respectively. The 0Si5Ca and 2Si5Ca catalysts were rstly taken as examples for illustrating the successful generation of CaOSiO2
catalysts. Fig. 1a and b showed the SEM micrographs of these
two catalysts. As illustration, small differences were found
between the two samples. In comparison to the tiny particle of
the 0Si5Ca catalyst, the 2Si5Ca catalyst exhibited larger and more
regular blocks, which could be owing to the coverage of Si compounds on the CaO surface. The EDS curves of the 0Si5Ca and
2Si5Ca catalysts were also acquired and several typical points on
the surface of the samples were selected in order to determine
the elemental contents of Ca, Si, O and C. The results were summarized in Table 1. It was observed that the Si content of the three
points (A, B and C) ranged from 5.48% to 11.21% for the sample
in Fig. 1b, indicating a relative high content of Si content on the
surface of the 2Si5Ca catalyst. To further investigate the chemical
structure of Si compound at the atomic scale for the composite catalysts, 29Si NMR of the 2Si5Ca catalyst was conducted as an example. As shown in Fig. 2, the three peaks at 86.5, 100.1, and
114.3 ppm were generally assigned to Q2 [Si(OSi)2(OH)2], Q3
[Si(OSi)3(OH)], and Q4 [Si(OSi)4] with relative percentages of 7%,
87%, and 6%, respectively. The high content of Q3 and Q4 adding
up to 93% indicated a fairly high proportion of condensed silanols,
revealing that lysozyme existed within the eggshell helped to form
well-condensed silica networks.
The FTIR spectra of the 0Si5Ca, 1Si5Ca, 2Si5Ca, 3Si5Ca and
5Si5Ca catalysts were presented in Fig. 3. The absorption band at

Table 2
Basic strengths and surface area of series of the catalysts.

871 cm1 that was the out-of-plane band vibration mode of the
CO2
3 groups could be found. Accordingly, another absorption band
that was ascribed to the asymmetric stretch of CO2
3 groups was
also found at 1417 cm1 [36]. The appearance of this absorption
band with similar intensity for all the ve catalysts may be
attributed to the basic carbonates generated by the adsorption of
gaseous CO2 from atmosphere onto the catalysts [37].
Furthermore, the SiOCa bond (the adsorption band at around
991 cm1 [18]) on the catalysts surface appeared as the Na2SiO3
concentration reached 0.2 M, while the SiOSi bonds (the absorption band around 1050 cm1 [26]) appeared with the Na2SiO3 concentration reached 1.0 M. This was suggested that, in the lower
Na2SiO3 concentration, all the SiO2 generated from the biomimetic
silicication process was reacted with CaO to form the new phase
CaSiO3 or Ca2SiO4 compound. when the Na2SiO3 concentration was
too high, excessive SiO2 was found on the catalyst surface.
Fig. 4 showed the XRD patterns of the 0Si5Ca, 1Si5Ca, 2Si5Ca,
3Si5Ca and 5Si5Ca catalysts. The peaks those appeared at
2h = 37.2, 64.2, 76.1 and 2h = 20.2, 33.4, 39.7, 55.3, 59.8
were the characteristic peaks of CaO and Ca(OH)2, respectively.
When the Na2SiO3 concentration increased from 0 to 0.4 M, a mixture of CaO and Ca(OH)2 in the catalysts could be found. Besides,
with the increase of Na2SiO3 concentration, the peak intensity of
CaO and Ca(OH)2 decreased gradually. This might be primarily
due to the increased coverage ratio of Si compounds (SiO2 and
CaSiO3) on the CaO surface. Subsequently, as the Na2SiO3 concentration increased to 0.6 M, the peak of Si compounds could be
found and the peak intensity of CaO and Ca(OH)2 decreased sharply, suggesting that the occurrence of chemical reaction between
SiO2 and CaO/Ca(OH)2 phases in the catalyst. As the Na2SiO3 concentration increased to 1.0 M, the peak of SiO2 (2h = 26.8 and
42.2), CaSiO3 (2h = 39.2) and Ca2SiO4 (2h = 35.1) could be then
observed, while CaO and Ca (OH)2 phases completely disappeared.
Interestingly, no CaCO3 peak could be found on the catalyst surface
for all the catalysts. This result seemed to be contrast with the FTIR
and EDS analysis (only 1.242.24 wt.% content in catalyst surface)
results as described above. The probable explanation could be
described as follows: the content of CaCO3 in the catalysts was
too low to be detected by the XRD instrument. Similar results were
also found in some previous literatures [18,38,39].

100

80

Biodiesel yield (%)

-- CaO
-- Ca(OH)2
-- CaSiO3
-- Ca2SiO4
-- SiO2

0Si5Ca
1Si5Ca
2Si5Ca
3Si5Ca
5Si5Ca

60

40

20

Catalyst

Basic strength
(H_)

Total basicity
(mmol g1)

Surface area
(m2 g1)

Biodiesel
yield (%)

0Si5Ca
1Si5Ca
2Si5Ca
3Si5Ca
5Si5Ca

15.0 < H_ < 18.4


15.0 < H_ < 18.4
9.8 < H_ < 15.0
7.2 < H_ < 9.8
6.8 < H_ < 7.2

9.3
8.0
5.5
0.4
0.1

15.47
8.67
5.78
4.29
4.01

90.2
87.7
80.1
40.7
13.5

10

Reaction time (h)


Fig. 5. Effect of reaction time on the biodiesel yield for the catalysts. Conditions for
0Si5Ca: methanol-to-oil molar ratio, 9; catalyst loading, 5 wt.%; temperature, 65 C.
Conditions for 1Si5Ca: methanol-to-oil molar ratio, 12; catalyst loading, 7 wt.%;
temperature, 65 C. Conditions for 2Si5Ca: methanol-to-oil molar ratio, 15; catalyst
loading, 9 wt.%; temperature, 65 C. Conditions for 3Si5Ca and 5Si5Ca: methanolto-oil molar ratio, 21; catalyst loading, 11 wt.%; temperature, 65 C.

52

G. Chen et al. / Fuel 153 (2015) 4855

be as a result of the active site of CaO surface being covered


by Si compounds. Among those catalysts, the base strength of the
0Si5Ca, 1Si5Ca and 2Si5Ca catalysts was 15.0 < H_ < 18.4,
15.0 < H_ < 18.4 and 9.8 < H_ < 15.0, making them strong base catalysts for conducting the transesterication reaction [16].

100
90
80

0Si5Ca
1Si5Ca
2Si5Ca
3Si5Ca
5Si5Ca

60
50

3.2. Optimization of reaction conditions on the transesterication of


palm oil

40
30
20
10
0
50

55

60

65

70

Reaction temperature (oC)


Fig. 6. Effect of reaction temperature on the biodiesel yield for the catalysts.
Conditions for 0Si5Ca: methanol-to-oil molar ratio, 9; catalyst loading, 5 wt.%;
reaction time, 3 h. Conditions for 1Si5Ca: methanol-to-oil molar ratio, 12; catalyst
loading, 7 wt.%; reaction time, 4 h. Conditions for 2Si5Ca: methanol-to-oil molar
ratio, 15; catalyst loading, 9 wt.%; reaction time, 8 h. Conditions for 3Si5Ca and
5Si5Ca: methanol-to-oil molar ratio, 21; catalyst loading, 11 wt.%; temperature,
65 C.

Biodiesel yield (%)

Table 2 showed the surface area of the CaOSiO2 catalysts.


Specically, the surface area of the catalysts was reduced (from
15.47 to 4.01 m2 g1) as the Na2SiO3 concentration increased (from
0 to 1.0 M). This decrease in surface area (similar phenomenon could
be observed by other CaO based catalysts [40,41]) may be ascribed to
the coverage of Si compounds on the CaO surface [40,41].
Meanwhile, Table 2 also showed the base strength and total
basicity of the catalysts evaluated by the Hammett indicator
method. The total basicity of the catalysts slightly decreased from
9.3 to 5.5 mmol g1 with the increase of Na2SiO3 concentration
from 0.2 to 0.4 M, then decreased signicantly (from 5.5 to
0.1 mmol g1) when the amount of Na2SiO3 concentration
increased from 0.6 to 1.0 M. The decrease in total basicity could

3.2.1. Effect of reaction time


The reaction time is known to be one of the most important
parameters affecting the biodiesel yield. As illustrated in Fig. 5,
the 0Si5Ca catalyst (directly derived from calcinated eggshell without loading Si) exhibited the highest catalytic activity with biodiesel yield of 90.2% after 3 h-reaction. This result was comparable to
the level that reported in the recent relevant literatures [17,39].
With the increase of Si compounds in catalyst, the catalytic activity
decreased gradually (the biodiesel yield of the catalysts decreased
from 90.2% to 13.5%). Nevertheless, the 1Si5Ca and 2Si5Ca catalysts
(the maximum biodiesel yields were 87.7% and 80.1% and optimum reaction time was 4 h and 8 h, respectively) showed minor
differences in the catalytic activity with the 0Si5Ca catalyst. In case
of other catalysts, much longer reaction time was required for
reaching the reaction equilibrium. Particularly, after 10 h-reaction,
the 5Si5Ca catalyst with the highest amount of Si compounds only
possessed a biodiesel yield of 13.5% (the lowest catalytic activity).
This was mainly due to the rather low basicity of this catalyst (as
illustrated in Table 2).
3.2.2. Effect of reaction temperature
Reaction temperature is also one of the factors affecting the
transesterication reaction. The catalytic activities of the catalysts
were investigated with the reaction temperature varying from 50
to 70 C. When the reaction was performed at a temperature
higher than 65 C (as the boiling point of methanol is 64.55 C), a
condenser (with a glass stopper) was attached to the side neck of
three-neck glass batch reactor for reuxing methanol. The reaction
system was sealed and no methanol could be escaped from the
reaction system. In Fig. 6, similar trends were found for all the ve

100

100

90

90

80

80

70

0Si5Ca
1Si5Ca
2Si5Ca
3Si5Ca
5Si5Ca

60
50
40

Biodiesel yield (%)

Biodiesel yield (%)

70

70

0Si5Ca
1Si5Ca
2Si5Ca
3Si5Ca
5Si5Ca

60
50
40

30

30

20

20

10

10
0

0
6

12

15

18

21

Methanol-to-oil molar ratio


Fig. 7. Effect of methanol-to-oil molar ratio on the biodiesel yield for the catalysts.
Conditions for 0Si5Ca: temperature, 65 C; catalyst loading, 5 wt.%; reaction time,
3 h. Conditions for 1Si5Ca: temperature, 65 C; catalyst loading, 7 wt.%; reaction
time, 4 h. Conditions for 2Si5Ca: temperature, 65 C; catalyst loading, 9 wt.%;
reaction time, 8 h. Conditions for 3Si5Ca and 5Si5Ca: temperature, 65 C; catalyst
loading, 11 wt.%; reaction time, 10 h.

10

11

Catalyst loading (wt.%)


Fig. 8. Effect of catalyst loading on the biodiesel yield for the catalysts. Conditions
for 0Si5Ca: temperature, 65 C; methanol-to-oil molar ratio, 9; reaction time, 3 h.
Conditions for 1Si5Ca: temperature, 65 C; methanol-to-oil molar ratio, 12; reaction
time, 4 h. Conditions for 2Si5Ca: temperature, 65 C; methanol-to-oil molar ratio,
15; reaction time, 8 h. Conditions for 3Si5Ca and 5Si5Ca: temperature, 65 C;
methanol-to-oil molar ratio, 21; reaction time, 10 h.

53

G. Chen et al. / Fuel 153 (2015) 4855

3.2.3. Effect of methanol-to-oil molar ratio


The stoichiometric methanol-to-oil molar ratio for the transesterication reaction is 3:1. Since the transesterication reaction is
reversible, excessive methanol is needed to drive the reaction forwards as far as possible. The effect of methanol-to-oil molar ratio
on the biodiesel yield was illustrated in Fig. 7. Seen from this gure, when the methanol-to-oil molar ratio increased from 3:1 to
21:1, the biodiesel yields of the catalysts increased gradually until
the maximum values were acquired. Correspondingly, the optimum methanol-to-oil molar ratios were 9:1, 12:1, 15:1, 21:1 and
21:1 for the 0Si5Ca, 1Si5Ca, 2Si5Ca, 3Si5Ca and 5Si5Ca catalysts,
respectively. Further increase in the methanol-to-oil molar ratio
beyond its optimum value would decrease the biodiesel yield,
which was primarily owing to the following aspect: the excessive
methanol made it difcult to separate the biodiesel with glycerol.

90

0Si5Ca
1Si5Ca
2Si5Ca

Biodiesel yield (%)

80

70

60

50

40
0

10

12

Number of cycle
Fig. 9. Reusability study of the 0Si5Ca, 1Si5Ca and 2Si5Ca catalysts. Conditions for
0Si5Ca: reaction time, 3 h; methanol-to-oil molar ratio, 9; catalyst loading, 5 wt.%;
temperature, 65 C. Conditions for 1Si5Ca: reaction time, 4 h; methanol-to-oil
molar ratio, 12; catalyst loading, 7 wt.%; temperature, 65 C. Conditions for 2Si5Ca:
reaction time, 8 h; methanol-to-oil molar ratio, 15; catalyst loading, 9 wt.%;
temperature, 65 C.

catalysts. In general, as the temperature rose from 50 to 65 C, the


biodiesel yields of the ve catalysts signicantly increased. More
specically, when temperature reached 65 C, all the reactions
attained equilibrium and the maximum biodiesel yields of the
0Si5Ca, 1Si5Ca, 2Si5Ca, 3Si5Ca and 5Si5Ca catalysts were 90.2%,
87.7%, 80.1%, 40.7% and 13.5%, respectively. However, when the
temperature was beyond 65 C, the biodiesel yields of the catalysts
slightly decreased (At 70 C, the biodiesel yields of the 0Si5Ca,
1Si5Ca, 2Si5Ca, 3Si5Ca and 5Si5Ca catalysts were 84.4%, 85.2%,
75.1%, 38.7% and 11.2%, respectively). At lower reaction temperature of 5055 C, the much decreased catalytic activities of
the catalysts was due to the higher mass transfer resistance and
solubility limitations [41], while the decreased biodiesel yield at
higher temperature (70 C) was probably caused by its similarity
with the boiling point of methanol (65 C). More specically,
methanol kept boiling and caused the reaction to move backwards,
decreasing the biodiesel yield. Thus, 65 C was the optimum
temperature for the three catalysts.

3500

(a)

3000

3.2.4. Effect of catalyst loading


The effect of the catalyst loading on the biodiesel yield was also
investigated by varying the catalyst mass ratio from 3 to 10 wt.%
(Fig. 8). The biodiesel yield exhibited a gradually increased trend
as the catalyst loading increased. In detail, the 0Si5Ca, 1Si5Ca,
2Si5Ca, 3Si5Ca and 5Si5Ca catalysts showed equilibrium biodiesel
yields when the catalyst loading was 5 wt.%, 7 wt.%, 9 wt.%, 11 wt.%
and 11 wt.%, respectively. It was recognized that the catalyst activity was extremely inuenced by the basicity of the catalysts. The
catalyst possessing lower basicity required more amount of catalysts to reach the reaction equilibrium. However, further increase
in the catalyst loading beyond its optimum value, a decreased yield
was observed because of the difculty in blending of the catalyst,
reactant and product caused by the excessive applied catalyst.
Since the activity of 3Si5Ca and 5Si5Ca catalysts were rather low
as illustrated above, further investigations would only focus on
the other three catalysts, the 0Si5Ca, 1Si5Ca and 2Si5Ca catalysts.
3.3. Catalyst reusability
Besides the catalyst activity, the reusability study of the 0Si5Ca,
1Si5Ca and 2Si5Ca catalysts (with minor differences in the catalytic
activity) was also investigated for consecutive 12 cycles under the
optimum condition. Fig. 9 indicated that the 0Si5Ca catalyst in the
absence of Si compound showed poor reusability (after 12
consecutive cycles, the biodiesel yield dropped to 43%). When the
amount Si compound increased, the reusability of the catalysts

1Si5Ca
0Si5Ca
2Si5Ca

2500

--Ca(OH)2

2000

--Ca-glyceroxide

1500
500
0
1

200

10

(b)

11

12

1Si5Ca
0Si5Ca
2Si5Ca

150

Intensity (a.u.)

Ca content (ppm)

1000

(a)

100

(b)
50
0
1

10

11

12

Number of cycle
Fig. 10. Leaching content of Ca

2+

ions in (a) glycerol and (b) biodiesel phase.

10

15

20

25

30

35

40

2 ()
Fig. 11. XRD patterns of recovered (a) 0Si5Ca and (b) 2Si5Ca catalysts.

54

G. Chen et al. / Fuel 153 (2015) 4855

CaO to the bulk reaction medium, thus preventing not only the
leaching of Ca2+ ions in some extent but also the unnecessary
reaction between CaO and glycerol (or H2O) [26,42,43]; and (2)
the leached Ca2+ ions from the catalysts could be adsorbed by
SiO2 on the catalyst surface to form CaOSi bond (as shown
in Fig. 3), which could also enhance the catalyst stability. As
was well known, once applying CaO as a catalyst, both the
homogeneous system and the heterogeneous system contributed
to activity in the reaction [17], and if the leaching problem was
well solved, the activity of catalyst may decrease, but the stability could be improved. For practical applications in the transesterication reaction, it was important to acquire a more stable
catalyst rather than a more active one [26].

90
80

Biodiesel yield (%)

70

Palm Oil
1 wt.% water
3 wt.% water
1 wt.% FFA
3 wt.% FFA

60
50
40
30
20
10

3.4. Effect of FFA and water content


0
2

10

Reaction time (h)


Fig. 12. Effect of FFA and water content. Reaction conditions for 2Si5Ca: reaction
time, 8 h; methanol-to-oil molar ratio, 15; catalyst loading, 9 wt.%; temperature,
65 C.

was improved. Although the catalytic activity of the 2Si5Ca catalyst was lower than that of the 0Si5Ca catalyst, little deactivation
could be found up to 12 consecutive cycles. Additionally, the catalytic activity of the 2Si5Ca catalyst was higher than that of the
0Si5Ca catalyst after conducting the reaction for more than four
cycles. This indicated that the CaOSiO2 catalyst prepared through
the biomimetic silicication approach had a much better
reusability.
Moreover, the biodiesel phase after the removal of the catalysts
was measured by ICP analysis to detect the leached Ca2+ ions. As
shown in Fig. 10, the Ca2+ ions were indeed leached into the biodiesel and glycerol phase after each reaction cycle. Typically, the highest amount of leached Ca2+ ions in either biodiesel phase or
glycerol phase could be detected for the rst reaction cycle. The
concentrations of the dissolved Ca2+ ions in biodiesel phase for
the 0Si5Ca, 1Si5Ca and 2Si5Ca catalysts were 168, 108 and
40 ppm, respectively. In glycerol phase, the concentration of dissolved Ca2+ ions for the three catalysts was 3240, 2514 and
2001 ppm, respectively. However, the concentration of Ca2+ ions
for the three catalysts decreased signicantly with subsequent
reaction cycles, becoming lower than 5 ppm (in biodiesel phase)
and 100 ppm (in glycerol phase) after the 12th cycle. And the order
of Ca2+ content leaching in the reaction media for the catalysts was
as follows: 0Si5Ca > 1Si5Ca > 2Si5Ca.
Additionally, in order to investigate the chemical composition
revolution of catalysts before and after use, the XRD patterns of
recovered catalysts (0Si5Ca and 2Si5Ca) after the 12th cycle were
studied. As illustrated in Fig. 11, no peaks of CaO were found in
the XRD patterns of the recovered 0Si5Ca and 2Si5Ca catalysts,
which was mainly owing to the formation of new phases such as
Ca(OH)2 (2h = 18 and 34) and calcium diglyceroxide (2h = 8,
10, 21, 24 and 26) on the catalyst surface. It was notable that
the intensities of calcium diglyceroxide and Ca(OH)2 peaks for
the 2Si5Ca catalyst were much lower than that of the 0Si5Ca catalyst. The result might be owing to the Si compounds on the 2Si5Ca
preventing the formation of the new phases calcium diglyceroxide
and Ca(OH)2 during reaction.
In light of the experiments mentioned above, the good stability of the 2Si5Ca catalyst may be owing to the following
two reasons: (1) the better distribution of Si compounds on
the CaO surface protected the direct exposure of Ca2+ ions or

Considering that the production of second generation biofuels


from low quality feedstock, often contained water and FFA, the
evaluation of the catalytic behavior in the transesterication of
vegetable oils containing water and FFA were also examined.
Specically, the palm oil containing FFA (palmitic acid, 1 wt.%
and 3 wt.%) or water (1 wt.% and 3 wt.%) was adopted in this study
for the transesterication reaction. The reaction was conducted
using the 2Si5Ca catalyst with reaction temperature of 65 C,
methanol-to-oil molar ratio of 15 and catalyst loading of 9 wt.%.
The results were shown in Fig. 12. Specically, with the amount
of water added from 0 to 1 wt.% and 1 to 3 wt.%, the maximum
of biodiesel yield decreased from 80% to 75% and 75% to 62%,
respectively. This suggested that the catalytic activity of the
2Si5Ca catalyst was inuenced by the addition of water due to
the hydrolysis of palm oil with water. The similar phenomenon
was observed by the addition of FFA to palm oil. When the amount
of FFA added from 0 to 1 wt.% and 1 to 3 wt.%, the maximum of biodiesel yield decreased from 80% to 72% and 72% to 56%, respectively. The loss of catalytic activity of the 2Si5Ca catalyst was mainly
owing to the neutralization of the base sites on the surface of the
catalyst by FFA [44]. Note that the order of catalytic activity on
the 2Si5Ca catalyst with different feed stocks was as follows: palm
oil > 1 wt.% water > 1 wt.% FFA > 3 wt.% water > 3 wt.% FFA suggesting that FFA has more obvious inuence on the decrease of catalytic activity.

4. Conclusions
In summary, the CaOSiO2 catalysts were prepared through a
biomimetic silicication approach and exploited as novel solid
base catalysts for biodiesel production. Along with the increase
of the amount of Si compounds in the CaOSiO2 catalysts, the catalytic activity was gradually decreased, while the reusability was
signicantly improved. Particularly, in comparison to the 0Si5Ca
catalyst, the 2Si5Ca catalyst only showed a slightly lower catalytic
activity (biodiesel yield: 80.1% vs. 90.2%), however, little deactivation
could be found after 12 cycles reuse during the transesterication
reaction. Hopefully, the CaOSiO2 catalyst prepared through this
biomimetic silicication approach will be a promising solid base
catalyst for biodiesel or other fuels/chemicals production.
Acknowledgements
This paper is nancially supported by National Natural Science
Funds of China (21406163), National High-tech Research and
Development Program (863 Program) through project
(2012AA051801).

G. Chen et al. / Fuel 153 (2015) 4855

References
[1] Yan Y, Li X, Wang G, Gui X, Li G, Su F, Wang X, Liu T. Biotechnological
preparation of biodiesel and its high-valued derivatives: a review. Appl Energy
2014;113:161431.
[2] Canakci M. The potential of restaurant waste lipids as biodiesel feedstocks.
Bioresour Technol 2007;98:18390.
[3] Aransiola EF, Ojumu TV, Oyekola OO, Madzimbamuto TF, Ikhu-Omoregbe DIO.
A review of current technology for biodiesel production: state of the art.
Biomass Bioenergy 2014;61:27697.
[4] Zhang Y. Biodiesel production from waste cooking oil: 1. Process design and
technological assessment. Bioresour Technol 2003;89:116.
[5] Marchetti JM, Miguel VU, Errazu AF. Possible methods for biodiesel production.
Renew Sustain Energy Rev 2007;11:130011.
[6] Kawashima A, Matsubara K, Honda K. Development of heterogeneous base
catalysts for biodiesel production. Bioresour Technol 2008;99:343943.
[7] Thitsartarn W, Kawi S. An active and stable CaOCeO2 catalyst for
transesterication of oil to biodiesel. Green Chem 2011;13:342330.
[8] Semwal S, Arora AK, Badoni RP, Tuli DK. Biodiesel production using
heterogeneous catalysts. Bioresour Technol 2011;102:215161.
[9] Liu Q, Wang B, Wang C, Tian Z, Qu W, Ma H, Xu R. Basicities and
transesterication activities of ZnAl hydrotalcites-derived solid bases.
Green Chem 2014;16:260413.
[10] Helwani Z, Othman MR, Aziz N, Kim J, Fernando WJN. Solid heterogeneous
catalysts for transesterication of triglycerides with methanol: a review. Appl
Catal A 2009;363:110.
[11] Jiang Y, Liu X, Chen Y, Zhou L, He Y, Ma L, Gao J. Pickering emulsion stabilized
by lipase-containing periodic mesoporous organosilica particles: a robust
biocatalyst
system
for
biodiesel
production.
Bioresour
Technol
2014;153:27883.
[12] Christopher LP, Hemanathan K, Zambare VP. Enzymatic biodiesel: challenges
and opportunities. Appl Energy 2014;119:497520.
[13] Gao J, Shi L, Jiang Y, Zhou L, He Y. Formation of lipase Candida sp. 99125
CLEAs in mesoporous silica: characterization and catalytic properties. Catal.
Sci Technol. 2013;3:33539.
[14] Kawashima A, Matsubara K, Honda K. Acceleration of catalytic activity of
calcium oxide for biodiesel production. Bioresour Technol 2009;100:696700.
[15] Wan Z, Hameed BH. Transesterication of palm oil to methyl ester on
activated carbon supported calcium oxide catalyst. Bioresour Technol
2011;102:265964.
[16] Boey PL, Maniam GP, Hamid SA, Ali DMH. Utilization of waste cockle shell
(Anadara granosa) in biodiesel production from palm olein: optimization using
response surface methodology. Fuel 2011;90:23538.
[17] Kouzu M, Hidaka JS. Transesterication of vegetable oil into biodiesel
catalyzed by CaO: a review. Fuel 2012;93:112.
[18] Chen GY, Shan R, Shi JF, Yan BB. Transesterication of palm oil to biodiesel
using rice husk ash-based catalysts. Fuel Process Technol 2015;133:813.
[19] Witoon T, Bumrungsalee S, Vathavanichkul P, Palitsakun S, Saisriyoot M,
Faungnawakij K. Biodiesel production from transesterication of palm oil with
methanol over CaO supported on bimodal meso-macroporous silica catalyst.
Bioresour Technol 2014;156:32934.
[20] Granados ML, Alonso DM, Sdaba I, Mariscal R, Ocn P. Leaching and
homogeneous contribution in liquid phase reaction catalysed by solids: the
case of triglycerides methanolysis using CaO. Appl Catal B 2009;89:26572.
[21] Albuquerque MCG, Azevedo DCS, Cavalcante CL, Santamara-Gonzlez J,
Mrida-Robles JM, Moreno-Tost R, Rodrguez-Castelln E, Jimnez-Lpez A,
Maireles-Torres P. Transesterication of ethyl butyrate with methanol using
MgO/CaO catalysts. J Mol Catal A 2009;300:1924.
[22] Rubio-Caballero JM, Santamara-Gonzlez J, Mrida-Robles J, Moreno-Tost R,
Jimnez-Lpez A, Maireles-Torres P. Calcium zincate as precursor of active
catalysts for biodiesel production under mild conditions. Appl Catal B
2009;91:33946.

55

[23] Kumar D, Ali A. Transesterication of low-quality triglycerides over a Zn/CaO


heterogeneous catalyst: kinetics and reusability studies. Energy Fuel
2013;27:375868.
[24] Ngamcharussrivichai C, Totarat P, Bunyakiat K. Ca and Zn mixed oxide as a
heterogeneous base catalyst for transesterication of palm kernel oil. Appl
Catal A 2008;341:7785.
[25] Xie W, Zhao L. Heterogeneous CaO-MoO3-SBA-15 catalysts for biodiesel
production from soybean oil. Energy Convers Manage 2014;79:3442.
[26] Sun H, Han J, Ding Y, Li W, Duan J, Chen P, Lou H, Zheng X. One-pot synthesized
mesoporous Ca/SBA-15 solid base for transesterication of sunower oil with
methanol. Appl Catal A 2010;390:2634.
[27] Nassif N, Livage J. From diatoms to silica-based biohybrids. Chem Soc Rev
2011;40:84959.
[28] Patwardhan SV. Biomimetic and bioinspired silica: recent developments and
applications. Chem Commun (Camb) 2011;47:756782.
[29] Forsyth C, Yip TW, Patwardhan SV. CO2 sequestration by enzyme immobilized
onto bioinspired silica. Chem Commun (Camb) 2013;49:31913.
[30] Steven CR, Busby GA, Mather C, Tariq B, Briuglia ML, Lamprou DA, Urquhart AJ,
Grant MH, Patwardhan SV. Bioinspired silica as drug delivery systems and
their biocompatibility. J Mater Chem B 2014;2:502842.
[31] Ramanathan M, Luckarift HR, Sarsenova A, Wild JR, Ramanculov EK, Olsen EV,
Simonian AL. Lysozyme-mediated formation of protein-silica nano-composites
for biosensing applications. Colloid Surface B 2009;73:5864.
[32] Betancor L, Luckarift HR. Bioinspired enzyme encapsulation for biocatalysis.
Trends Biotechnol 2008;26:56672.
[33] Garakani TM, Wang H, Krappitz T, Liebeck BM, van Rijn P, Boker A. Lysozymesilica hybrid materials: from nanoparticles to capsules and double emulsion
mineral capsules. Chem Commun (Camb) 2012;48:102102.
[34] Liu C, Yang D, Jiao Y, Tian Y, Wang Y, Jiang Z. Biomimetic synthesis of TiO2
SiO2Ag nanocomposites with enhanced visible-light photocatalytic activity.
ACS Appl Mater Interfaces 2013;5:382432.
[35] Coradin T, Coup A, Livage J. Interactions of bovine serum albumin and
lysozyme with sodium silicate solutions. Colloid Surface B 2003;29:18996.
[36] Boro J, Thakur AJ, Deka D. Solid oxide derived from waste shells of Turbonilla
striatula as a renewable catalyst for biodiesel production. Fuel Process Technol
2011;92:20617.
[37] Yaakob Z, Sukarman IS, Narayanan B, Abdullah SR, Ismail M. Utilization of
palm empty fruit bunch for the production of biodiesel from Jatropha curcas
oil. Bioresour Technol 2012;104:695700.
[38] Girish N, Niju SP, Meera Sheriffa Begum KM, Anantharaman N. Utilization of a
cost effective solid catalyst derived from natural white bivalve clam shell for
transesterication of waste frying oil. Fuel 2013;111:6538.
[39] Correia LM, Saboya RM, Campelo Nde S, Cecilia JA, Rodriguez-Castellon E,
Cavalcante CL, Vieira RS. Characterization of calcium oxide catalysts from
natural sources and their application in the transesterication of sunower oil.
Bioresour Technol 2014;151:20713.
[40] Atadashi IM, Aroua MK, Aziz AA. Biodiesel separation and purication: a
review. Renew Energy 2011;36:43743.
[41] Boro J, Konwar LJ, Thakur AJ, Deka D. Ba doped CaO derived from waste shells
of T striatula (TS-CaO) as heterogeneous catalyst for biodiesel production. Fuel
2014;129:1827.
[42] Sun H, Duan J, Chen P, Lou H, Zheng X. Room temperature transesterication of
soybean oil to biodiesel catalyzed by rod-like CaxSiOx+2 solid base. Catal
Commun 2011;12:10058.
[43] Jain D, Khatri C, Rani A. Fly ash supported calcium oxide as recyclable solid
base catalyst for Knoevenagel condensation reaction. Fuel Process Technol
2010;91:101521.
[44] Pasupulety N, Rempel GL, Ng FTT. Studies on MgZn mixed oxide catalyst for
biodiesel production. Appl Catal A 2015;489:7785.

You might also like