Anda di halaman 1dari 14

Energy Conversion and Management 88 (2014) 484497

Contents lists available at ScienceDirect

Energy Conversion and Management


journal homepage: www.elsevier.com/locate/enconman

A review on the performance of glycerol carbonate production via


catalytic transesterication: Effects of inuencing parameters
Wai Keng Teng, Gek Cheng Ngoh , Rozita Yusoff, Mohamed Kheireddine Aroua
Department of Chemical Engineering, Faculty of Engineering, University of Malaya, 50603 Kuala Lumpur, Malaysia

a r t i c l e

i n f o

Article history:
Received 6 June 2014
Accepted 15 August 2014
Available online 18 September 2014
Keywords:
Glycerol
Crude glycerol
Biodiesel
Glycerol carbonate
Catalytic transesterication
Hydrotalcite

a b s t r a c t
Driven by high energy demand and environmental concerns, biodiesel as a substitute for fossil fuels is
recognized to be promising renewable and clean energy. The increase in the biodiesel plant dramatically
leads to the oversupply of its by-product glycerol in the biodiesel industries. Developing new industrial
uses for glycerol is essential to increase the net energy and sustainability of biodiesel. Moreover, glycerol
has great potential to be converted into marketable and valuable chemicals. The conversion of glycerol to
glycerol carbonate (GC) has been extensively studied and transesterication of glycerol to GC has been
proven to be the most promising route. Aimed to reveal the underlying mechanism of this successful conversion path, this paper reviews the chemo- and biocatalytic transesterication of glycerol with different
carbonates sources. Also, a detail elucidation of the inuence of the catalysts and operating conditions on
the GC yield is included to provide an insight into the process. In addition, the future direction of glycerol
carbonate production via catalytic transesterication is provided in this review.
2014 Elsevier Ltd. All rights reserved.

1. Introduction
Biodiesel industry is booming attributing to the threat from
petroleum depletion since the last decade [1,2]. Glycerol is generated at the proportion of 10% (w/w) of the total biodiesel production [3] and the rapid growing industry has led to a surplus of
glycerol. For instance, the forecasted world biodiesel supply by
2016 will be reaching 37 billion gallons with 4 billion gallons of
crude glycerol production [3]. The abundance of crude glycerol is
expected to exert a great impact on the rened glycerol market.
As a consequence, the price of glycerol has been dropped dramatically since 2006 [4]. In late 2013, the price of rened glycerol was
around 900965 US$/ton depending on the biodiesel feedstock
while crude glycerol with purity 80% obtained directly from biodiesel plant had been reported at lower value of 240 US$/ton in
mid 2014 [5]. This has prompted the conversion of the low cost
glycerol to value-added products.
Inevitably, the paradigm shift has attracted much attention of
researchers to explore the possibilities of converting glycerol to
value-added chemicals such as fuel, chemical intermediate and
chemicals. The attempts made include the transformation of glycerol to 1,3-propanediol, epichlorohydrin, acrolein, fuel additive,

Corresponding author. Tel.: +60 3 79675301; fax: +60 3 79675371.


E-mail addresses: tengwaikeng@yahoo.com (W.K. Teng), ngoh@um.edu.my (G.C.
Ngoh), ryusoff@um.edu.my (R. Yusoff), mk_aroua@um.edu.my (M.K. Aroua).
http://dx.doi.org/10.1016/j.enconman.2014.08.036
0196-8904/ 2014 Elsevier Ltd. All rights reserved.

glycerol carbonate (GC) and glycidol [6]. One of the most celebrated products reported in the last 5 years is GC. It is a high
value-added product with market price greater than 8141 US$/
ton [7]. Due to the high cost, it is still not widely used in commercial application [8,9]. A limited usage of only a few kt per year was
reported [7]. On the other hand, GC can be produced from biogenic
glycerol [10] and has great potential to be used as substitution for
petro-derivative compounds [11]. The potential industrial uses of
GC are presented in Fig. 1 [8,1114].
GC can be synthesized from different routes by using glycerol as
alcohol OH-source and chemicals such as CO/O2, organic carbonate,
urea or carbon dioxide as carbonate source [8,12]. Among the
routes for GC synthesis, transesterication of glycerol with
dimethyl carbonate (DMC) is one of the most direct and industrial
feasible pathways to produce high GC yield [12]. This synthesis
route will be systematically discussed.

2. Transesterication of glycerol
Transesterication is the carbonate exchange reaction between
alcohols and carbonate sources [8]. Glycerol, also known as glycerin, glycerin, or 1,2,3-propanetriol is the simplest trihedric alcohol. It is produced conventionally through saponication, fatty
alcohol plant and hydrolysis [15]. Utilization of glycerol to synthesize GC is possible either through direct or indirect synthetic route
as summarized in Fig. 2 [8,12]. Among the routes, the indirect

W.K. Teng et al. / Energy Conversion and Management 88 (2014) 484497

485

Fig. 1. The potential uses of glycerol carbonate in various industries [8,1114].

Fig. 2. Various glycerol carbonate synthesis routes [8,12].

route which involved phosgenation and transesterication of glycerol with CO2 derivatives such as alkylene carbonate and dialkyl
carbonate have favored high yield of GC. The carbon atom of the
carbonate group is nucleophilic which can be attacked by the oxygen atom of the glycerols hydroxyl group [8].
The earliest GC production method reported [16] used carbon
monoxide or phosgene with metallic catalysts. The process is hazardous due to toxicity of phosgene. A safer method involves a
much lower reaction temperature is the transesterication of glycerol with alkylene carbonate. Currently, GC is produced by chemical companies such as Huntsman from propylene carbonate in the

presence of catalyst at 100150 C and 35 mmHg [17]. The most


studied route for the synthesis of GC is via the environmentally
benign transesterication of glycerol with dialkyl carbonate. This
carbonate source can be prepared from methanol and urea and is
widely applied in the production of GC [12].
Catalyst plays a crucial role in transesterication of glycerol to
GC. A wide spectrum of catalysts with different property combination i.e. alkaline or acid, homogeneous or heterogeneous can be
employed. Table 1 summarizes the reaction conditions for various
types of catalysts and their advantages and limitations are
presented in Table 2 to exhibit the inuence of catalyst on GC

Type of catalysts

486

Table 1
Comparison of reaction conditions and performance of various catalysts in glycerol transesterication.
Performanceb

Reaction conditions
a

References

Molar ratio (carbonate:


glycerol)

Reaction time
(h)

Catalyst loading (wt or


mol%)

Solvent

Y/C/S (%)

3:1 (DMC)
3:1 (DMC)
5:1 (DMC)
10:1 (DMC)
3:1 (DEC)
5:1 (DMC)
5:1 (DMC)
4:1 (DMC)
3.5:1 (DMC)

3
5
1.5
48
4
1.5
1.5
2.5
0.33

4.5 wt%
4.5 wt%
15 wt%
4.5 wt%
3 mol%
6 wt%
4 wt%
10 mol%
2.6 mol%

Y = 97
N.A.
Y = 100
Y = 18c
Y = 97
Y = 100
Y = 98.5
Y = 98
Y = 95.7

[18]
[19,20]
[21]
[18]
[18]
[21]
[21]
[22]
[23]

1,3-Dichlorodistannoxanes
Ionic liquid tetra-n-butylammonium
Ionic liquid (BMIM-2-CO2)
Ionic liquid (BMIM-2-CO2)
Ionic liquid [Mor1,4][N(CN)2]
Ionic liquid ([TMA][OH])
Calcium complex Ca(C3H7O3)(OCO2CH3)
Calcium complex Ca(C3H7O3)(OCO2CH3)

7375
7176
75
7375
7375
75
75
6888
Room
temperature
100
120
74
74
120
80
75
75

5:1 (DEC)
2:1 (DMC)
3.2:1 (DMC)
3.2:1 (DMC)d
3:1 (DMC)
3:1 (DMC)
2:1 (DMC)
2:1 (DMC)

2
6
1.33
5
13
1.5
0.5
0.5

0.5 mol%
3.3 mol%
1 mol%
5 mol%
17 mol%
1 mol%
0.01 wt%
0.01 wt%

Y = 99.1
Y = 92
Y = 100
Y = 93
Y = 95
Y = 47e, C = 95
Y = 91.4
Y = 41.8

[24]
[25]
[26]
[26]
[27]
[28]
[29]
[29]

Homogeneous acid catalyst


H2SO4
p-Toluenesulfonic acid

75
75

5:1 (DMC)
5:1 (DMC)

1.5
1.5

10 mol%
10 mol%

Y = 3.5
Y = 4.3

[21]
[21]

Heterogeneous base catalyst


CaOf
CaO
CaO
CaOf
CaO
CaO
Calcium complex Ca(C3H7O3)2
CaCO3f
Na2O
ZnO
MgOf
MgO
n-Bu2Sn(OMe)2
Mg/La mixed oxides
Mg1 + xCa1 xO2 mixed oxides
Mg/Zr/Sr mixed oxides
Mg/Al/Zr mixed oxides
Al/Mg hydrotalcitef
Al/Mg hydrotalcitef (rehydrated)
Al/Mg hydrotalcitef
Al/Li hydrotalcitef
Al/Ca hydrotalcitef
Mg/Al hydrotalcite
Mg/Al hydrotalcite
Mg/Al hydrotalcite
Mg/Al hydrotalcite
Mg/Al hydrotalcite-hydromagnesium 0.1 g
Mg/Al hydrotalcite-hydromagnesium 0.5 g
Mg/Al hydrotalcitef

75
60
75
75
35
35
60
75
75
75
75
50
180
85
70
90
75
50
50
35
35
35
100
100
100
100
100
100
100

5:1 (DMC)
1:1 (DMC)
2:1 (DMC)
2:1 (DMC)
2:1 (EC)
2:1 (EC)
2.5:1 (DMC)
5:1 (DMC)
2:1 (DMC)
2:1 (DMC)
2:1 (DMC)
2:1 (EC)
1:1 (DMC)
2:1 (DMC)
2:1 (DMC)
5:1 (DMC)
5:1 (DMC)
2:1 (EC)
2:1 (EC)
2:1 (EC)
2:1 (EC)
2:1 (EC)
5:1 (DMC)
5:1 (DMC)
5:1 (DMC)
5:1 (DMC)
5:1 (DMC)
5:1 (DMC)
3:1 (DMC)

1.5
2
0.5
0.5
0.25
1
3
1.5
0.5
0.5
3
5
15
1
1.5
1.5
1.5
5
5
1
1
1
1
2
3
9g
1.16
9g
2

6 wt%
2 mol%
3 mol%
3 mol%
0.5 wt%
0.5 wt%
8 mol%
10 mol%
3 mol%
3 mol%
3 mol%
7 wt%
6 mol%
5 wt%
3 wt%
15 wt%
0.1 wt%
7 wt%
7 wt%
0.5 wt%
0.5 wt%
0.5 wt%
54 wt%
54 wt%
54 wt%
11 wt%
54 wt%
27 wt%
10 wt%

DMF
DMF

DMF
DMF

Y = 91.1
Y = 69
Y = 90.2
Y = 94
Y = 81
Y = 83
Y = 95
Y = 90.6
Y = 92.6
Y = 0.5
Y = 10.2
Y = 78
Y = 65
Y = 65
Y = 100
Y = 56e, C = 96
Y = 94
Y = 82
Y = 68
C = 57
Y = 85
Y = 87
Y = 75
Y = 99
Y = 82
Y = 98
Y = 88
Y = 71
Y = 66

[21]
[30]
[29]
[29]
[31]
[31]
[30]
[21]
[29]
[29]
[29]
[31]
[32]
[33]
[34]
[35]
[36]
[31]
[31]
[31]
[31]
[31]
[37]
[37]
[37]
[37]
[38]
[38]
[39]

Homogeneous base catalyst


K2CO3
K2CO3
K2CO3
K2CO3
K2CO3
KOH
NaOH
Triethylamine
N-heterocyclic carbenes

W.K. Teng et al. / Energy Conversion and Management 88 (2014) 484497

Temperature (C)

100
130
130

3:1 (DMC)
17:1 (DEC)
21:1 (DEC)

2
10
8

Mg/Al hydrotalcite supported on carbon nanoberf


Mg/Al hydrotalcite
Zeolite (NaY)f
KF/hydroxyapatite
NaOH/c-Al2O3
K2CO3/MgO
Extruded CaO-based/Al2O3
Ionic liquids (Tri-n-butylamine immobilized on mesoporous
MCM41)

130
140
70
78
78
80
80
80

17:1 (DEC)
5:1 (DEC)
3:1 (DMC)
2:1 (DMC)
2:1 (DMC)
2.5:1 (DMC)
3:1 (DMC)
2:1 (EC)

Heterogeneous acid catalyst


Amberlyst 131wet
Amberlyst 39wet

75
75

Enzymatic catalyst
C.A. lipase B immobilized on Novozym 435
C.A. lipase B immobilized on Novozym 435
C.A. lipase B immobilized on Novozym 435
C.A. lipase B immobilized on Novozym 435
C.A. lipase B immobilized on Novozym 435
A.N. lipase immobilized on magnetic nano particles

DMSO

Y = 55
Y = 65
Y = 84c

[40]
[41]
[42]

2
3
4
0.83
1
2
5
1.5

10 wt%
16 wt%
2.16 g/g (continuous
system)
16 wt%
54 wt%
10 wt%
3 wt%
3 wt%
1 wt%
15 mol%
9 wt%

DMF
Methanol

Y = 58c
Y = 77
Y = 80
Y = 99
Y = 97h
Y = 99
Y = 90.57
Y = 79h

[43]
[37]
[44]
[45]
[46]
[47]
[48]
[49]

5:1 (DMC)
5:1 (DMC)

1.5
1.5

10 mol%
10 mol%

Y = <5
Y = 6.2

[21]
[21]

60
70
60
50
60
60

1:1 (DMC)
10:1 (DMC)
2:1 (DMC)
3:1 (DMC)
1.5:1 (DMC)
10:1 (DMC)

30
48
48
12
14
6

54 wt%
5 wt%
75 g/L
5 wt%
54 wt%
28 wt%

THF

Acetonitrile
t-butanol
t-butanol

[50]
[51]
[52]
[53]
[54]
[55]

A.N. lipase (free enzyme)

60

10:1 (DMC)

12 wt%

A.N. lipase (cross-linked enzyme aggregates on magnetic


particles)
A.N. lipase immobilized on magnetic nano particles

60

10:1 (DMC)

28.6 wt%

Y = 88, C = 94
Y = 80, C = 93
Y = 96.25
Y = 95, C = nearly 100
C = 94.85
Y = 41.3, C = 48.6,
S = 85
Y = 59.3, C = 74,
S = 80.3
Y = 55, C = 61, S = 90

60

10:1 (DMC)i

5 wt%

[58]

A.N. lipase immobilized on magnetic nano particles

60

10:1 (DMC)

5 wt%

5 wt%

Y = 32.4h, C = 36,
S = 90
Y = 41.4h, C = 45,
S = 92
Y = 25.7h, C = 27,
S = 95

A.N. lipase immobilized on magnetic nano particles

60

10:1 (DMC)

[56]
[57]

[58]
[58]

Note: Y = yield of glycerol carbonate, C = conversion of glycerol, S = selectivity of glycerol carbonate, DMC = dimethyl carbonate, DEC = diethyl carbonate, EC = ethylene carbonate. C.A. = candida antarctica, A.N. = aspergillus niger.
a
(Amount of catalyst/amount of glycerol)  100%.
b
Y = (g glycerol carbonate produced/g glycerolinitial)  100%, C = (glycerolinitialglycerolresidual)/glycerolinitial  100%, S = Y/C.
c
By-product diglycerol tricarbonate was formed.
d
Crude glycerol obtained from industrial biodiesel plant was used.
e
By-product glycidol was formed.
f
Calcination of catalyst at 450 C 6 T 6 900 C for 3 h to overnight.
g
Reaction was scaled-up.
h
Yield is calculated from Y = C * S.
i
Crude glycerol obtained from transesterication of residual sun-ower oil was used.
j
Crude glycerol obtained from transesterication of crude sun-ower oil was used.

W.K. Teng et al. / Energy Conversion and Management 88 (2014) 484497

Mg/Al hydrotalcitef doped with nickel


Mg/Al hydrotalcite (rehydrated)
Mg/Al hydrotalcite supported on a-Al2O3f

487

488

W.K. Teng et al. / Energy Conversion and Management 88 (2014) 484497

Table 2
Merits and demerits of various types of catalyst used in transesterication of glycerol.
Types of catalyst
Homogeneous
base catalyst

Merits

Demerits

High catalytic activity

Catalyst difcult to be separated from product

Sensitive to water, CaO can deactivate in presence of water

Mixed metal oxides

Simple separation method of catalyst from


product
Catalyst can be easily recovered and recycled

Hydrotalcite

Catalyst are economical

Zeolite

High possibility to create in-house unique catalyst


by varying the composition

K2CO3
KOH
NaOH
Ionic liquid

Heterogeneous
base catalyst

CaO

Leaching of catalyst active sites could be happened, lead to product


contamination
Energy intensive process such as calcination of catalyst required to
achieve high yield

Catalyst with support


H2SO4

Catalyst are widely available and economical

Very low catalytic activity

p-Toluenesulfonic
acid

No diffusion problem in reaction mixture

Catalyst difcult to be separated from product

Heterogeneous
acid catalyst

Amberlyst ion
exchange resins

Catalyst can be regenerated and reused

Enzyme

Candida Antarctica
(Novozym 435)
Aspergillus Niger

Mild reaction temperature is sufcient to carry


out the reaction
Simple purication step of enzyme from product
Enzyme can be reused

Homogeneous
acid catalyst

Corrosion on reactor and pipelines could be happened


Extremely low catalytic activity due to mass transport limitation
Cause corrosion on reactor and pipelines

synthesis process. It has been reported that the acidic catalysts in


both homogeneous and heterogeneous forms are not suitable for
GC synthesis as their catalytic activities are extremely low [12].
GC yields achieved by acidic catalysts were less than 5% even with
the employment of strong acid such as sulfuric and p-Toluenesulfonic acids or Amberlyst ion exchange resins in the catalytic reaction. Acidic heterogeneous catalysts give poor reaction
performance due to the mass transport limitation. This could be
explained by the hydrophobic surface of the resins that hindered
the diffusion of hydrophilic reactants like DMC towards the catalytic sites inside the resins. On the other hand, acidic homogeneous
catalysts do not encounter the diffusion problem and they can
achieve a much improved GC yield of 50% after 24 h. The rate-controlling step which cyclizes the methyl glyceryl carbonate intermediate to GC has led to a slow GC formation rate [12]. Unlike the
acidic catalysts, both the homogeneous and heterogeneous strong
basic catalysts successfully produce more than 90% GC yield.
Apparently, the key criteria to ensure a viable transesterication reaction of glycerol lies in the efciency of the catalysts
applied. A vast range of base catalysts of either homogeneous or
heterogeneous form has proven to be effective in GC synthesis
despite individually possessing certain merits and demerits. The
following sections discuss the performance of various catalysts in
transesterication of glycerol in terms of product yield, catalyst
reusability and etc.
2.1. Homogeneous base catalyzed transesterication
Transesterication of glycerol had been successfully carried out
using homogeneous basic catalysts such as Na or K hydroxides,
carbonates or alkoxides with high catalytic activity. Similar to
the transesterication of oil, strong bases such as KOH, NaOH,
K2CO3 and ionic liquids were employed in the transesterication
of glycerol [1821,2528].

Very slow reaction rate, thus energy intensive


High cost
Enzyme can be easily deactivated methanol inhibited the enzyme
performance [59]
Additional chemicals like molecular sieves, detergent or solvent are
needed to obtain high yield [5052,54]

In transesterication of glycerol, the reaction equilibrium can


be shifted towards the product either by removing the by-product
methanol continuously or by using excess DMC in the reaction. The
former is not recommended as methanol forms a minimum boiling
azeotrop with DMC at a weight ratio of 30:70 for DMC: methanol
composition. Moreover, DMC could also be removed together with
the methanol [21,60]. Alternatively, methanol could be removed by
adding molecular sieves [50,51,54] at agitation speed below
600 rpm to avoid breaking of molecular sieves and concurrently
to create extra separation between catalyst and molecular sieves.
The other method which can shift the reaction equilibrium was
deployed by Ochoa et al. [21]. The researchers proved that when
excess DMC used at DMC/glycerol molar ratio of 5 in the reaction,
nearly 100% of glycerol was converted to yield 100% of GC in
shorter reaction time and lower catalyst loading of KOH, NaOH
and K2CO3. A pioneering work on using K2CO3 reported a 97% of
GC under mild reaction conditions [1820]. Another ndings on
high glycerol conversions using NaOH and K2CO3 [44] reported
the formation of large amounts of byproduct glycidol, which had
unfavorably affected the GC yield.
Generally, the base catalyzed transesterication gave remarkable GC yield. However, the emergence of green solvent such as
ionic liquid with unique properties had gradually gaining its versatile application as catalyst and solvent in process synthesis [61]. As
shown in Table 1, ILs such as imidazolium-based ILs [26,27],
ammonium-based ILs [25,28] and dicyanamide-based ILs [27] have
been applied in the transesterication of glycerol. Interesting to
note that 1-n-butyl-3-methylimidazolium-2-carboxylate (BMIM2-CO2) achieved 100% GC yield within 80 min [26] which has
outshone its base counterpart K2CO3 as far as reaction time is
concerned [21]. Furthermore, the chain length of ILs with the
presence of ion halide, hydroxide and bicarbonate have altered
the catalytic activity of ILs [26,28]. The glycerol conversion
had increased from 30% to 100% when ionic liquid was tested on

W.K. Teng et al. / Energy Conversion and Management 88 (2014) 484497

1-n-dodecyl-imidazoles to 1-n-butyl-imidazoles. The catalytic performance of 1-n-butylimidazole was better than higher alkyl derivatives as it promotes the 2-carboxylate formation with DMC [26]. It
has been reported that ionic liquids having hydroxide and bicarbonate counter ions are effective to convert 7795% of glycerol at
80 C, 90 min reaction time and molar ratio (DMC: glycerol) of 3
[28] as presented in Table 1.
It is economically viable for ILs to be reused. Inspite of dicyanamide-based ILs i.e. N-methyl-N-butylmorpholinium dicyanamide
[Mor1,4][N(CN)2] showed good recyclability without any signicant reduction in the conversion yield after 4 recycles [27], the
reaction conditions undermine the feasibility as higher reaction
temperature and longer duration are required moreover, it is difcult to separate the product from ILs.
Though basic homogeneous catalyst has high catalytic activity,
problem associated with the separation of the dissolved catalyst
from the product is undesirable. Unlike biodiesel reaction, water
could not be used to wash the dissolved catalyst off the product
in the reaction as GC and water are miscible. As a result, this might
incur additional separation cost. For instance, cation-exchanging
resin such as Amberlit IR 120 that was used to remove K2CO3 from
the reaction mixtures [18]. In view of that, heterogenenous catalyst
with high activity that can ease the separation process should be
considered.
2.2. Heterogeneous base catalyzed transesterication
Many solid base catalytic systems have been investigated for
the synthesis of GC via transesterication of glycerol. Catalysts
such as alkaline earth metal oxides, basic zeolites, mixed metal
oxides derived from hydrotalcites are suitable candidates. The
top most sought after catalysts are alkaline earth metal oxides like
CaO. These catalysts have relatively strong basicity and are available from cheap sources such as calcium carbonate and calcium
hydroxide [62]. GC synthesis via CaO catalyzed transesterication
had been reported by several groups of researchers [21,29
31,60]. The high catalytic activity of CaO was attributed to the
effect of its calcination at high temperature [21]. Better performance of calcined CaO than uncalcined CaO had also been proven
by Simanjuntak et al. [29]. The small deviation in the performance
of uncalcined catalyst was simply caused by the impurities such as
Ca(OH)2 and CaCO3 when water was reacted with CO2 in the atmosphere. Subsequently, the appearance of these impurities led to a
reduction in the basicity of catalyst [29].
In addition, Li and Wang [30] pointed out that high activity of
CaO might be caused by the soluble species of CaO that formed
through the interaction with glycerol and DMC. Simanjuntak
et al. [29] further identied the soluble substance as calcium complex Ca(C3H7O3)(OCO2CH3). Other metal oxides such as Na2O, MgO
and ZnO were also investigated by the same authors [29]. The catalytic activities of MgO and ZnO produced only 10.2% and 0.5%
yield of GC, whereas Na2O produced 92.6% GC with high basicity.
The low activity of MgO was expected as it has the weakest basic
strength among group II oxides [63,64]. Nevertheless, MgO had
been used as efcient catalyst support for K2CO3 and together they
demonstrated 99% yield in the catalyzed transesterication reaction [47]. Noticeably, the basicity of Mg-contained catalyst could
be improved by combining Mg with other types of metals such
as Ca, Al, Li and Zn through co-precipitation method. Many
researchers have prepared mixed magnesium-alumina (MgAl)
oxide by using hydrotalcites (HT) (Mg6Al2(OH)16CO34H2O) as precursor and calcined at high temperature [31,3743,65].
3+
Hydrotalcite or Layered Double Hydroxide (LDH), [M2+
(1 x)Mx+
n
yH2O are anionic and basic clay minerals [66].
x(OH)2] (Ax/n)
The metal cations M2+ and M3+ and anion An reside in the interlayer space of the hydroxides as shown in Fig. 3 [66,67]. The most

489

common hydrotalcite is Mg6Al2(OH)16CO34H2O and its conventional preparation method is co-precipitation [68]. Hydrotalcite is
a powerful catalyst as many of its physical and chemical properties
resemble those of clay minerals. The acid/basic properties can be
easily controlled by varying their composition making it widely
used in base-catalyzed or -assisted reactions such as alkylation,
Michael addition, ClaisenSchmidt condensation, Knoevenagel
condensation, aldol condensation, hydrogenation, olen epoxidation, alcohol oxidation and transesterication [39,43].
Eshuis and co-workers had rst patented a process of converting glycerol to a mixture of oligomers by using commercial Mg/
Al-hydrotalcite Macrosorb CT100 hydrotalcite [69]. The heterogeneous catalysts applied only managed to obtain 23% GC. The simple
preparation method for HT and its adjustable basicity encouraged
researchers to opt for catalyst preparation in house. The basic
properties of HT could be tuned via pretreatment of HT such as calcination [31,3943,65], rehydration [4143,65], changing the
anion composition in HT [65] and doping transition metal cations
on the calcined hydrotalcite [39,40]. After pretreatment at high
temperature, catalytic activity would be enhanced. As in the case
of mixed oxides (Al/Mg, Al/Li, Al/Ca) derived from hydrotalcites
by calcinations which would contain higher Lewis basic sites than
the uncalcined HT making the catalyst more effective during GC
synthesis [31].
Further increase in the catalytic activity of HT is possible via
rehydration of calcined HT with the presence of Bronsted basic
sites [41]. The rehydrated calcined HT though almost had 4 times
lower surface area, presented a 99% glycerol conversion compared
with the 76% obtained using the calcined HT despite both of them
have similar total number of basic sites. This implies that the
accessibility and the number of basic sites are not as important
as the basicity of the solid. This shows that the Brnsted basic sites
are better in extracting proton from glycerol (which presents
higher acidity compared with DEC) and thus stabilized the alkoxide anion on the surface of the solid [41].
The promotional effects on the basicity of calcined hydrotalcites
with transition metals doping was conrmed by Liu et al. [39,40].
They discovered that nickel doped hydrotalcites (HTC-Ni) exhibited 10 times higher catalytic activity than the uncalcined hydrotacites precursor in the transesterication reaction as their
reconstructed HT possess more open structure and higher basicity.
Furthermore, it is economically viable as it could be recycled and
reused at least ve times without signicant loss of performance,
giving 100% GC selectivity even after repetitive use.
The effectiveness of hydrotalcite does not limit to the calcinated
hydrotalcite in GC synthesis. The use of uncalcined Mg/Al hydrotalcite at 25 ratio in the hydromagnesite phase for the transesterication of glycerol with DMC increases the yield of GC [37,38]. The
high catalysts activity of HTHM (hydrotalciteshydromagnesite)
was attributed to the increased HT surface area and adsorption
sites for glycerol during the reaction. Besides coupling calcinations
and rehydration with methods such as ultrasound, microwave has
also been applied to alter the structure of HT [4143,65,70,71].
Regardless of its effectiveness in the catalytic transesterication
reaction, HT has limited industrial application largely due to the
difculty in obtaining desirable small particle sizes for continuous
ow reactors. However, this drawback could be overcome by
impregnating Mg/Al HT precursor salts onto the nano-scale materials supports such as a-Al2O3 and c-Al2O3 in a continuous ow
reactor [42].
A number of studies employed mixed oxides have high catalytic
activity as shown by Mg-containing bimetallic [33,34] and tri
metallic [35,36] in Table 1. Mixed metal oxides are generally preferred to the single metal oxides owing to their stronger basic
property and larger surface area. In the transesterication of
DMC with glycerol using MgLa mixed oxide [33], it was suggested

490

W.K. Teng et al. / Energy Conversion and Management 88 (2014) 484497

Fig. 3. Structure of hydrotalcite (adapted from [67]).

that the modication of catalyst preparation condition i.e. the


metal oxide molar ratio, the presence of precipitation agent and
calcination temperature alter the catalyst composition, structure,
surface area and basic site concentration and which had signicantly affected the catalytic activity. To achieve high yield of GC,
the best conditions determined were at mixed oxide molar ratio
of 3 with KOH/K2CO3 and calcined between 650 C and 720 C. In
the Mg1+xCa1 xO2 catalyzed glycerol transesterication process
[34], 100% GC yield was achieved after 90 min of reaction time at
70 C and 2:1 DMC/glycerol molar ratio. The remarkable ndings
could be explained by the synergistic interaction between Ca and
Mg species in the catalyst structure that contributing to the high
catalytic activity. Furthermore, the researchers [3334] agreed that
the nature of active sites play an inuential role in the catalytic
activity.
Evidences of mixed oxides molar ratio and pretreatment temperature that inuenced the catalysts activity are demonstrated
by the tri metallic mixed oxides Mg/Al/Zr [35] and Mg/Zr/Sr [36].
These base catalysts performed well in GC synthesis and the
well-dispersed mixed oxides in stabilized tetragonal phase of zirconia at calcination temperature of 650 C with Mg/Al/Zr mole
ratio of 3:1:1 showed higher activity of 94% yield than the bimetal
oxides (Mg/Al and Mg/Zr) which are only 15% and 25% yield,
respectively [35].
The basic site clearly has certain inuence on the transesterication yield. From the investigation of Na based zeolite and hydrotalcite in the transesterication process [44], the reaction shows
higher dependency on the catalyst structure than the basic sites.
The activity of the zeolite is not proportional to the number of
basic sites but the pore structure of zeolites. With small pore diameter such as 3A, 4A, and NaZSM-5, the catalyst was inactive
whereas those NaY and Nab with pore diameter containing a 12numbered ring-structured pore channel under mild conditions
yielded excellent GC. As for recycle possibility, NaY indicates a negligible 1% difference in the rst three cycles, and remained stable
thereafter [44].
Supplementary information on other high performance
catalysts with support is listed in Table 1. Examples shown are

KF/hydroxyapatite (HAP) [45], NaOH/c-Al2O3 [46] and K2CO3/


MgO [47] that have their acid/base properties modied by loading
with other compounds according to the transesterication reactions.
2.3. Enzymatic transesterication
Catalytic transesterications by enzymes such as lipases are
commonly found in biodiesel industry. Through the application
of enzymatic transesterication, many downstream processing
problems are simplied and eventually the production cost can
be reduced. As listed in Table 2, enzyme under mild reaction condition without the generation of by-products not only eases the
recovery of the desired product but also enables the reusability
of the enzyme. This might make enzyme replacing chemical catalyst worth to be investigated in the context of a greener GC production route. However, high cost of enzymes, their comparatively
slow reaction rates and the likelihood of enzyme deactivation have
often undermined the choice of enzyme for industrial scale production. Thus, this review only discusses the enzymatic transesterication of glycerol at lab scale.
GC synthesis via transesterication can be catalyzed with
enzymes such as Candida Antarctica lipase [5054] and Aspergillus
niger lipase [5559]. Kim et al. [50] successfully synthesized 99%
GC from glycerol and DMC using Candida Antarctica lipase B immobilized on resins Novozym 435 in the presence of solvent THF. Stoichiometric molar ratio of reactants (DMC: Glycerol) at 1:1 was
applied but the reaction took 30 h to complete and moreover the
organic solvent THF is toxic. To overcome the toxicity problem, solvent-free system was proposed [51] at 10:1 DMC: glycerol molar
ratio with the excess DMC played the dual role as reactant and solvent. To prevent two phases from forming due to the poor solubility of glycerol in hydrophobic DMC, glycerol was coated with silica
gel in equal amount though this had little effect on the reaction
rate that involving free enzyme. The problem associated with
the formation of glycerol coating could one hand be resolved
by the addition of surfactant Tween 80 to enhance mixing but
on the other hand this induces downstream processing problems. Another exploration on the use of both hydrophilic and

W.K. Teng et al. / Energy Conversion and Management 88 (2014) 484497

hydrophobic solvents by Jung et al. [52] attained 96.25% glycerol


conversion at DMC:Glycerol molar ratio of 2:1 using acetonitrile
as solvent and with the addition of detergent. The 48 h reaction
time is as lengthy as for the solvent free system [51]. Thus, tertbutanol was used to reduce the reaction time to 12 h [53,54] in
their respective reactions.
Besides Candida Antarctica lipase, Aspergillus niger lipase had
been extensively studied in GC synthesis via transesterication of
glycerol with DMC [5559]. GC was successfully produced under
mild reaction condition at shorter time but the yield of GC obtained
using Aspergillus niger lipase was only between 25% and 60%. It is
imperative therefore to develop an improved enzymatic process
scheme besides searching for suitable and effective catalysts to
synthesize GC feasibly via examining the operational factors critically. The factors inuencing the process efciency will be discussed next.
3. Factors inuencing transesterication reaction
The conversion of glycerol and the yield of GC in transesterication reaction are signicantly affected by the reaction conditions
and the parameters such as reaction temperature, reaction time,
molar ratio of reactants and solvent, as well as impurity particularly water content and methanol in the reaction mixture. Most
of the studies evaluate these factors one at a time and limited work
[21] has been done on evaluating their synergistic effects on GC
synthesis. Therefore, these pertinent factors affecting the performance of GC synthesis will be elucidated individually.
3.1. Effect of temperature
Glycerol Carbonate (GC) synthesis via transesterication of
glycerol can produce high GC yield at mild reaction temperature
and atmosphere pressure. The synthesis route has been widely
studied using different carbonates and catalysts at varying temperature range between 35 C and 140 C at atmosphere pressure.
Conventionally at temperature above 140 C, a reduced pressure
of 10 3 MPa is required in the glycerolysis reaction whereas high
pressure of 5 MPa requires the temperature range of 80180 C
in the carboxylation process [8]. The carboxylation reaction condition can achieve only 65% conversion of glycerol after 15 h [32].
The reaction temperature is obviously a critical parameter for
the transesterication of glycerol as this process is reversible and
favorable in producing GC which is thermodynamically related to
chemical equilibrium constant [72]. According to Arrhenius equation, an increase in reaction temperature can increase the collision
rate between the reactants and thus leading to a higher reaction
rate and product yield [73]. Nevertheless, the optimum reaction
temperature in transesterication is closely dependent on the heat
sensitive carbonates and catalysts used. In the transesterication
of glycerol with ethylene carbonate (EC), lower reaction temperature is favorable to achieve a better GC yield. As reported, 87% yield
of GC at 35 C was achieved within 1 h as compared to a lower
yield of GC at 50 C even with higher catalyst loading and prolonged reaction time [31]. This nding reported by Li and Wang
[72] through their calculation of chemical equilibrium constant
suggesting that chemical equilibrium constant decreased when
the temperature increased within the range of 2580 C. The reaction shifted to backward reaction at higher temperature causing a
reduction in GC formation. However, this trend does not apply to
the transesterication reaction of glycerol with DMC or diethyl carbonate (DEC). Different from the case of EC, temperature increases
within the range of 4080 C increases the chemical equilibrium
constant which in turn has increased the reaction rate of glycerol
and DMC [72]. Furthermore, the effect of temperature on the

491

conversion and yield had been witnessed by recent works which


applied metal oxides [29,3336] and ionic liquid [28,49] as
catalyst. The works involved metal oxides [29,3336] have similar
trend of temperature effect whereby less than 40% GC yield at
50 C and further increased in temperature increased the yield
up to the optimum temperature. In most of the studies, GC synthesis reaction performed well between 70 C and 90 C
[18,21,22,26,29,3336,4448] using chemical catalysts and 60 C
for biocatalysts [5059]. Higher temperature of 100120 C are
essential for the reaction system that involved hydrotalcite [37
40] and ionic liquid [2527].
To have a better understanding on the types of catalysts and
their suitable temperature range for GC synthesis, the physical
properties of the reactants should be taken into account when
determining the reaction temperature. DMC and glycerol are thermally stable and do not decompose below 390 C [74] and 150 C
[75], respectively. This implies that the suitable temperature
applied to DMC and glycerol should not be exceeded the mentioned values. This contradicted to the observation on the solid
base-catalyzed decarboxylation [76] whereby, partial decomposition of DMC to dimethyl ether and CO2 occurred at temperature
>200 C. Furthermore, it is worth noting that the dialkyl carbonates
are volatile organic compounds and can easily be evaporated. To
avoid excessive and unnecessary evaporation, the boundary of
reaction temperature should be set below their boiling points. This
was evidenced from the successful cases of transesterication of
glycerol with DMC or DEC carried out below 140 C. The temperature range required for transesterication of glycerol with DEC by
hydrotalcite was 130140 C [37,4143,65]. Even though the
transesterication of glycerol with DEC could withstand harsher
reaction conditions, there is a trade-off between the improved
yield and the heat requirement to carry out the reaction at higher
temperature.
On the other hand, enzymatic catalyzed transesterication can
be carried out at relatively low temperature range between 40 C
and 70 C [5059]. Too low a temperature i.e. 40 C, cannot adequately dissolve the enzyme and reactants [52]. In such case, an
increase in temperature is needed to reduce the viscosity of reaction mixture. The less viscous and fast moving reactant molecules
would promote the effective collisions with enzyme to increase the
reaction rate. At temperature greater than 70 C, the active conformation of enzyme might be disrupted and resulted in activity loss
and hence the conversion rate and the yield [54] are decreased.
Hence, the enzyme activity should increase with increasing temperature provided the stability of enzyme pertained.
Increase the operating temperature beyond the optimum temperature might not improve the conversion of glycerol but instead
it would probably decrease the yield. When the reaction temperature exceeds 100 C, side-reactions involved dehydrogenation and
condensation of the by-product methanol might have occurred
on the basic sites [39,40]. This was witnessed in the decarbonylation of GC to glycidol as shown in Fig. 4 [28,35]. The temperature
increment promoted the formation of glycidol and maximized
the glycerol conversion from 90 C to 110 C [35] which was manifested by the similar trend when temperature increased from
70 C to 80 C [28]. The selectivity of GC decreased as it was
consumed to form glycidol, it is therefore crucial to control the

Fig. 4. Decarbonylation of glycerol carbonate to glycidol [28,35].

492

W.K. Teng et al. / Energy Conversion and Management 88 (2014) 484497

optimum reaction temperature properly in achieving maximum


yield of GC.
Most of the GC synthesis through transesterication are in
batch modes rather than in the continuous modes. In the batch
operational mode, round bottom glass asks or jacketed glass reactors were not sufcient to withstand the high reaction temperature.
High pressure reactor was used for reactions above 75 C [29,34] and
above 90 C [21]. In the continuous system, a tubular quartz packed
bed reactor was applied at 130 C [42]. It can be seen that higher temperature tends to favor the continuous mode of operation. For
instance, continuous operation is recommended to replace batch process at reaction temperature of 175 C due to the decomposition of
product [77]. To prevent the undesired consumption of DMC or
DEC and the extensive decomposition of GC, the operating temperature limit should be established case by case for individual catalytic
system. Moreover, safety aspect should also be highlighted to eliminate the formation of highly ammable ethers as a result of the undesired decomposition of DMC at high temperature [78].
The research ndings so far suggest that chemical catalyzed
transesterication in general required higher temperature than
enzymatic transesterication. Except the case of using EC as carbonate source, more glycerol converted with the increasing temperature. For the reaction carried out beyond the optimum
temperature, the yield of GC was decreased due to the formation
of undesired products. However, the optimum temperature range
is also very much dependent on the carbonates used. Since most
of the researchers focused using different type of catalysts on
one carbonate source in the synthesis, more effort should therefore
be devoted to apply different carbonates using same catalyst
[18,53]. With that, thorough evaluation on the trend of temperature with different carbonates can be performed. Furthermore,
the reaction temperature is also related to the other reaction conditions, such as molar ratio of DMC/glycerol and reaction pressure
[21]. When transesterication of glycerol takes place under atmosphere pressure, DMC and the by-product methanol could simultaneously be recovered and owed back to the reactor, this would
signicantly affect the reaction temperature [21]. Hence, simple
distillation is normally applied in the transesterication reaction
to enable the reaction to be carried out either under total reux
condition [1821,2831,33,3944,48,65] or with removal of a substantial portion of unreacted DMC and the produced methanol
[45,46,79].
3.2. Effect of reaction time
Reaction time has effect in the synthesis of GC. Short reaction
time is predictably attractive for industrial manufacturing of GC.
In the absence of catalyst, transesterication was slow and only
5% GC yield was achieved after 5 h [31]. The shortest reaction time
required to yield 81% of GC from transesterication with EC using
catalyst CaO was around 15 min [31,49]. Generally, the reaction
rate of the base catalyzed transesterication is much faster than
that of acid catalysis due to higher catalytic activity [21]. Many different types of base catalysts as previously mentioned are suitable
for transesterication of glycerol as the reactions are rapid with no
apparent induction period [39]. Furthermore, the conversion rate
of glycerol, selectivity and yield of GC increase with reaction time
[28,29,31,3335,39,40] in the catalyzed-transesterication reaction. The yield of GC improved from 75% to 99% when the reaction
time had increased from 1 h to 2 h [37].
When immobilized ionic liquid [49] was used in the transesterication of glycerol with EC, 86% glycerol conversion in 5 min was
accompanied with the by-product of ethylene glycol of greater
than 75% selectivity. This has unfavorably lowered the GC yield.
To have higher GC yield of more than 90%, optimum reaction time
ranges from 30 min to 3 h were applied in the homogeneous cata-

lyzed transesterications [18,21,22,26,29]. For instances, CaO


achieved 94% GC yield in 30 min and K2CO3 and KOH both yielded
100% GC in 1.5 h [21]. With the employment of heterogeneous
catalyst, the reaction time were mostly between 1 and 2 h
[21,30,3340,4547,49,60,72]. One of the highly recommended
heterogeneous catalysts is Mg1+xCa1 xO2 [34] which resembles
the organometallic catalysts such as Layered Double Hydroxide
(LDH) based heterogeneous catalyst are efcient and reusable for
the transesterication reaction. The highest yield of GC synthesis
by LDH catalyst achieved so far is 99% by using uncalcined Mg/Al
hydrotalcite in 2 h [37] and 87% GC yield in 1 h using calcined
Al/Ca hydrotalcite [31]. This suggests that carbonate source used
in transesterication affects the reaction time as reected from
the slighltly longer time required for hydrotalcite in the transesterication with DMC regardless of it is well acclaimed catalytic
activity. When hydrotalcite was involved in the transesterication
of glycerol with DEC, 65% GC yield was achieved in 10 h [41]. The
low GC yield could have been caused by the long reaction time
which prolonged the exposure of reaction to high temperature
and resulted in the forming of 33% of side product, glycerol dicarbonate [77]. The formation of glycerol dicarbonate was conrmed
by another study on hydrotalcite-catalyzed transesterication
[39,40]. The side product formation decreases the GC selectivity
whereas, glycerol conversion was enhanced as the reaction time
was extended to 5 h. This can be explained by the strong adsorption of side products methanol and its derivatives on the basic sites
during the reaction [39,40].
Prolonging reaction time does not benet the conversion and
yield in the solid base- catalyzed transesterication with the
occurrence of undesirable reactions such as decomposition of cyclic carbonates [31] and the decarbonylation of GC to glycidol
[28,35]. In the transesterication of GC with carbonate sources
using strong bases catalysts, glycerol dicarbonate or diglycerol tricarbonate could be produced in 48 h [18]. A much reduced reaction
time of 8 h in the Mg/Zr/Sr mixed oxide base catalyzed-transesterication showed an increased selectivity of glycidol and a
decrease in GC selectivity [35]. Also it has been reported that for
the reaction beyond 90 min, a slight decrease in GC yield was
reported which could have been caused by the effect of solubility
of by-product methanol on the reaction mixture [34].
A few studies on supported catalyst employed in the transesterication of glycerol have been attempted to replace homogeneous
catalyst by adding support to the catalyst [4547]. When the polarity and structure of a catalyst is altered, the catalytic activity would
be inuenced [43]. For instance, by applying carbon nanober
(CNF) supported Mg/Al hydrotalcite (HT) in the transesterication
reaction, the reaction time had been reduced from 10 h to 2 h. This
is because the surface area and polarity of the HTCNF catalyst
were reduced which had enhanced the adsorption of reactants
and resulted in higher reaction rate [43].
A different yet interesting approach using glycerol-coated silica
gel in the synthesis of GC from glycerol applying lipase in DMC was
adopted to get rid of large glycerol droplet formation [51]. This
approach enhanced the accessibility of glycerol and DMC to the
enzyme and it signicantly has increased the transesterication
rate by tenfold than that of free glycerol. The reaction rate can also
be sped up by preheating the viscous glycerol before mixing with
DMC to increase the miscibility of both reactants. Similar concept
was applied by Leung and Guo [80] to increase the rate of biodiesel
reaction and shorten the reaction time through heating oil prior to
the mixing. Thus, proper mixing between reactants and catalyst
can determine the completion time of a reaction. To ensure the
operational feasibility, molar ratio of reactants must be adjusted
accordingly and also the problems of hydrophilic glycerol and
hydrophobic DMC need to be resolved by suitable choice of
solvents.

W.K. Teng et al. / Energy Conversion and Management 88 (2014) 484497

While taking short reaction time and high GC yield as selection


criteria for catalyst in the transesterication processes, reusability
and greenness of the catalytic reaction had also been considered by
many. This can be witnessed from the work done on enzymatic
catalyzed transesterication to synthesize GC as well as to overcome the drawback of lengthy reaction time (3048 h) associated
with the process. In particular, Cushing and Peretti [53] and Lanjekar et al. [54] have greatly reduced their respective reaction time
by 2-fold to 12 h and 14 h using catalyst Candida Antarctica lipase
B (Novozym 435). A more efcient enzyme Aspergillus niger lipase
that took only 46 h was employed in the biosynthesis of GC but
the end result showed a conversion of glycerol less than 75% and
a GC yield below 60% [5559].
Another uctuating trend between the reaction time and the
catalyst applied in the GC synthesis can be seen from the use of
the unconventional catalysts ionic liquids. This type of catalyst
showed a vast difference in the transeterication time range from
1.33 to 13 h [2528,49]. The inconsistent ndings probably are
caused by the interaction between the different anion and cation
species in the ionic liquid with the reaction mixture.
Higher reaction time could be advantageous as it allows more
time for the glycerol to react in both the chemical catalyzed and
enzymatic transesterication. Similar to the temperature effect,
prolonging reaction time increases the operating cost but reduces
GC yield due to undesired decomposition reaction involved. Other
than the environmental friendliness demonstrated by the enzymatic transesterication, the production route is not benecial as
it gives extremely low reaction rate and its performance is non
comparable with those of chemical catalysts. In general, production route be it chemical or enzymatically catalyzed, proper
screening for the compatibility of different carbonates and catalysts in GC synthesis is required.
3.3. Effect of molar ratio and solvent
Theoretically, the ratio for transesterication reaction requires
only 1 mol of carbonates and 1 mol of glycerol to produce 1 mol
of GC and 1 or 2 mol of relevant by-product as shown in Fig. 5.
The by-products could be ethylene glycol, propylene glycol, methanol or ethanol depending on the carbonate source used. In the
transesterication between the hydrophilic glycerol and hydrophobic carbonate source, the reactants are not miscible and the
reaction is reversible which is in need of an excess carbonate
source to give positive effect on the conversion and yield [27].
The prevention of the two-phase formation between the reactants
can be achieved either by applying excessive amount of carbonate
source over the glycerol moiety to act as reactant and solvent or by
adding organic solvent.
As reported in the literature, reactions normally carried out at
molar ratio of carbonate source to glycerol in the range of 25 to
shift the chemical equilibrium towards GC formation for greater
glycerol conversion in a shorter time as shown in Table 1. Low
GC yield was observed when equimolar of reactants was used in
the transesterication [60]. The yield of GC was increased when
the carbonate glycerol ratio is raised beyond 2 and reached a maximum molar ratio of 5 [36]. Further increase in the amount of carbonate beyond the optimal ratio has adverse effect on the yield and
would incur additional cost for the reactants. The yield would drop
when GC reacted with excess carbonate to form by-product i.e.
GDC when the DMC/glycerol ratio increased [22].
Also, molar ratio of the reactants is inuenced by the type of
catalyst used. When homogeneous catalyst was applied, molar
ratio between 2 and 5 produces 90100% GC yield [18,21,22,25
27,29] without the need of adding solvent. Furthermore, homogeneous catalyst such as ionic liquid could also serve as solvent in the
reaction [27]. In most of the heterogeneous bases catalyst reaction,

493

DMC mixed well with glycerol at low molar ratio of 2


[29,33,34,45,46]. With exceptional cases, organic solvent is added
to compensate for the low molar ratio of carbonate/glycerol
[37,38,42,44]. The reaction between the strongly polarized glycerol
and aprotic DMC or DEC allows the polar aprotic solvents such as
dimethylformamide (DMF), dimethyl sulfoxide (DMSO), acetonitrile and hydrophilic solvents such as tetrahydrofuran (THF), tertbutanol, ethanol or methanol to enhance their solubility in the
transesterication reaction. Adding suitable polar solvents DMF
[37,38] or DMSO [42] greatly improved the GC yield from 17% to
99% in the uncalcined Mg/Al hydrotalcite-catalyzed transesterication [37]. The glycerol conversion increased with the polarity of the
solvent in the descending order of DMSO, DMF and DMA respectively at 50%, 30% and 20% [42]. However, incorrect solvent choices
may lead to the formation of by-products such as glycidol and
other unknown compounds when ethanol was applied [44]. This
signies the effect of solvent polarity on the glycerol conversion
and selectivity [42,44].
In the enzymatic catalyzed transesterication reaction, high
conversion of glycerol could be achieved either by using high molar
ratio or addition of solvent. For solvent free system, high molar
ratio of 10 was suggested [51,53,5559]. Other options such as
co-solvent THF [50] or acetonitrile [52] can also be used. Regardless of the methods applied, 93% glycerol conversion was achieved
when Candida Antarctica lipase B was employed and a comparatively lower conversion of less than 75% was obtained when Aspergillus niger lipase was used. The performances between these two
biocatalysts varied largely [53] due to the un-similar reaction condition applied.
The product yield is critical to the chemical synthesis as the
selection of solvent is to the design of experiment [42]. To be in
tandem with green chemistry, non-toxic solvent must be considered for the enzymatic catalyzed reaction. Therefore, it is not surprise that green solvent i.e. tert-butanol had been selected for
catalytic transesterication of glycerol and promising conversion
of 94.85% [54] and nearly 100% [53] were reported. These ndings
have placed tert-butanol, a non-toxic and chemically inert solvent,
favorable for enzymatic reaction to promote singe phasic system
[53,54]. Other advantages of tert-butanol are its ability to increase
enzyme exibility, as it allows the enzyme to be more easily bound
with substrates and alter their conformation to facilitate reaction
for improving the catalytic activity [53].
Solvent could also play an essential role as azeotropic agent on
the synthesis of GC. In such case, azeotropic agent was required for
the rupture of methanol-DMC azeotropic mixture formed due to
their narrow boiling points [81]. Addition of a suitable solvent to
the reaction mixture can alter the relative volatility of the original
binary mixture to eliminate the azeotrope, thus facilitate the
separation of methanol from DMC [82]. In azeotropic distillation
coupled with glycerol transesterication reaction, the suitable azeotropic agent or entrainer are cyclohexane, n-hexane, n-heptane,
isooctane, ethyl acetate, cyclohexene, benzene, dichloroethane as
they could form the azeotrope with methanol [60]. Benzene was
found as the most effective azeotropic agent which gave a 98%
GC yield even at the stoichiometric molar ratio of DMC/glycerol
[60]. All of the DMC was retained in the reactor for reaction
whereas the resulting mixture of methanol and benzene was
removed as a distillate in the distillation column [60]. The high
yield of GC was attributed to the positively shift of the reversible
transesterication reaction by continuously removal of the produced methanol [60]. In short, azeotropic distillation is capable
to overcome the limitation caused by the thermodynamic phase
equilibrium with high yield of GC achieved without excess DMC
used. The drawback of this process are energy intensive and
environmental unfriendly due to the carcinogenic nature of the
entrainer used [83].

494

W.K. Teng et al. / Energy Conversion and Management 88 (2014) 484497

Fig. 5. Transesterication of glycerol with different carbonate sources [8].

From economic and environment viewpoint, low molar ratio of


carbonate/glycerol and solvent free are preferable in the transesterication reaction. Nonetheless, high molar ratio or addition of
solvent is essential to increase the miscibility of glycerol with
DMC or DEC. If high molar ratio of DMC/glycerol is adopted, excess
DMC could be easily separated from the reaction product together
with methanol via distillation. If solvent or azeotropic agent were
to be applied to the system, more efcient separation and purication methods are needed. In brief, carbonates should present in
excess of stoichiometric proportion for the effective transesterication of GC.
3.4. Effect of impurities
Impurities are undesirable for reaction generally. They can be
appeared in both reactants carbonate source and crude glycerol
in the transesterication reaction. Crude glycerol obtained from
biodiesel plant contain many impurities such as methanol, water
and soaps, fatty acids, salts like phosphates and sulfates, and
metals such as Na, K, Ca, Mg and Mn depending on the type of oils
and catalysts used in the production [15]. These impurities may be
low in concentration but is sufcient to affect the conversion of
glycerol and the GC yield.
Pure (99%) glycerol was used in most of the transesterication
of glycerol with very few using crude glycerol at content varying
from 40% to 90% [26,58,59,84]. Crude glycerol can be obtained
either before neutralization/acid treatment (4070%) or after acid
treatment (above 80%). A 88 wt% (59 mol%) crude glycerol gave a
GC yield of 93% and 100% yield of GC when pure glycerol was used
[26]. A slightly lower yield of GC from the crude glycerol could be
caused by the water and salt content present in the crude glycerol.
The crude glycerol used by Ilham et al. [84] for GC synthesis with

supercritical DMC have the same impurities as reported by Naik


et al. [26], whereby the glycerol obtained from alkaline-catalyzed
biodiesel contained 70 wt% glycerol, 10 wt% of water and 20 wt%
of sodium salt. When these impurities reacted with supercritical
DMC, the GC produced further decomposed to form glycidol as
shown in Fig. 4. This further demonstrated that water and salt
too could lower the GC yield under supercritical condition.
The presence of water and methanol in crude glycerol also
affected the GC yield signicantly in biosynthesis of GC with lipase
[59]. It was reported that a 17 wt% of water in crude glycerol could
lead to a decrease of 11% in the GC yield. The negative effect of
water on GC yield is more obvious for higher water content in
crude glycerol. The reduction of yield was caused by the inhibitory
effect of lipase activity by water as the enzyme required preponderantly a hydrophobic environment. The presence of water
jeopardized the stability of GC as well as the short-chain alcohols
like methanol and subsequently destabilized the protein [59]. Conversely, trace amount of water (hydrophilic phase) can conserve
the spatial enzyme structure. Besides, impurities like soap have
no effect on the biocatalytic reaction and do not inuence the
transesterication reaction [59].
Crude glycerol produced from different oils contains different
impurities and thus can result in different GC yield under identical
reaction conditions. Glycerol content varies from 40% to 75% in
crude glycerol derived from different oils. The nature of the fatty
ester precursor determines the yield, which is further dependent
on the percentage of methanol and water present in the crude
glycerol produced. Apart from that, lipase activity could be affected
by metal ions like sodium, potassium and magnesium but they
normally were not detected in most of the crude glycerol. Among
the metals and salts, only manganese which was found in crude
glycerol derived corn oil can affect the GC yield [59].

W.K. Teng et al. / Energy Conversion and Management 88 (2014) 484497

As a main contaminant, water presents in crude glycerol and


commercial pure glycerol [18] as well as in the reagent grade
DMC containing impurities such as water and methanol [53].
Water showed negative effect to both the chemical catalyzed synthesis [18,28,48] and enzymatic reactions [52,53]. When ionic
liquid was used as catalyst, the glycerol conversion decreased
signicantly from 95% to 35% with the increase amount of water
in the reactants from 0.059 g to 0.238 g [28]. Extruded CaO/Al2O3
has certain water-resistant performance, 92.1% glycerol conversion and 97.7% GC yield were obtained within 0.4 wt% of water
in glycerol. As the water content in glycerol increased continually
to 2 wt%, the glycerol conversion and GC yield drop to half [48]
as the catalyst was probably deactivated with large quantity of
water in glycerol. For enzymatic reaction, small amount of water
improves the enzymatic activity by providing more interfacial area
[52] but beyond 0.25% (v/v) of water an inhibitory effect on the
enzyme activity would be induced. In addition, excess water
shifted the reaction equilibrium in favor of hydrolysis and thus
limiting the overall glycerol conversion [53]. Thus, it is essential
to reduce the water content to an acceptable level in the reactants
prior to be used in transeterication reaction. Water could be
removed to less than 2% of water content in glycerol via pretreating the commercial pure glycerol by azeotropic [18] or via mixing
the reagent grade DMC or crude glycerol with molecular sieves
overnight at room temperature [53].
The known positive aspect of impurities as far as transesterication of crude glycerol is concerned is associated with the promotion of the recycle capacity of the biocatalyst. The GC yield was
enhanced by 2050% after the rst 23 cycles, and the catalytic
capacity of the biocatalyst was preserved within the next 10 reaction cycles [58]. Also, biocatalyst can perform better with the
removal of the supercial enzyme molecules kept by protein
protein interactions on the biocatalyst surface making the catalytic
sites accessible. The soap impurities which act as surfactants can
serve the purpose to breakdown the interactions [58].
Weighing the pros and cons of crude glycerol, the energy intensive distillation that currently applied in the industry for removing
the impurities from crude glycerol to obtain pure glycerol is undesirable. One promising report suggests that KF/HAP [45] is stable in
the presence of water and soap [45] which signies the possibility
of producing high GC yield via transesterication of crude glycerol.
Thus, further exploration to ascertain the feasibility of transesterication of crude glycerol in GC production is encouraged.

4. Future direction of glycerol carbonate production via


catalytic transesterication
Catalytic transesterication of glycerol is a simple and efcient
route to produce GC. Most of the studies focused on developing
new catalysts to produce high yield of GC and very few studies
are on purication of GC. To have feasible GC production at industrial scale, process optimization of the production route and
improved product isolation techniques are essential. Furthermore,
the study of the inuencing factors on the performance of GC production that conned in a single factor designed in most of the
works reported has raised the need to conduct design of experiment more thoroughly to examine the synergistic interactions of
the parameters investigated. Also, it is prudent to point out that
though chemical catalyzed transesterication of glycerol is a promising route to produce GC, the plausible yields had been derived
mainly from the use of pure glycerol. Due to the fact that pure glycerol obtained from crude glycerol that contains impurities incurs
high purication cost, future studies should gear towards using
crude glycerol directly obtained from the biodiesel plant to produce GC. This would greatly reduce the production cost of GC

495

and broaden its industrial usages. Along with that, this shall open
up an avenue for making biodiesel production a truly economic
viable and integrated process.
5. Conclusions
High yield of GC can be achieved by transesterication of glycerol using different type of catalysts including pure and mixed
metal oxides, ionic liquids, hydrotalcites and lipase. It is worth noting that the effects of the operating parameters on the GC production are closely dependent on the type of catalyst applied. In
chemical catalyzed transesterication, homogeneous based catalyst gave good yield in relatively short time meanwhile the use
of heterogeneous based catalysts have also become popular due
to their high activities, simple recovery methods and their abilities
to be recycled. Enzymatic transesterication with milder operating
condition could circumvent the disadvantages imposed by the
chemical catalyzed reaction. The proper control of the operating
parameters in transesterication is important to ensure the success in GC production as well as to minimize the formation of
undesired intermediates and side products which would complicate the subsequent downstream processing. Organic solvent,
molecular sieves, surfactant and silica gel could be used in GC
enhancement provided the cost for product purication involved
is justiable.
Acknowledgement
The authors thank University of Malaya for supporting this
research under the Grants of HIR (High Impact Research) with Project no. UM.C/625/1/HIR/MOHE/ENG/59 and University of Malaya
Research Grant (UMRG)RP002B-13AET.
References
[1] Abbaszaadeh A, Ghobadian B, Omidkhah MR, Naja G. Current biodiesel
production technologies: a comparative review. Energy Convers Manage
2012;63:13848.
[2] Demirbas A. Progress and recent trends in biodiesel fuels. Energy Convers
Manage 2009;50:1434.
[3] Yang F, Hanna MA, Sun R. Value-added uses for crude glycerol a byproduct of
biodiesel production. Biotechnol Biofuels 2012;5:110.
[4] Quispe CAG, Coronado CJR, Carvalho Jr JA. Glycerol: production, consumption,
prices, characterization and new trends in combustion. Renew Sust Energy Rev
2013;27:47593.
[5] Ciriminna R, Pina CD, Rossi M, Pagliaro M. Understanding the glycerol market.
Eur J Lipid Sci Technol 2014. n/an/a.
[6] Pagliaro M, Rossi M. The future of glycerol. Cambridge: RSC Publishing; 2010.
[7] Schols E. Production of cyclic carbonates from CO2 using renewable feedstocks.
Lille, France; 2014. p. CEOPS Workshop R&D on CO2 utilization.
[8] Sonnati MO, Amigoni S, Tafn de Givenchy EP, Darmanin T, Choulet O, Guittard
F. Glycerol carbonate as a versatile building block for tomorrow: synthesis,
reactivity, properties and applications. Green Chem 2013;15:283306.
[9] Posey ML, Zhao H, Zhang V. Glycerin carbonate a unique and versatile
chemical. In: The 12th annual green chemistry and engineering conference;
2008.
[10] GlaconChemie GmbH: isopropylidene glycerine, glycerine formal, glycerine
carbonate. <http://www.glaconchemie.de/cms/upload/yer/glaconews_engl_02_
screen.pdf> [accessed 15.07.14].
[11] Ang GT, Tan KT, Lee KT. Recent development and economic analysis of
glycerol-free processes via supercritical uid transesterication for biodiesel
production. Renew Sust Energy Rev 2014;31:6170.
[12] Ochoa-Gmez JR, Gmez-Jimnez-Aberasturi O, Ramrez-Lpez C, Belsu M. A
brief review on industrial alternatives for the manufacturing of glycerol
carbonate, a green chemical. Org Process Res Dev 2012;16:38999.
[13] Huntsman corporation: JEFFSOL glycerine carbonate. <http://www. huntsman.
com/portal/page/portal/performance_products/Media%20Library/global/les/jeffsol_
glycerine_carbonate.pdf> [accessed 15.07.14].
[14] Huntsman corporation: glycerine carbonate in beauty and personal
care. <http://www.huntsman.com/performance_products/a/Home> [accessed
15.07.14].
[15] Tan HW, Abdul Aziz AR, Aroua MK. Glycerol production and its applications as
a raw material: a review. Renew Sust Energy Rev 2013;27:11827.
[16] Nemirowsky J. Ueber die Einwirkung von Chlorkohlenoxyd auf
Glycolchlorhydrin. J Prakt Chem 1885;31:1735.

496

W.K. Teng et al. / Energy Conversion and Management 88 (2014) 484497

[17] Huntsman
corporation:
JEFFSOL
alkylene
carbonate
<http://
www.huntsman.com/performance_products/Media%20Library/global/les/
jeffsol_alkylene_carbonates_brochure.pdf> [accessed 6.06.14].
[18] Rokicki G, Rakoczy P, Parzuchowski P, Sobiecki M. Hyperbranched aliphatic
polyethers obtained from environmentally benign monomer: glycerol
carbonate. Green Chem 2005;7:52939.
[19] Herseczki Z, Tams V, Gyula M. Synthesis of glycerol carbonate from glycerol, a
by-product of biodiesel production. Int J Chem Reactor Eng 2009:7.
[20] Herseczki Z, Tams V, Gyula M. Enhanced used of renewable resources:
transesterication of glycerol. Hung J Ind Chem 2011:1837.
[21] Ochoa-Gmez JR, Gmez-Jimnez-Aberasturi O, Maestro-Madurga B,
Pesquera-Rodrguez A, Ramrez-Lpez C, Lorenzo-Ibarreta L, et al. Synthesis
of glycerol carbonate from glycerol and dimethyl carbonate by
transesterication: catalyst screening and reaction optimization. Appl Catal
A Gen 2009;366:31524.
[22] Ochoa-Gomez JR, Gomez-Jimenez-Aberasturi O, Ramirez-Lopez C, MaestroMadurga B. Synthesis of glycerol 1,2-carbonate by transesterication of
glycerol with dimethyl carbonate using triethylamine as a facile separable
homogeneous catalyst. Green Chem 2012;14:336876.
[23] Hervert B, McCarthy PD, Palencia H. Room temperature synthesis of glycerol
carbonate catalyzed by N-heterocyclic carbenes. Tetrahedron Lett
2014;55:1336.
[24] Patel Y, George J, Pillai SM, Munshi P. Effect of liophilicity of catalyst in cyclic
carbonate formation by transesterication of polyhydric alcohols. Green Chem
2009;11:105660.
[25] Grey RA. Preparation of cyclic carbonates using alkylammonium and tertiary
amine catalysts 1992.
[26] Naik PU, Petitjean L, Refes K, Picquet M, Plasseraud L. Imidazolium-2carboxylate as an efcient, expeditious and eco-friendly organocatalyst for
glycerol carbonate synthesis. Adv Synth Catal 2009;351:17536.
[27] Chiappe C, Rajamani S. Synthesis of glycerol carbonate from glycerol and
dimethyl carbonate in basic ionic liquids. Pure Appl Chem 2012;84:75562.
[28] Gade SM, Munshi MK, Chherawalla BM, Rane VH, Kelkar AA. Synthesis of
glycidol from glycerol and dimethyl carbonate using ionic liquid as a catalyst.
Catal Commun 2012;27:1848.
[29] Simanjuntak FSH, Kim TK, Lee SD, Ahn BS, Kim HS, Lee H. CaO-catalyzed
synthesis of glycerol carbonate from glycerol and dimethyl carbonate:
Isolation and characterization of an active Ca species. Appl Catal A Gen
2011;401:2205.
[30] Li J, Wang T. On the deactivation of alkali solid catalysts for the synthesis of
glycerol carbonate from glycerol and dimethyl carbonate. Reac Kinet Mech Cat
2011;102:11326.
[31] Climent MJ, Corma A, De Frutos P, Iborra S, Noy M, Velty A, et al. Chemicals
from biomass: synthesis of glycerol carbonate by transesterication and
carbonylation with urea with hydrotalcite catalysts. The role of acidbase
pairs. J Catal 2010;269:1409.
[32] Aresta M, Dibenedetto A, Nocito F, Pastore C. A study on the carboxylation of
glycerol to glycerol carbonate with carbon dioxide: the role of the catalyst,
solvent and reaction conditions. J Mol Catal A Chem 2006;257:14953.
[33] Simanjuntak FSH, Widyaya VT, Kim CS, Ahn BS, Kim YJ, Lee H. Synthesis of
glycerol carbonate from glycerol and dimethyl carbonate using magnesium
lanthanum mixed oxide catalyst. Chem Eng Sci 2013;94:26570.
[34] Khayoon MS, Hameed BH. Mg1 + xCa1 xO2 as reusable and efcient
heterogeneous catalyst for the synthesis of glycerol carbonate via the
transesterication of glycerol with dimethyl carbonate. Appl Catal A Gen
2013.
[35] Parameswaram G, Srinivas M, Hari Babu B, Sai Prasad PS, Lingaiah N.
Transesterication of glycerol with dimethyl carbonate for the synthesis of
glycerol carbonate over Mg/Zr/Sr mixed oxide base catalysts. Catal Sci Technol
2013.
[36] Malyaadri M, Jagadeeswaraiah K, Sai Prasad PS, Lingaiah N. Synthesis of
glycerol carbonate by transesterication of glycerol with dimethyl carbonate
over Mg/Al/Zr catalysts. Appl Catal A Gen 2011;401:1537.
[37] Takagaki A, Iwatani K, Nishimura S, Ebitani K. Synthesis of glycerol carbonate
from glycerol and dialkyl carbonates using hydrotalcite as a reusable
heterogeneous base catalyst. Green Chem 2010;12:57881.
[38] Kumar A, Iwatani K, Nishimura S, Takagaki A, Ebitani K. Promotion effect of
coexistent hydromagnesite in a highly active solid base hydrotalcite catalyst
for transesterications of glycols into cyclic carbonates. Catal Today
2012;185:2416.
[39] Liu P, Derchi M, Hensen EJM. Synthesis of glycerol carbonate by
transesterication of glycerol with dimethyl carbonate over MgAl mixed
oxide catalysts. Appl Catal A Gen 2013.
[40] Liu P, Derchi M, Hensen EJM. Promotional effect of transition metal doping on
the basicity and activity of calcined hydrotalcite catalysts for glycerol
carbonate synthesis. Appl Catal B Environ 2014;144:13543.
[41] Alvarez MG, Segarra AM, Contreras S, Sueiras JE, Medina F, Figueras F.
Enhanced use of renewable resources: transesterication of glycerol catalyzed
by hydrotalcite-like compounds. Chem Eng J 2010;161:3405.
[42] lvarez MG, Plkov M, Segarra AM, Medina F, Figueras F. Synthesis of
glycerol carbonates by transesterication of glycerol in a continuous system
using supported hydrotalcites as catalysts. Appl Catal B Environ 2012;113
114:21220.
[43] lvarez MG, Frey AM, Bitter JH, Segarra AM, de Jong KP, Medina F. On the role
of the activation procedure of supported hydrotalcites for base catalyzed

[44]

[45]

[46]

[47]

[48]

[49]

[50]

[51]

[52]

[53]
[54]
[55]

[56]

[57]

[58]

[59]

[60]

[61]
[62]
[63]

[64]

[65]

[66]

[67]
[68]

[69]
[70]

[71]
[72]
[73]

[74]

reactions: glycerol to glycerol carbonate and self-condensation of acetone.


Appl Catal B Environ 2013;134135:2317.
Pan S, Zheng L, Nie R, Xia S, Chen P, Hou Z. Transesterication of glycerol with
dimethyl carbonate to glycerol carbonate over Nabased zeolites. Chinese J
Catal 2012;33:17727.
Bai R, Wang S, Mei F, Li T, Li G. Synthesis of glycerol carbonate from glycerol
and dimethyl carbonate catalyzed by KF modied hydroxyapatite. J Ind Eng
Chem 2011;17:77781.
Bai R, Wang Y, Wang S, Mei F, Li T, Li G. Synthesis of glycerol carbonate from
glycerol and dimethyl carbonate catalyzed by NaOH/c-Al2O3. Fuel Process
Technol 2013;106:20914.
Du M, Li Q, Dong W, Geng T, Jiang Y. Synthesis of glycerol carbonate from
glycerol and dimethyl carbonate catalyzed by K2CO3/MgO. Res Chem Intermed
2012;38:106977.
Lu P, Wang H, Hu K. Synthesis of glycerol carbonate from glycerol
and dimethyl carbonate over the extruded CaO-based catalyst. Chem Eng J
2013.
Cho H-J, Kwon H-M, Tharun J, Park D-W. Synthesis of glycerol carbonate from
ethylene carbonate and glycerol using immobilized ionic liquid catalysts. J Ind
Eng Chem 2010;16:67983.
Kim SC, Kim YH, Lee H, Yoon DY, Song BK. Lipase-catalyzed synthesis of
glycerol carbonate from renewable glycerol and dimethyl carbonate through
transesterication. J Mol Catal B Enzym 2007;49:758.
Lee KH, Park CH, Lee EY. Biosynthesis of glycerol carbonate from glycerol by
lipase in dimethyl carbonate as the solvent. Bioprocess Biosyst Eng
2010;33:105965.
Jung H, Lee Y, Kim D, Han SO, Kim SW, Lee J, et al. Enzymatic production of
glycerol carbonate from by-product after biodiesel manufacturing process.
Enzyme Microb Technol 2012;51:1437.
Cushing KA, Peretti SW. Enzymatic processing of renewable glycerol into
value-added glycerol carbonate. RSC Adv 2013;3:18596604.
Lanjekar K, Rathod VK. Utilization of glycerol for the production of glycerol
carbonate through greener route. J Environ Chem Eng 2013;1:12316.
Tudorache M, Protesescu L, Negoi A, Parvulescu VI. Recyclable biocatalytic
composites of lipase-linked magnetic macro-/nano-particles for glycerol
carbonate synthesis. Appl Catal A Gen 2012;437438:905.
Tudorache M, Protesescu L, Coman S, Parvulescu VI. Efcient bio-conversion of
glycerol to glycerol carbonate catalyzed by lipase extracted from Aspergillus
niger. Green Chem 2012;14:47882.
Tudorache M, Nae A, Coman S, Parvulescu VI. Strategy of cross-linked enzyme
aggregates onto magnetic particles adapted to the green design of biocatalytic
synthesis of glycerol carbonate. RSC Adv 2013;3:40528.
Tudorache M, Negoi A, Protesescu L, Parvulescu VI. Biocatalytic alternative for
bio-glycerol conversion with alkyl carbonates via a lipase-linked magnetic
nano-particles assisted process. Appl Catal B Environ 2014.
Tudorache M, Negoi A, Tudora B, Parvulescu VI. Environmental-friendly
strategy for biocatalytic conversion of waste glycerol to glycerol carbonate.
Appl Catal B Environ 2014.
Li J, Wang T. Coupling reaction and azeotropic distillation for the synthesis of
glycerol carbonate from glycerol and dimethyl carbonate. Chem Eng Process
Process Intensif 2010;49:5305.
Mohammad Fauzi AH, Amin NAS. An overview of ionic liquids as solvents in
biodiesel synthesis. Renew Sust Energy Rev 2012;16:577086.
Zabeti M, Wan Daud WMA, Aroua MK. Activity of solid catalysts for biodiesel
production: a review. Fuel Process Technol 2009;90:7707.
Lam MK, Lee KT, Mohamed AR. Homogeneous, heterogeneous and enzymatic
catalysis for transesterication of high free fatty acid oil (waste cooking oil) to
biodiesel: a review. Biotechnol Adv 2010;28:50018.
Kouzu M, Kasuno T, Tajika M, Sugimoto Y, Yamanaka S, Hidaka J. Calcium oxide
as a solid base catalyst for transesterication of soybean oil and its application
to biodiesel production. Fuel 2008;87:2798806.
lvarez MG, Chimento RJ, Figueras F, Medina F. Tunable basic and textural
properties of hydrotalcite derived materials for transesterication of glycerol.
Appl Clay Sci 2012;58:1624.
Helwani Z, Aziz N, Bakar MZA, Mukhtar H, Kim J, Othman MR. Conversion of
Jatropha curcas oil into biodiesel using re-crystallized hydrotalcite. Energy
Convers Manage 2013;73:12834.
Yong Z, Rodrigues ArE. Hydrotalcite-like compounds as adsorbents for carbon
dioxide. Energy Convers Manage 2002;43:186576.
Sharma YC, Singh B, Korstad J. Latest developments on application of
heterogenous basic catalysts for an efcient and eco friendly synthesis of
biodiesel: a review. Fuel 2011;90:130924.
Eshuis JJW. Polyglycerol production 1995.
Benito P, Guinea I, Labajos FM, Rocha J, Rives V. Microwave-hydrothermally
aged Zn, Al hydrotalcite-like compounds: inuence of the composition and the
irradiation conditions. Micropor Mesopor Mat 2008;110:292302.
Rivera JA, Fetter G, Bosch P. Microwave power effect on hydrotalcite synthesis.
Micropor Mesopor Mat 2006;89:30614.
Li J, Wang T. Chemical equilibrium of glycerol carbonate synthesis from
glycerol. J Chem Thermodyn 2011;43:7316.
Yadav GD, Kadam AA. Selective engineering using MgAl calcined hydrotalcite
and microwave irradiation in mono-transesterication of diethyl malonate
with cyclohexanol. Chem Eng J 2013;230:54757.
Cross J, Hunter R, Stimson V. The thermal decomposition of simple carbonate
esters. Aust J Chem 1976;29:147781.

W.K. Teng et al. / Energy Conversion and Management 88 (2014) 484497


[75] Benavides PT, Salazar J, Diwekar U. Economic comparison of continuous and
batch production of biodiesel using soybean oil. Environ Prog Sustainable
Energy 2013;32:1124.
[76] Selva M, Benedet V, Fabris M. Selective catalytic etherication of glycerol
formal and solketal with dialkyl carbonates and K2CO3. Green Chem
2012;14:188200.
[77] Bell JB, Arthur CV. Method for preparing glycerin carbonate 1959.
[78] Selva M, Fabris M, Perosa A. Decarboxylation of dialkyl carbonates to
dialkyl ethers over alkali metal-exchanged faujasites. Green Chem 2011;13:
86372.
[79] Zhao H, Posey M. Reactive recovery of dimethyl carbonate from dimethyl
carbonate/methanol mixtures 2013.
[80] Leung D, Guo Y. Transesterication of neat and used frying oil: optimization
for biodiesel production. Fuel Process Technol 2006;87:88390.

497

[81] Holtbruegge J, Wierschem M, Steinruecken S, Voss D, Parhomenko L, Lutze P.


Experimental investigation, modeling and scale-up of hydrophilic vapor
permeation membranes: separation of azeotropic dimethyl carbonate/
methanol mixtures. Sep Purif Technol 2013;118:86278.
[82] Brignole E, Pereda S. Separation of azeotropic mixtures. In: Esteban B, Selva P,
editors. Supercritical uid science and technology. Elsevier; 2013. p. 179213
[chapter 8].
[83] Kunnakorn D, Rirksomboon T, Siemanond K, Aungkavattana P, Kuanchertchoo
N, Chuntanalerg P, et al. Techno-economic comparison of energy usage
between azeotropic distillation and hybrid system for waterethanol
separation. Renew Energy 2013;51:3106.
[84] Ilham Z, Saka S. Conversion of glycerol as by-product from biodiesel
production to value-added glycerol carbonate. In: Zero-carbon energy Kyoto
2011. Springer; 2012. p. 12733.

Anda mungkin juga menyukai