You are on page 1of 13

Journal of Cleaner Production 295 (2021) 126437

Contents lists available at ScienceDirect

Journal of Cleaner Production


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

Valorization of glycerol into glycerol carbonate using the stable


heterogeneous catalyst of Li/MCM-41
Shivali Arora , Vijayalakshmi Gosu , U.K. Arun Kumar , Verraboina Subbaramaiah *
Department of Chemical Engineering, Malaviya National Institute of Technology Jaipur, 302017, India

a r t i c l e i n f o a b s t r a c t

Article history: The present study explored the catalytic activity of the heterogeneous catalyst for transesterification of
Received 19 September 2019 glycerol into glycerol carbonate, a versatile compound. Transesterification of glycerol was investigated
Received in revised form with different active metals (Li, La, Ce, Mg, K) impregnated on MCM-41 (Mobil Composition of Matter No.
12 February 2021
41) framework. Among these, lithium incorporated MCM-41 proved the better catalytic activity towards
Accepted 17 February 2021
the formation of glycerol carbonate. The shortening of a long-range hexagonal array was observed with
Available online 21 February 2021
active metal incorporation due to the accumulation of non-framework Li species in the MCM-41
Handling editor: Dr Sandra Caeiro structure. BET study revealed that Li/MCM-41 possess type IV hysteresis loop according to IUPAC stan-
dards. The average pore diameter was increased from 25.43 Å to 62.02 Å with active metal incorporation
Keywords: in the MCM-41 framework. The catalytic activity of Li/MCM-41 was observed by varying different weight
MCM-41 ratios of active metal and the calcination temperature. The results demonstrated that 5 wt% Li
Glycerol carbonate impregnated on MCM-41, calcined at 450  C, appeared to have a maximum yield of glycerol carbonate.
Glycerol Additionally, the influence of reaction operating parameters was also investigated. The results showed
Transesterification
that 99 ± 1.89% glycerol conversion and 93.14 ± 2.52% glycerol carbonate yield was achieved at dimethyl
Kinetics
carbonate-to-glycerol molar ratio of 3, catalyst dosage of 4 wt% (relative to glycerol mass) and a reaction
temperature of 90  C in 3 h. The recyclability and stability of the screened catalyst was also studied under
optimized conditions. The activation energy of the catalyst was determined by solving the differential
equation using MATLAB ODE15s tool, and the obtained value was 53.77 ± 3.26 kJ/mol. The E-factor and
PMI (Process mass intensity) values of glycerol carbonate synthesis were determined as 1.16 and 2.16,
respectively, which demonstrate the less waste generation during the process.
© 2021 Elsevier Ltd. All rights reserved.

1. Introduction Amongst, transesterification of GLY into glycerol carbonate (GLC)


had received significant attention as a promising chemical due to
The consumption of fossil fuels has increased drastically with its vast industrial applicability. As this GLC possess potential
rise in concern about greenhouse gas emissions, and the stringent properties such as non-toxicity, high density, excellent biodegrad-
laws imposed by government demands bio-based fuel alternatives. ability, low freezing point, and high chemical reactivity (Song et al.,
Among various renewable resources, biodiesel has gained accept- 2017; Rokicki et al., 2005). GLC can be used as a monomer in the
ability as an alternative fuel to the global petroleum calamity. polymer and plastic industry, as a electrolyte in the semiconductor
Transesterification of vegetable oil/fat with alcohol generates bio- industry, a curing agent in the cement industry, and plant activator
diesel. During the biodiesel producing process, approximately 10% in the agricultural sector (Teng et al., 2014; Simanjuntak et al.,
of glycerol (GLY) is produced as a by-product, which needs to be 2011).
utilized effectively to make the biodiesel industry sustainable The transesterification of GLY with dimethyl carbonate (DMC)
(Arora et al., 2020). has proved to be an environmentally benign route for GLC syn-
Therefore, faster commercialization of processes with low cap- thesis. Since the process occurs under mild reaction conditions
ital cost and the short energy-intensive route is highly desirable. using non-toxic reactants (DMC and GLY). To improve process
selectivity and yield, a highly efficient alkaline active site catalyst is
required. Several homogeneous catalysts were explored to facilitate
* Corresponding author. the transesterification reaction, such as, KOH, K2CO3, and NaOH
E-mail address: vsr.chem@mnit.ac.in (V. Subbaramaiah). (Rokicki et al., 2005; Herseczki et al., 2011; Ochoa-Go mez et al.,

https://doi.org/10.1016/j.jclepro.2021.126437
0959-6526/© 2021 Elsevier Ltd. All rights reserved.
S. Arora, V. Gosu, U.K.A. Kumar et al. Journal of Cleaner Production 295 (2021) 126437

enough basic sites, so strengthening the basic site incorporation of


Nomenclature active metal is highly required. Apart from all alkali/alkaline earth
metals, lithium provides strong alkaline sites because of its strong
CGLY Concentration of glycerol at time t, mol/mL ion size effect (Corma et al., 2005). Besides, various authors
CDMC Concentration of dimethyl carbonate at time t, investigated Liþ as an active metal on different supports to boost
mol/mL the basic sites of catalyst for the synthesis of biodiesel through
CGLC Concentration of glycerol carbonate at time t, mol/ transesterification (Xie et al., 2007; Kaur and Ali, 2011; Wang et al.,
mL 2012; Boro et al., 2014; Puna et al., 2014). The aforementioned
CM Concentration of methanol at time t, mol/mL phenomenon necessitate the author’s interest to use Li as an active
Ct Concentration of total active sites present on the metal on MCM-41 for transesterification of GLY into GLC. To the
catalyst best of the author’s knowledge, the investigation of Li/MCM-41
Ea Activation energy, kJ/mol catalyst as a heterogeneous catalyst was not performed in the
k1 Forward reaction rate constant for the adsorption transesterification of GLY with DMC to GLC conversion.
of glycerol on active site of catalyst, min1 In the present manuscript, Li species were incorporated on
k-1 Backward reaction rate constant for the MCM-41 by wet-impregnation technique. Detailed characterization
adsorption of glycerol on active site of catalyst, has been used to investigate the prepared catalyst’s physico-
min1 chemical properties using FT-IR, SEM, XRD, N2 adsorption-
k2 Forward reaction rate constant for adsorption of desorption isotherm, and Hammett titration method. Trans-
dimethyl carbonate on active site of catalyst, esterification reaction was carried out, and the effect of reaction
min1 influencing conditions as well as the catalyst recyclability was
k-2 Backward reaction rate constant for adsorption of optimized. Apart from the optimization of reaction conditions, re-
dimethyl carbonate on active site of catalyst, action kinetics for the transesterification reaction was also exam-
min1 ined. The Langmuir-Hinshelwood-Hougen-Watson (LHHW) model
k3 Forward reaction rate constant for was used to develop the kinetics of the reaction since it was
transesterification of glycerol to glycerol identified as the most trust-worthy model to study heterogeneous
carbonate, min1 catalytic reactions.
ke3 Backward reaction rate constant for
transesterification of glycerol to glycerol 2. Material and method
carbonate, min1
k4 Forward reaction rate constant for desorption of 2.1. Materials
methanol, min1
k-4 Backward reaction rate constant for desorption of All chemicals used in this study were of analytical grade, and
methanol, min1 used without further purification. Glycerol (99% purity), methanol
K1 Equilibrium constant for adsorption of glycerol on (99% purity), phenolphthalein, ammonium hydroxide, 2,4 nitro-
active site of catalyst aniline, and cetyl trimethyl ammonium bromide (CTAB, surface
K2 Equilibrium constant for adsorption of dimethyl directing agent) were purchased from Fisher Scientific, Mumbai,
carbonate on active site of catalyst India. Dimethyl carbonate (98% purity) and lithium nitrate were
K3 Equilibrium constant for desorption of glycerol acquired from Loba Chemie, Mumbai, India. Tetraethylorthosilicate
carbonate on active site of catalyst (TEOS, silica source) and 4-nitroaniline were obtained from
K4 Equilibrium constant for desorption of methanol Chemical Drug House (CDH) Pvt. Ltd., New Delhi, India. Ethanol was
on active site of catalyst purchased from Merck, Darmstadt, Germany. The standards used
kˈ Apparent reaction rate constant for for calibration was Glycerol carbonate (purity > 99.00%) and was
transesterification of glycerol to glycerol acquired from TCI Chemical Pvt. Ltd., India, pyridine (pu-
carbonate, min1 rity > 99.99%), and glycerol (purity > 99.50%) were purchased from
-rGLY Rate of depletion of GLY, mol/mL.min Sigma-Aldrich, Missouri, USA.

2.2. Catalyst preparation

2009). However, reusability and separation of catalyst from the MCM-41 (catalyst support) was prepared using CTAB and TEOS
products are the significant constraints associated with these cat- as surface directing agent and silica source, respectively. In the
alysts (Song et al., 2017). These drawbacks initiate the use of the typical procedure, a clear and homogeneous solution of CTAB was
heterogeneous alkaline active catalyst for the substantial conver- prepared by dissolving 4.8 g of CTAB in 240 ml of distilled water
sion of GLY. Heterogeneous catalysts, such as CaeAl hydrocalumite under continuous stirring. After that, 16 ml of 25 wt% ammonium
(Zheng et al., 2015), SreAl mixed oxide (Algoufi et al., 2017), hydroxide was added into it, and stirred for 5 min. Subsequently,
hydrotalcite (Takagaki et al., 2010), hydroxyapatite (Bai et al., 2011), 40 ml of TEOS was added drop-wise into the solution. The solution
and CaO (Simanjuntak et al., 2011) had been reported in the liter- was stirred for 12 h, and the resultant white material was filtered,
ature for tailoring transesterification reaction to produce GLC. followed by its washing with ethanol and distilled water. There-
However, these catalysts have certain limitations, including long after, the obtained powder was kept for drying in a hot air oven at
reaction time and rapid deactivation rate. Hence, the high surface 100  C for 12 h. Then, the dried material was calcined in the muffle
area and stable catalyst are highly needed to overcome the existing furnace at 550  C for 5 h. The white powder was designated as
barrier. In the last few decades, MCM-41 received much attention in MCM-41.
material science due to its versatile properties such as large pore Lithium (LiNO3), as an active species was incorporated into the
volume, high surface area, high thermal stability, and ease to framework of MCM-41 through wet impregnation technique.
incorporate active sites to enhance catalytic activity (Chen et al., Further, the effect of synthesis conditions i.e. calcination tempera-
2002; Ravikovitch et al., 1997). Bare MCM-41 does not possess tures and metal loading amounts were investigated. The resultant
2
S. Arora, V. Gosu, U.K.A. Kumar et al. Journal of Cleaner Production 295 (2021) 126437

catalysts were labeled as Li/MCM-41 (x%, T), where x and T repre- pore volume.
sent lithium loading amount and calcination temperature in  C,
respectively. 2.4. Transesterification reaction

2.3. Catalyst characterization Transesterification reaction of GLY and DMC was conducted in a
100 mL three-necked round-bottom flask equipped with a reflux
Fourier transforms infrared (FT-IR) spectrometer (PerkinElmer, condenser. The whole system was placed in an oil bath kept on a
Massachusetts, US) was used to analyze the functional group pre- magnetic stirrer (5 MLH Plus, REMI, India) to maintain the desired
sent on the sample. FT-IR was conducted over the range of temperature and stirring speed. In a typical experiment, 5.02 g
4000 cm1 to 400 cm1. Catalyst samples were finely grounded (0.05 mol) of GLY and 14.41 g (0.16 mol) of DMC were charged into
with KBr and pressed in a hydraulic press to obtain a homogeneous the reactor and allowed to heat until the desired temperature was
transparent pallet. The morphology of catalyst samples were attained. After achieving the desired reaction temperature, 0.2 g of
investigated using field emission scanning electron microscope catalyst was added into the flask to initiate the reaction, and the
(FE-SEM) (Nova Nano FE-SEM 450 (FEI), Thermo Fisher Scientific, reaction was preceded for 3 h. After the completion of the reaction,
Massachusetts, US). The samples were mounted on aluminum the catalyst was separated from the mixture using centrifugation
stubs using double-sided tape, coated with gold on Q150T ES (Re12C plus, REMI, India). Subsequently, the product was filtered
(Quorum Technologies Ltd, England) sputter coater and examined and analyzed by gas chromatography (Trace 1310, Thermofisher
under Nova Nano FE-SEM to determine their microscopic Scientific, USA) equipped with a capillary column (Quadrex, 007e5,
morphology at an accelerating voltage of 40 kV. X-ray diffraction is 30 m  0.25 mm X 0.25 mm) and flame ionization detector. The
an analytical technique used to determine the crystal structure and injector and detector temperatures were kept at 290  C and 325  C.
crystalline phases. The small-angle and wide-angle XRD patterns of The oven temperature was initialized at 70  C and elevated to
the samples were recorded on the X’Pert Powder XRD (PANalytical 300  C at the rate of 20  C/min. Pyridine was used as an internal
X’Pert Pro, B.V., Netherlands) system. The uniformly powdered standard for the quantification of GLY and GLC. All the calculations
samples were placed on sample holder, and the analysis was car- were done assuming GLY as a limiting reactant. GLC’s yield, con-
ried out at operating condition of 40 kV and 40 mA using Cu Ka version of GLY, the selectivity of GLC, and turn over frequency of
radiation (l ¼ 1.540 ). The small angle X-ray scattering was catalyst (TOF) were determined using the following Eqs. (1)e(4):
measured in the range of 1.5 < 2q < 10 , whereas wide angle X-ray

Initial moles of GLY in feed  Unreacted moles of GLY


Conversion of GLY ¼ X 100 (1)
Initial moles of GLY in feed

Moles of GLC produced


Yield of GLC ¼ X 100 (2)
Initial moles of GLY in feed

Yield of GLC
Selectivity of GLC ¼ X 100 (3)
Conversion of GLY

Initial moles of GLY X Conversion of GLY X Molecular weight of GLY


TOF ¼ (4)
100 X Mass of catalyst used for the reaction X Reaction time

diffraction was performed in the range of 10 < 2q < 80 . PAN-
alytical software was used to analyze the XRD data. Thermal sta-
bilities of the samples were investigated through the
thermogravimetric analyzer (TGA). TGA of samples were performed
2.5. Kinetic study
on PerkinElmer simultaneous thermal analyzer (STA 6000).
Approximately, 5 mg of sample was placed in an alumina crucible
The obtained values of concentration of the species were used to
and heated at a rate of 20  C/min under the flow of nitrogen from
minimize the least-squares objective function (Eq. (5)). The opti-
ambient temperature to 900  C. The basic strength of the catalyst
mization stopped when the constraints satisfied within the limits
was determined through a Hammett indicator method. Phenol-
of the constraint tolerance. The concentration-time profiles at
phthalein (H_ ¼ 9.3), 2, 4-dinitroaniline (H_ ¼ 15), and 4-
constant temperatures were obtained by using stiff integrators-
nitroaniline (H_ ¼ 18.4) were adopted as indicators to measure
ODE15s from the MATLAB.
the basic strength of synthesized catalysts. The adsorption-
The principle for the evaluation of the rate constants is that, the
desorption isotherms of MCM-41 and Li/MCM-41(5%, 450) was
values of the sum of square error (SSE), root mean square error
obtained on Micromeritics (3Flex Version 4.01, USA). All the sam-
(RMSE), and chi-square (c2) should tend to zero (Eqs. (7)e(9))
ples were degassed at 350  C for 4 h before analysis. The total
(Hamid et al., 2012), whereas the value of the coefficient of deter-
surface area of the samples was calculated utilizing Brunauer-
mination should tend to one (Eq. (6)). This criterion was further
Emmet-Teller (BET) method and Barrett-Joyner-Halen (BJH)
used to authenticate the fit of the model with the estimated
method was employed to determine average pore diameter and
parameters.
3
S. Arora, V. Gosu, U.K.A. Kumar et al. Journal of Cleaner Production 295 (2021) 126437

# water molecule and surface silanols bonds (Sikarwar et al., 2018).


N 
X 2  2
i i i i The bands at 1085 cm1 and 808 cm1 were assigned to condensed
f¼ CGLY;exp  CGLY;pred þ CGLC;exp  CGLC;pred (5)
asymmetric and symmetric silica network (SieOeSi)
i¼1
(Kaiprommarat et al., 2016). The absorption band’s existence at
459 cm1 was due to SieO- bending vibration, indicating the for-
N 
P 2
1 Cexp;j;i  Cpred;j;i mation of hexagonal mesoporous silica structure (Sikarwar et al.,
N
j¼1 2018; Shu et al., 2015; Zhou et al., 2015; Yang et al., 2015). The
R2 ¼ 1  (6)
N 
P 2 band at 1384 cm1 was ascribed to NeO in NO2 3 , while the peak at
1 Cexp;j;i  Ci
N 1641 cm1 was endorsed to the OeH group (Song et al., 2017). Both
j¼1
peaks were strengthened with increase in Li species concentration
(Fig. 1b-f). In contrast, the reflection of the peak at 1384 cm1
2
decreased considerably in the spectra of regenerated catalyst
1 XN  
SSE ¼ C  Cpred;j;i (7) (Fig. 1g) due to leaching of active metal (Li) from the channels of
N j¼1 exp;j;i
MCM-41. The escape of Li metal from the catalyst or the deposition
of GLY molecule in the pores of catalyst reduced the catalytic ac-
N  0:5 tivity towards GLC formation; this was further advocated from the
1 X
RMSE ¼ Cpred;j;i  Cexp;j;i (8) reusability study.
N j¼1
Powder small-angle x-ray scattering (SAXS) of Li/MCM-41 (x%,
450) catalysts at different lithium loaded amount is represented in
N 
P 2 Fig. 2A. In SAXS profile of MCM-41 (Fig. 2A (a)), three characteristic
Cexp;j;i  Cpred;j;i peaks were observed corresponding to plane (1 0 0), (1 1 0), and (2
c2 ¼
j¼1
(9) 0 0) at 2q ¼ 2.0021, 2.4241 and 4.3932 respectively. The reflec-
Nn tion of a broad peak at 2q ¼ 2.0021 displayed the typical feature of
the hexagonal structure of MCM-41, which also justifies the highly-
where, Cexp,j,i is the experimental concentration of component i at
ordered mesoporous structure of silica (Wu et al., 2015). The (1 0 0)
jth number of the experiment data. Cpred,j,i is the simulated con-
plane’s reflection shifted towards a higher scattering angle
centration of component i at jth number of simulated data. The
(2q ¼ 2.04 ) when the Li ions were introduced into the MCM-41
number of the experimental data is represented by N, while n in-
framework. This could be ascribed to the reduction in inter-
dicates the number of proposed kinetic model constants.
planar d spacing at (1 0 0) plane from 45.70 Å to 41.27 Å
(Sikarwar et al., 2018). As a result shortening of a long-range hex-
3. Results and discussion agonal array of mesoporous MCM-41 was observed and it may be
beacuse of the accumulation of non-framework lithium species in
3.1. Characterization the structure of MCM-41, as compared to bare support material
(Fig. 2A (b-f)). This behavior might be addressed to the local
The FT-IR spectrum of prepared catalysts are presented in Fig. 1. distortion of pores of MCM-41 at higher loading of lithium species
In Fig. 1a, broadband at 3433 cm1 was assigned to the SieOH (Tang and Hong, 2016). Additionally, it was also observed that the
surface hydroxyl group due to physically intruded water molecule diffraction intensities of peaks at 2.4241 and 4.3932 decreased
(Sikarwar et al., 2018). The spectra of MCM-41 exhibited clear gradually with an increase in Li-ion loading and finally disappeared
doublet at 2926 and 2855 cm1 corresponding to symmetric CeH in the SAXS spectra of 10% Li loaded MCM-41.
vibrations and asymmetric CH2 vibration of adhered surfactant The wide-angle XRD profiles of bare MCM-41 and Li-ion incor-
molecule (Deshmane et al., 2015). A peak at 1641 cm1 was porated MCM-41, calcined at 450  C are presented in Fig. 2B. The
ascribed to the SieOH group formed due to the freely absorbed XRD pattern of bare MCM-41 depicted a broad hump at 2q ¼ 22 ,
justifies its amorphous nature. No peaks were detected corre-
sponding to lithium oxide and/or lithium nitrate in Li/MCM-41 (1%
and 3%, 450) catalyst, as seen in Fig. 2B (b, c). This may be due to
complete dispersion of lithium species into the pores of MCM-41.
Since lithium species possess lower atomic weight and reflect
weak scattering which was dominated by higher intensities of silica
framework (Song et al., 2017). However, lithium silicates and
lithium oxide phases were detected in XRD profiles bearing lithium
weight percentage greater than three weight percentage. The
characteristic peaks of Li2SiO3 were identified at 18.93 and 26.99
(Pan et al., 2017; Ortiz-Landeros et al., 2011), while the reflection of
the peak at 33.04 was associated with the development of Li2O
phases formed due to the decomposition of lithium nitrate in the
structure of MCM-41 (Puna et al., 2014). The peaks intensified with
an increase in lithium metal (5 wt% - 10 wt%) in the assembly of
MCM-41 as shown in Fig. 2B (d-f), whereas no peaks of lithium
were observed in the XRD profile of reused catalyst (Fig. 2B (g)).
The morphological images of MCM-41 and Li/MCM-41 (5%, 450)
are displayed in Fig. 3. It can be clearly observed that the meso-
structured MCM-41 exhibited evenly dispersed spherical shaped
Fig. 1. FT-IR spectra of (a) MCM-41; (b) Li/MCM-41 (1%, 450); (c) Li/MCM-41 (3%, 450);
monolithic particles (Zhou et al., 2015; Dhokte et al., 2011; Liu et al.,
(d) Li/MCM-41 (5%, 450); (e) Li/MCM-41 (7%, 450); (f) Li/MCM-41 (10%, 450); (g) Li/ 2014). Li incorporated MCM-41 shows the uniform film on the
MCM-41 (5%, 450) regenerated to 4th cycle. spherical shaped particle due to multilayer dispersion of active
4
S. Arora, V. Gosu, U.K.A. Kumar et al. Journal of Cleaner Production 295 (2021) 126437

Fig. 2. XRD Pattern of (a) MCM-41; (b)Li/MCM-41 (1%, 450); (c) Li/MCM-41 (3%, 450); (d) Li/MCM-41 (5%, 450), (e) Li/MCM-41 (7%, 450), (f) Li/MCM-41 (10%, 450), (g) Li-MCM-41
(5%, 450) regenerated to 4th cycle.

Fig. 3. SEM micrographs of (A) MCM-41 (B) Li/MCM-41 (5%,450).

Fig. 4. TGA profile and N2 adsorption-desorption isotherms of MCM-41 and Li/MCM-41 (5%, 450).

metal in the MCM-41 framework (Fig. 3B), elucidating that the Thermal stability of MCM-41 and Li/MCM-41(5%, 450) was
active metal was successfully amalgamated into the framework of investigated through TGA analysis, and their profiles are presented
MCM-41 (Parida and Rath, 2009). in Fig. 4A. The Thermogram of MCM-41 displayed a total weight

5
S. Arora, V. Gosu, U.K.A. Kumar et al. Journal of Cleaner Production 295 (2021) 126437

loss of 7% in the temperature ranging from room temperature to calcined at 450  C, and the results depicted in Fig. 5A. As observed
900  C. However, the initial weight loss (below 200  C) was due to from the experiments, the conversion of GLY escalated (35.24 ± 2.16
loss of water molecules from the pores of MCM-41, while the to 99.00 ± 3.15%) with an increase of active metal loading (1e5 wt%)
further losses (above 200  C) were due to the leftover surfactant on MCM-41 (as shown in Table 2). The conversion of GLY increased
decomposition in the pores of MCM-41. However, Li incorporated due to the improvement of the basicity of catalysts with lithium
MCM-41 shows significant weight loss in the region 250e550  C species incorporation. However, when Li loading increased beyond
due to the decomposition of nitrate spices from lithium nitrate 5 wt%, a slight decrease in yield and conversion of GLC and GLY was
precursor (Liu et al., 2014; Parida et al., 2009). observed, respectively. However, the highest total basicity,
The N2 adsorption-desorption isotherms of MCM-41 and Li/ 11.0 mmol/g (shown in Table 2), was acquired for 10 wt% Li/MCM-41,
MCM-41 (5%, 450) are shown in Fig. 4B. It was noticed that the calcined at 450  C. The catalytic activity was declined even at higher
isotherms depict the characteristic feature of type IV hysteresis loading (>5 wt%), due to multilayer dispersion of lithium species on
loop according to International Union of Pure and Applied Chemistry MCM-41 framework takes place which leads to decrease in surface
(IUPAC) standards, with capillary condensation in the relative area by blocking of pores of MCM-41 channel or agglomeration of
pressure range (P/P0) of 0.3e0.5. This trend particularly signifies active metal takes places. Furthermore, the turn over frequency
that the formed material exhibit typical mesoporous nature. The (TOF) was calculated to investigate the catalyst performance. With
BET surface area, pore volume and average pore diameter of bare an increase in active metal loading amount from 1 wt% to 5 wt%, the
MCM-41 was 861 m2/g, 0.577 cm3/g and 25.431 Å, respectively. TOF value increased from 2.94 ± 0.40 h1 to 8.25 ± 0.45 h1. Further
After incorporation of active metal in MCM-41 framework, the increase of active species on MCM-41 (>5 wt%), TOF value decreased.
surface area and pore volume declined significantly to 254 m2/g This is due to an in-accessibility of the reactant to the active sites
and 0.330 cm3/g, respectively. The decrease in BET surface area and because of multilayer dispersion or may be due to pore blocking.
pore volume might be due to the incorporation of lithium species in Kumar et al. (2015) investigated the GLC synthesis by using CaLa
the framework of MCM-41, and the scattering of metal oxides on catalyst: TOF was estimated in a range of 5.15e5.78 h1. Hence, it is
the boundaries of mesoporous channels. However, the average pore logical to accept that a moderate amount of basic sites are required to
diameter increased to 62.022 Å owing to the formation of metal obtain the maximum GLC yield. Therefore, 5 wt% of Li loading was
oxide bond on the wall of MCM-41 channels (Xu et al., 2017). chosen as an optimum active metal loading on MCM-41 for the
synthesis of GLC.
3.2. Screening of active metal for GLY transesterification
3.3.2. Effect of calcination temperature
A series of catalysts were prepared by impregnating different The impact of calcination temperature on basicity of the catalyst
active metals on MCM-41 for the transesterification of GLY with surface was investigated with varying calcination temperature
DMC. All the experiments were conducted thrice, and confidence from 250  C to 550  C. Fig. 5B and Table 2 showed the yield of GLC
interval were calculated at a ¼ 0.05. Table 1 demonstrates the and conversion of GLY. As expected, GLY conversion increased from
catalytic performance of different active metal loaded on MCM-41. 60 ± 2.58% to 99 ± 3.15% with an increase in a calcination tem-
Bare MCM-41 exhibited only 5% conversion of GLY. The GLY con- perature from 250  C to 450  C. At lower temperature calcined
version (10.56 ± 1.95%) was improved slightly when Ni/MCM-41 catalyst showed a considerably lower performance (60% conversion
(5%, 450) was used as a catalyst. Further, the conversion of GLY of GLY). This may be due to an insufficient amount of basic sites
increased to 20.30 ± 1.53% with Mg/MCM-41 (5%, 450) catalyst. The formation at a lower calcination temperature. Since at lower
conversion of GLY influenced by different active metals including K, calcination temperature, active metal precursor i.e. lithium nitrate,
La, and Ce was also determined. The improvement in trans- may not decompose into lithium oxide species. Besides, TGA
esterification reaction was due to the development of basic sites by investigation also confirmed that lithium nitrate species on MCM-
incorporation of active metal in MCM-41 framework, which is 41 partially decomposed below 450  C, thereby there was less
essential for GLY transesterification reaction. Nonetheless, lithium formation of LiO2 as an oxide layer, which contributes to the basic
species incorporated MCM-41 catalyst showed a better catalytic sites on the catalyst surface, and enhances higher yield and con-
activity. The conversion of GLY increased significantly to 75% version. It was perceived that the yield of GLC and conversion of
because lithium species provides strong basic sites due to its strong GLY was declined when a further increase of calcination tempera-
ion size effect. Therefore, Li/MCM-41 was used as a potential ture from 450  C to 550  C. The observed negative trend may be due
catalyst for the transesterification reaction. to the sintering of active metal in MCM-41 framework at higher
calcination temperature leads to the destruction of basic sites of the
3.3. Preparation parameters of catalyst catalyst (Wang et al., 2015). Hence, calcination temperature plays a
vital role in obtaining the maximum yield and selectivity of GLC. In
3.3.1. Effect of lithium loaded amount this study, 5 wt% Li/MCM-41 calcined at 450  C was adopted as an
The influence of active metal loading was investigated by varying optimized catalyst for investigating the effect of reaction variables
different weight percentage of lithium active species on MCM-41, on transesterification reaction.

Table 1
Effect of various catalysts on conversion of GLY and yield of GLC with an interval of confidence at a ¼ 0.05.

Entry Catalyst Conversion of GLY Interval of confidence Yield of GLC Interval of confidence

1 MCM-41 4.00 0.27 3.20 0.19


2 Ni/MCM-41 10.56 1.95 9.79 1.37
3 Mg/MCM-41 20.30 1.53 15.93 1.05
4 K/MCM-41 40.31 2.01 31.48 1.85
5 La/MCM-41 43.14 1.49 39.03 1.25
6 Ce/MCM-41 54.79 3.01 52.45 2.96
7 K/Mg-MCM-41 68.51 1.56 65.37 1.87
8 Li/MCM-41 75.00 2.23 73.97 2.10

6
S. Arora, V. Gosu, U.K.A. Kumar et al. Journal of Cleaner Production 295 (2021) 126437

Fig. 5. Effect of catalyst preparation parameters on yield and conversion (A): DMC-to-GLY molar ratio: 3, temperature: 90  C, catalyst amount: 4 wt% of Li/MCM-41 (5%, 450),
reaction time: 3 h; (B): DMC-to-GLY molar ratio: 3, temperature: 90  C, catalyst amount: 4 wt% of Li/MCM-41 (5%, 450), reaction time: 3 h.

Table 2
Textural properties of Li/MCM-41.

Catalyst Basicity (mmol/g) Basic strength(H_) Yield of GLC Interval of confidenceb Conversion of GLY Interval of Confidenceb

MCM-41 5.70 H_  7.2 3.20 0.19 4.00 0.27


Li/MCM-41 (1%, 450) 6.80 7.2 H_ 9.8 35.08 2.89 35.24 2.16
Li/MCM-41 (3%, 450) 8.50 7.2  H_ 9.8 60.90 2.13 61.39 3.05
Li/MCM-41 (5%, 450) 9.80 9.8 H_ 15 93.13 3.11 99.00 3.15
Li/MCM-41 (7%, 450) 10.70 9.8 H_ 15 74.13 1.78 80.42 2.35
Li/MCM-41 (10%, 450) 11.00 15 H_ 18 49.29 2.23 53.94 2.61
Li/MCM-41 (5%, 250) 8.00 7.2  H_ 9.8 59.04 1.95 60.06 2.58
Li/MCM-41 (5%, 350) 8.60 9.8 H_ 15 74.48 3.00 75.28 2.36
Li/MCM-41 (5%, 550) 9.10 9.8 H_ 15 79.91 1.05 88.48 2.11
a
Li/MCM-41 (5%, 450) 8.90 9.8 H_ 15 80.85 2.63 81.49 3.11
a
Regenerated for four runs.
b
Interval of confidence at a ¼ 0.05.

3.4. Effect of reaction parameters on transesterification of GLY into of DMC-to-GLY molar ratio was investigated and results are
GLC depicted in Fig. 6B. At an equal molar ratio of DMC-to-GLY (1:1),
only 50.40 ± 1.61% of GLY conversion, and 49.40 ± 1.13% of GLC yield
3.4.1. Effect of catalyst dosage were observed. Further increase of DMC-to-GLY molar ratio (3:1),
The effect of catalyst dose (with reference to GLY weight) was the GLC yield increased and maximized to 93.14 ± 2.52%. It was
investigated by varying catalyst (5 wt% Li/MCM-41) dose from 1 to observed that when DMC-to-GLY molar ratio was increased from
6 wt%, and the results are depicted in Fig. 6A. It observed that GLY 3:1 to 5:1, the yield of GLC slightly declined from 93.14 ± 2.52% to
conversion and GLC yield increased from 45.74 ± 3.15 to 99 ± 1.89% 86.30 ± 1.58%. At a higher molar ratio of DMC-to-GLY (>3:1), the
and 45.34 ± 2.01% to 93.14 ± 2.52%, respectively, on intensifying the catalyst concentration may be diluted due to the addition of a large
catalyst dosage from 1 to 4 wt% (based on GLY mass). This behavior quantity of reactant molecule (DMC) into the reaction mixture,
was anticipated due to the increase in active sites concentration to which leads to a decrease in contact rate between the catalyst and
initiate the transesterification reaction. However, with the further the reactants (Okoye et al., 2017). At a greater molar ratio of DMC-
increase of catalyst dose from 4 wt% to 6 wt% of GLY mass, GLC yield to-GLY, the TOF also decreased from 8.25 ± 0.45 h1 to 8.00 ± 0.32
was declined from 93.14 ± 2.52% to 85.37 ± 1.25%, but the con- h1. Hence, a DMC-to-GLY molar ratio of 3:1 was chosen as an
version of GLY was not influenced significantly. At higher catalyst optimum molar ratio for further studies.
mass, the available excess active basic sites may participate in
transesterification reaction further leading to decarboxylation of
formed GLC into glycidol, due to which, the yield and selectivity of 3.4.3. Effect of reaction temperature
GLC was declined (Algoufi et al., 2017; Liu et al., 2014). Hence the Influence of reaction temperature on the transesterification re-
catalyst dose of 4 wt% based on GLY mass was chosen as an effective action is displayed in Fig. 6C. It is a well-known fact that the vis-
dosage for further studies. cosity of reactant decreases with an increase in reaction
temperature: hence the miscibility of GLY and DMC increases. The
rate of collision between the reactant molecules increases, which
3.4.2. Effect of DMC-to-GLY molar ratio enhances the rate of transesterification reaction towards GLC for-
Transesterification reaction strongly influenced by reactant mation (Algoufi et al., 2017; Okoye et al., 2017; Malyaadri et al.,
concentration, in order to investigate the effect of reactant con- 2011). When the reaction temperature was raised from 60 to
centration, the study was conducted by varying molar ratio of 90  C, the GLC yield increased from 34.18 ± 1.47% to 93.14 ± 2.52%.
DMC-to-GLY. Since the transesterification reaction is reversible in Correspondingly, the TOF value of catalyst risen from 2.92 ± 0.26 to
nature, so to shift the equilibrium towards the forward direction 8.25 h1 ± 0.45. This can be attributed to the well-established fact
(GLC production), an excess amount of DMC is required. The effect that the reaction rate increases with an increase in reaction
7
S. Arora, V. Gosu, U.K.A. Kumar et al. Journal of Cleaner Production 295 (2021) 126437

Fig. 6. Effect of reaction influencing parameters on GLY conversion and GLC yield (A): DMC-to-GLY molar ratio: 3, temperature: 90  C, reaction time: 3 h; (B): Catalyst dosage 4 wt %
of Li/MCM-41 (5%, 450), temperature: 90  C, reaction time: 3 h; (C): Catalyst dosage 4 wt % of Li/MCM-41 (5%, 450), DMC-to-GLY molar ratio: 3, reaction time: 3 h; (D): DMC-to-GLY
molar ratio: 3; reaction temperature: 90  C; catalyst dosage: 4 wt% of Li/MCM-41 (5%, 450); reaction time: 3 h.

temperature (Yadav and Chandan, 2014). However, with a further catalyst’s basicity was dropped from 9.80 mmol/g to 8.90 mmol/g
increase in reaction temperature from 90 to 100  C, a slight (Table 2) after the 4th cycle. The spent catalyst must be regenerated
decrease in the GLC yield was observed. However, the GLY con- with a suitable technique to get a sufficient quantity of basic sites
version was not affected by increasing the reaction temperature on the catalyst surface. However, the present study was not
beyond 100  C. This may be due to decarboxylation of formed GLC explored any regeneration technique but asses the reusability of
into glycidol owing to ring-opening of GLC structure on available the catalyst. Besides, the absence of lithium species was observed in
vacant primary sites into other by-products. Thus, the reaction the spent catalyst XRD profile (Figs. 1g and Fig 2B (g)). It may be
temperature of 90  C has chosen as an adequate temperature for another possible reason for the decrease in catalyst activity because
tailoring the transesterification reaction of GLY into GLC. of the leaching of active sites from the MCM-41 framework. The
reasons, as mentioned earlier, are in good agreement with the
available literature. Malhotra and Ali (2018) investigated lithium-
3.5. Reusability study
doped ceria supported SBA-15 for biodiesel production, the yield
of biodiesel was declined from 98% to 60% after six successive re-
The reusability of the catalyst is an important criterion to eval-
cycles and the same trend was observed due to leaching of active
uate the catalyst stability and its industrial feasibility. The reus-
metal on the surface of catalyst during catalytic reaction.
ability of Li/MCM-41 (5%, 450) was carried out by separating the
catalyst from the product mixture through centrifugation. The
separated catalyst was washed with ethanol, dried in an oven at
100  C for 5 h, and reused for further transesterification reaction.
3.6. Green matrix study
The reusability has been carried before the next subsequent run.
After the 4th cycle, the GLC yield and GLY conversion reduced
The atom economy, environmental factor (E-factor), and process
significantly from 93.14 ± 2.52% to 81 ± 3.01% and 99 ± 1.89% to
mass intensity (PMI) were calculated to assess the efficiency and
85.49 ± 1.85%, respectively, as shown in Fig. 6D.
environmental performance of mesoporous Li/MCM-41 (5%, 450)
The reason behind gradual loss in the catalytic activity has been
catalyst. The atom economy, E-factor, and PMI of the process were
investigated using FTIR, Hammet titration, and XRD analysis for
determined using Eqs. (10)e(12), respectively.
spent catalyst after the 4th cycle. The spent catalyst FTIR spectra
revealed that the intensity of peak at 1384 cm1 (NeO in NO2 3 ) was
Molecular mass of GLC
decreased and because of a decline in the basic site, which is highly Atom economy ¼ X 100 (10)
responsible for reduction in yield and conversion. The aforemen- Molecular mass of reactants
tioned hypothesis was also confirmed with Hammet titration. The
8
S. Arora, V. Gosu, U.K.A. Kumar et al. Journal of Cleaner Production 295 (2021) 126437

values of R2 (greater than 0.9499) suggest that the model best fits
Total waste ðgÞ with experimental data. Table 4 represents the equilibrium con-
E  factor ¼ (11)
Product ðgÞ stants and apparent reaction rate constant evaluated at different
temperatures (60e100  C). It was seen that the values of apparent
where, total waste ¼ total mass of material used in the process-total reaction rate constants increased with an increase of reaction
mass of product formed temperature, and facilitated the concentration of GLY and GLC to
attain equilibrium at a faster rate. However, a negligible difference
Total mass used in the process ðgÞ
PMI ¼ (12) in apparent reaction rate constants was seen at 90  C and 100  C.
Mass of final product ðgÞ The acquired values of apparent rate constants were used to eval-
During transesterification of GLY into GLC, methanol is the only uate the activation energy reaction using the Arrhenius equation
by-product produced, which could serve as a feedstock for bio- (shown in Table 4). Similar findings were reported by pankajakshan
diesel production. The catalyst’s reusability was not considered et al. (Pankajakshan et al., 2018) for the acetylation of GLY on
during the calculation of the atom economy, environmental factor sulfated alumina catalyst. A plot of ln (kˈ) vs 1/T used to calculate
(E-factor), and process mass intensity (PMI) values. Lower the the activation energy of 53.77 ± 3.26 kJ/mol (Fig. S1). Similar, re-
environmental factor (E-factor) and process mass intensity (PMI) sults were reported in the literature for the transesterification of
values, the lesser the waste generation during the process. The GLY into GLC. Devi et al. (2018), Esteban et al. (2015), and Yadav and
atom economy of the process was calculated as 0.64, whereas the E- Chandan (2014) reported the activation energies and are follows
factor and PMI values of GLC synthesis were determined as 1.16 and 39.2 kJ/mol using Ti/SBA-15 catalyst; 28.4 ± 1.5 kJ/mol using po-
2.16, respectively, which demonstrate the less waste generation tassium methoxide catalyst; and 53.77 52.50 kJ/mol employing
during the process. Pradhan and Sharma (2020) reported an E- calcined hydrotalcite supported hexagonal silica catalyst,
factor of 0.074 for GLC synthesis using NiMgOx catalyst by elimi- respectively.
nating solvent reusability and recyclability of catalyst. Olivares et al.
(2020) had observed E-factor of 0.5322 for GLC synthesis using an 3.8. Comprehensive analysis of GLC synthesis routes
apatite-like catalyst. Roy et al. (2020) reported E-factor of 0.153
(waste cooking oil methyl ester) and 0.157 (castor oil methyl ester) Various synthesis routes have been explored to synthesize GLC
for biodiesel production using potassium modified ceria oxide from GLY using various carbonate sources, including carbon diox-
catalysts. ide, urea, phosgene, and dialkyl carbonate. The yield and conver-
sion of the different processes were tabulated in Table 5. Liu et al.
(2018) investigated the catalytic conversion of GLY with CO2 into
3.7. Kinetic study GC using Ce0$98Zr0$02O2, and it was observed a 36.3% GLC yield and
40.90% GLY conversion. Further, this process requires continuous
Transesterification of GLY with DMC into GLC was carried out in water removal to eliminate thermodynamic limitations and ach-
the presence of Li/MCM-41 (5%, 450) as a catalyst in a heteroge- ieve a higher yield. Chaves and Silva (2019) investigated the urea
neous reaction system. The Langmuir-Hinshelwood-Hougen- route to convert GLY into GLC using Sn(OH)2. It obtained 87%
Watson (LHHW) model was adopted to formulate the probable conversion, but these processes need to operate under vacuum to
dual-site mechanism assuming that the rate of external and inter- remove formed ammonia as a by-product. Besides, researchers
nal mass transfer is very high as compared to the surface reaction explored the one-pot green synthesis route i.e., transesterification
rate route (Das and Mohanty, 2019). Hence the model was devel- of glycerol with dialkyl carbonate using homogeneous and het-
oped based on adsorption, surface reaction, and desorption steps. erogeneous catalysts. Ochoa-Go mez et al. (2012) investigated GLC
The heterogeneous reaction precedes in three major steps via (1) synthesis using the transesterification of GLY with DMC in the
adsorption of reactant species on to the active sites of catalyst presence of a homogeneous catalyst. It was found that 98% GLC
surface from bulk phase (2) surface reaction of GLY and DMC onto yield and 99% GLY conversion. However, homogeneous catalysts
the active sites to produce GLC (3) desorption of product molecules provide an excellent catalytic activity interms of yield and con-
(GLC and methanol) from the active sites of catalyst into bulk phase. version, but recovery and regeneration are major constraints. Devi
Eq. (13) was developed from the model, and its elaborated deri- et al. (2018) examined the one-pot green synthesis route using a
vation is provided in SI. heterogeneous catalyst. It was found that the Ti-SBA-15 catalyst has
potential catalytic activity with 82% yield and 94% conversion. The
dCGLY k0 CGLY 0 present study also investigated the feasibility of mesoporous het-
 ¼ 2 ; k
dt CGLC CM erogeneous catalyst (Li/MCM-41) for a one-pot green synthesis
1 þ K1 CGLY þ K2 CDMC þ K3 þ K4 route for GC synthesis. It was observed that the present catalyst has
a potential catalytic activity with a high yield (93.14 ± 2.52%) and
¼ CDMC kSR K1 K2 Ct2 (13)
conversion (99 ± 1.89%).
The kinetic parameters (equilibrium constants and apparent
reaction rate constant) of Eq. (13) were optimized using MATLAB 4. Conclusion
R2015b. The experimental and simulated concentrations at varying
temperatures were compared, and the plots are presented in Fig. 7. Transesterification of GLY with DMC was successfully investi-
The simulation was performed at a constant density system, i.e. at gated using the heterogeneous mesoporous catalyst. 5 wt% Li/
constant temperature and pressure. The ODE solver function was MCM-41, calcined at 450  C illustrates the better basic sites for
called by the main ‘fmincon’ optimization function during the transesterification of GLY to GLC. Li/MCM-41 catalyst showed stable
execution. The values of reaction rate constants were constrained thermal stability up to 900  C. With the incorporation of active
within positive real numbers. metal species (5% Li) in the MCM-41 framework, the surface area
Table 3 summarizes the statistical parameters of the kinetic (861 m2/g to 254 m2/g) and pore volume (0.577 cm3/g to 0.330 cm3/
model. The small value of SSE, RMSE, c2, and relatively higher g) was declined but basic sites were increased (5.70 mmol/g to

9
S. Arora, V. Gosu, U.K.A. Kumar et al. Journal of Cleaner Production 295 (2021) 126437

Fig. 7. Comparison plots of the simulated and experimental concentration of GLY and GLC at different reaction temperatures: (a) 60  C, (b) 70  C, (c) 80  C, (d) 90  C, (e) 100  C at
DMC-to-GLC molar ratio of 3, and 4 wt % of catalyst dosage using Li/MCM-41 (5%, 450) catalyst.

Table 3
Statistical analysis for the mathematical parameter estimation for GLY and GLC.

Component 60  C 70  C 80  C 90  C 100  C
09 07 08 07
GLY SSE 3.51  10 8.65  10 9.71  10 2.70  10 5.11  1007
RMSE 3.34  1004 6.17  1004 9.45  1004 5.84  1004 6.19  1004
c2 9.37  1009 2.30  1006 2.59  1007 7.18  1007 1.36  1006
R2 0.9624 0.9499 0.9503 0.9819 0.9786
GLC SSE 4.00  1009 8.71  1007 9.82  1008 2.74  1007 5.14  1007
RMSE 2.97  1004 6.06  1004 7.48  1004 6.17  1004 8.11  1004
c2 1.07  1008 2.31  1006 2.62  1007 7.31  1007 1.37  1006
R2 0.9680 0.9608 0.9651 0.9766 0.9560

9.80 mmol/g). Langmuir-Hinshelwood-Hougen-Watson kinetic research work confirms that GLC production from GLY through
model was adopted to explain the reaction kinetics of Li/MCM-41 transesterification reaction using lithium loaded MCM-41 and has a
(5%, 450) catalyst in the transesterification reaction. The present great potential to increase the economy of the biodiesel plant.

10
S. Arora, V. Gosu, U.K.A. Kumar et al.
Table 4
Fitted values of equilibrium constants, rate of GLY depletion and activation energy with an interval of confidence at a ¼ 0.05.

Temperature ( C) 60  C Interval of 70  C Interval of 80  C Interval of 90  C Interval of 100  C Interval of Activation energy Ea Interval of
confidence confidence confidence confidence confidence (kJ/mol) confidence

Equilibrium constant K1 11.358 1.389 11.025 1.648 11.109 2.035 11.226 2.035 10.966 1.235 53.77 3.26
K1 11.541 1.856 11.319 2.365 10.897 1.638 11.678 1.638 11.271 1.022
a
kˈˈ 0.123 0.019 0.294 0.031 0.443 0.042 0.955 0.042 0.909 0.025
K1 10.854 2.056 10.487 1.563 10.709 3.014 10.324 3.014 10.065 1.012
K1 10.399 1.956 10.057 2.635 10.071 1.032 9.585 1.032 9.383 1.145
Rate of reaction -rGLY 1.83  1005 5.08  1007 2.19  1005 1.15  1006 2.51  1005 1.38  1006 7.01  1006 1.38  1006 7.62  1006 6.46  1007
(mol/L.min)
a
Apparent reaction rate constant (min1).
11

Table 5
Comprehensive analysis of GLY synthesis routes.

S. Reaction system Catalyst Catalyst dosage Temperature Time Pressure Yield of GLC Conversion of GLY Reference

No. C (h) (%) (%)

1. GLY þ CO2 Ce0$98Zr0$02O2 0.52 g catalyst 150 5 CO2 (3.0 MPa) 36.3 40.9 Liu et al. (2018)
2 GLY þ urea Sn(OH)2 4.9 mol% Sn 140 4 e e 87 Chaves and Da Silva
(2019)
3 GLY þ CO CuCl2 0.60 mmol 130 3e4 O2 (0.7 MPa) and CO e >92 Casiello et al. (2014)
(3.3 MPa
4 GLY þ Ethylene Quaternary salt ionic liquid immobilized on MCM41 0.1 g 80 1.5 e e 94 Cho et al. (2010)
carbonate (RNX-MCM41)
5 GLY þ DMC Triethylamine (homogeneous catalyst) TEA/glycerol molar 90 2.5 e 98 99 Ochoa-Go mez et al.
ratio ¼ 0.1 (2012)

Journal of Cleaner Production 295 (2021) 126437


6 GLY þ DMC Ti-SBA-15 5.5 wt% 87.5 2 e 82 94 Devi et al. (2018)
7 GLY þ DMC Li/MCM-41 (5%, 450) 4 wt% based on GLY wt. 90 3 e 93.14 99 Present work
S. Arora, V. Gosu, U.K.A. Kumar et al. Journal of Cleaner Production 295 (2021) 126437

CRediT authorship contribution statement kinetics modeling. Chem. Eng. J. 200, 532e540.
Herseczki, Z., Marton, G., Varga, T., 2011. Enhanced use of renewable resources:
transesterification of glycerol, the byproduct of biodiesel production. Hung. I.
Shivali Arora: Methodology, Investigation, Writing e original Ind. Chem. 39 (2), 183e187.
draft. Vijayalakshmi Gosu: Validation, Writing e review & editing. Kaiprommarat, S., Kongparakul, S., Reubroycharoen, P., Guan, G., Samart, C., 2016.
U.K. Arun Kumar: Writing e review & editing. Verraboina Sub- Highly efficient sulfonic MCM-41 catalyst for furfural production: Furan-based
biofuel agent. Fuel 174, 189e196.
baramaiah: Conceptualization, Supervision, Writing e review & Kaur, M., Ali, A., 2011. Lithium ion impregnated calcium oxide as nano catalyst for
editing, Project administration. the biodiesel production from karanja and jatropha oils. Renew. Energy 36,
2866e2871.
Kumar, P., With, P., Srivastava, V.C., Gla €ser, R., Mishra, I.M., 2015. Glycerol carbonate
Declaration of competing interest synthesis by hierarchically structured catalysts: catalytic activity and charac-
terization. Ind. Eng. Chem. Res. 54, 12543e12552.
Liu, J., Li, Y., Liu, H., He, D., 2018. Transformation of CO2 and glycerol to glycerol
The authors declare that they have no known competing carbonate over CeO2ZrO2 solid solution-effect of Zr doping. Biomass Bioenergy
financial interests or personal relationships that could have 118, 74e83.
appeared to influence the work reported in this paper. Liu, P., Derchi, M., Hensen, E.J., 2014. Promotional effect of transition metal doping
on the basicity and activity of calcined hydrotalcite catalysts for glycerol car-
bonate synthesis. Appl. Catal. B. 144, 135e143.
Acknowledgments Malhotra, R., Ali, A., 2018. Lithium-doped ceria supported SBA-15 as mesoporous
solid reusable and heterogeneous catalyst for biodiesel production via simul-
taneous esterification and transesterification of waste cottonseed oil. Renew.
The author’s grateful to the Department of Science and Tech- Energy 119, 32e44.
nology (DST), Government of India for financial support under the Malyaadri, M., Jagadeeswaraiah, K., Prasad, P.S., Lingaiah, N., 2011. Synthesis of
Early Career Research (ECR) Award scheme (ECR/2015/000085, glycerol carbonate by transesterification of glycerol with dimethyl carbonate
over Mg/Al/Zr catalysts. Appl. Catal., A 401, 153e157.
Dated 20.08.2016). Ms. Shivali Arora thanks Ministry of Human Ochoa-Go mez, J.R., Go mez-Jime nez-Aberasturi, O., Ramírez-Lo pez, C., Maestro-
Resource Development (MHRD), Government of India for the award Madurga, B., 2012. Synthesis of glycerol 1, 2-carbonate by transesterification of
of a fellowship. glycerol with dimethyl carbonate using triethylamine as a facile separable ho-
mogeneous catalyst. Green Chem. 14, 3368e3376.
Ochoa-Go mez, J.R., Go mez-Jime nez-Aberasturi, O., Maestro-Madurga, B., Pesquera-
Appendix A. Supplementary data Rodríguez, A., Ramírez-Lo pez, C., Lorenzo-Ibarreta, L., Villara n-Velasco, M.C.,
2009. Synthesis of glycerol carbonate from glycerol and dimethyl carbonate by
transesterification: catalyst screening and reaction optimization. Appl. Catal. A-
Supplementary data to this article can be found online at Gen. 366 (2), 315e324.
https://doi.org/10.1016/j.jclepro.2021.126437. Okoye, P.U., Abdullah, A.Z., Hameed, B.H., 2017. Stabilized ladle furnace steel slag for
glycerol carbonate synthesis via glycerol transesterification reaction with
dimethyl carbonate. Energy Convers. Manag. 133, 477e485.
References Olivares, R.D.O., Okoye, P.U., Ituna-Yudonago, J.F., Njoku, C.N., Hameed, B.H.,
Song, W., Sebastian, P.J., 2020. Valorization of biodiesel byproduct glycerol to
Algoufi, Y.T., Akpan, U.G., Kabir, G., Asif, M., Hameed, B.H., 2017. Upgrading of glycerol carbonate using highly reusable apatite-like catalyst derived from
glycerol from biodiesel synthesis with dimethyl carbonate on reusable Sr-Al waste Gastropoda Mollusca. Biomass Convers. Biorefin. 1e13.
mixed oxide catalysts. Energy Convers. Manag. 138, 183e189. Ortiz-Landeros, J., Martínez-dlCruz, L., Gomez-Yan ~ ez, C., Pfeiffer, H., 2011. Towards
Arora, S., Gosu, V., Kumar, U.K.A., Subbaramaiah, V., 2020. A facile approach to understanding the thermoanalysis of water sorption on lithium orthosilicate
develop rice husk derived green catalyst for one-pot synthesis of glycerol car- (Li4SiO4). Thermochim. Acta 515, 73e78.
bonate from glycerol. Int. J. Chem. React. Eng. 18 (3) https://doi.org/10.1515/ Pan, Y., Zhang, Y., Zhou, T., Louis, B., O’Hare, D., Wang, Q., 2017. Fabrication of lithium
ijcre-2019-0078. silicates as highly efficient high-temperature CO2 sorbents from SBA-15 pre-
Bai, R., Wang, S., Mei, F., Li, T., Li, G., 2011. Synthesis of glycerol carbonate from cursor. Inorg. Chem. 56, 7821e7834.
glycerol and dimethyl carbonate catalyzed by KF modified hydroxyapatite. Pankajakshan, A., Pudi, S.M., Biswas, P., 2018. Acetylation of glycerol over highly
J. Ind. Eng. Chem. 17, 777e781. stable and active sulfated alumina catalyst: reaction mechanism, kinetic
Boro, J., Konwar, L.J., Deka, D., 2014. Transesterification of non-edible feedstock with modeling and estimation of kinetic parameters. Int. J. Chem. Kinet. 50, 98e111.
lithium incorporated egg shell derived CaO for biodiesel production. Fuel Pro- Parida, K.M., Rath, D., 2009. Amine functionalized MCM-41: an active and reusable
cess. Technol. 122, 72e78. catalyst for Knoevenagel condensation reaction. J. Mol. Catal. Chem. 310,
Casiello, M., Monopoli, A., Cotugno, P., Milella, A., Dell’Anna, M.M., Ciminale, F., 93e100.
Nacci, A., 2014. Copper (II) chloride-catalyzed oxidative carbonylation of glyc- Pradhan, G., Sharma, Y.C., 2020. Studies on green synthesis of glycerol carbonate
erol to glycerol carbonate. J. Mol. Catal. Chem. 381, 99e106. from waste cooking oil derived glycerol over an economically viable NiMgOx
Chaves, D.M., Da Silva, M.J., 2019. A selective synthesis of glycerol carbonate from heterogeneous solid base catalyst. J. Clean. Prod. 264, 121258.
glycerol and urea over Sn(OH)2: a solid and recyclable in situ generated catalyst. Puna, J.F., Gomes, J.F., Bordado, J.C., Correia, M.J.N., Dias, A.P.S., 2014. Biodiesel
New J. Chem. 43, 3698e3706. production over lithium modified lime catalysts: activity and deactivation.
Chen, H., Dai, W.L., Deng, J.F., Fan, K., 2002. Novel heterogeneous W-doped MCM-41 Appl. Catal. A-Gen 470, 451e457.
catalyst for highly selective oxidation of cyclopentene to glutaraldehyde by Ravikovitch, P.I., Wei, D., Chueh, W.T., Haller, G.L., Neimark, A.V., 1997. Evaluation of
aqueous H2O2. Catal. Lett. 81, 131e136. pore structure parameters of MCM-41 catalyst supports and catalysts by means
Cho, H.J., Kwon, H.M., Tharun, J., Park, D.W., 2010. Synthesis of glycerol carbonate of nitrogen and argon adsorption. J. Phys. Chem. B 101 (19), 3671e3679.
from ethylene carbonate and glycerol using immobilized ionic liquid catalysts. Rokicki, G., Rakoczy, P., Parzuchowski, P., Sobiecki, M., 2005. Hyperbranched
J. Ind. Eng. Chem. 16, 679e683. aliphatic polyethers obtained from environmentally benign monomer: glycerol
Corma, A., Hamid, S.B.A., Iborra, S., Velty, A., 2005. Lewis and Bro €nsted basic active carbonate. Green Chem. 7 (7), 529e539.
sites on solid catalysts and their role in the synthesis of monoglycerides. J. Catal. Roy, T., Sahani, S., Sharma, Y.C., 2020. Study on kinetics-thermodynamics and
234 (2), 340e347. environmental parameter of biodiesel production from waste cooking oil and
Das, B., Mohanty, K., 2019. A green and facile production of catalysts from waste red castor oil using potassium modified ceria oxide catalyst. J. Clean. Prod. 247,
mud for the one-pot synthesis of glycerol carbonate from glycerol. J. Environ. 119166.
Chem. Eng. 7 (1), 102888. https://doi.org/10.1016/j.jece.2019.102888. Shu, Y., Shao, Y., Wei, X., Wang, X., Sun, Q., Zhang, Q., Li, L., 2015. Synthesis and
Deshmane, V.G., Abrokwah, R.Y., Kuila, D., 2015. Synthesis of stable Cu-MCM-41 characterization of Ni-MCM-41 for methyl blue adsorption. Microporous Mes-
nanocatalysts for H2 production with high selectivity via steam reforming of oporous Mater. 214, 88e94.
methanol. Int. J. Hydrogen Energy 40 (33), 10439e10452. Sikarwar, P., Kumar, U.A., Gosu, V., Subbaramaiah, V., 2018. Catalytic oxidative
Devi, P., Das, U., Dalai, A.K., 2018. Production of glycerol carbonate using a novel Ti- desulfurization of DBT using green catalyst (Mo/MCM-41) derived from coal fly
SBA-15 catalyst. Chem. Eng. J. 346, 477e488. ash. J. Environ. Chem. Eng. 6, 1736e1744.
Dhokte, A.O., Khillare, S.L., Lande, M.K., Arbad, B.R., 2011. Synthesis, characterization Simanjuntak, F.S.H., Kim, T.K., Lee, S.D., Ahn, B.S., Kim, H.S., Lee, H., 2011. CaO-
of mesoporous silica materials from waste coal fly ash for the classical Mannich catalyzed synthesis of glycerol carbonate from glycerol and dimethyl carbonate:
reaction. J. Ind. Eng. Chem. 17, 742e746. Isolation and characterization of an active Ca species. Appl. Catal. A-Gen. 401
Esteban, J., Domínguez, E., Ladero, M., Garcia-Ochoa, F., 2015. Kinetics of the pro- (1e2), 220e225.
duction of glycerol carbonate by transesterification of glycerol with dimethyl Song, X., Wu, Y., Cai, F., Pan, D., Xiao, G., 2017. High-efficiency and low-cost Li/ZnO
and ethylene carbonate using potassium methoxide, a highly active catalyst. catalysts for synthesis of glycerol carbonate from glycerol transesterification:
Fuel Process. Technol. 138, 243e251. the role of Li and ZnO interaction. Appl. Catal. A-Gen. 532, 77e85.
Hamid, H.A., Yunus, R., Rashid, U., Choong, T.S.Y., Al-Muhtaseb, A.H., 2012. Synthesis Takagaki, A., Iwatani, K., Nishimura, S., Ebitani, K., 2010. Synthesis of glycerol car-
of palm oil-based trimethylolpropane ester as potential biolubricant: chemical bonate from glycerol and dialkyl carbonates using hydrotalcite as a reusable

12
S. Arora, V. Gosu, U.K.A. Kumar et al. Journal of Cleaner Production 295 (2021) 126437

heterogeneous base catalyst. Green Chem. 2, 578e581. Xie, W., Yang, Z., Chun, H., 2007. Catalytic properties of lithium-doped ZnO catalysts
Tang, K., Hong, X., 2016. Preparation and characterization of Co-MCM-41 and its used for biodiesel preparations. Ind. Eng. Chem. Res. 46, 7942e7949.
adsorption removing basic nitrogen compounds from fluidized catalytic Xu, X., Wu, J., Xu, W., He, M., Fu, M., Chen, L., Ye, D., 2017. High-efficiency non-
cracking diesel oil. Energy Fuels 30, 4619e4624. thermal plasma-catalysis of cobalt incorporated mesoporous MCM-41 for
Teng, W.K., Ngoh, G.C., Yusoff, R., Aroua, M.K., 2014. A review on the performance of toluene removal. Catal. Today 28, 527e533.
glycerol carbonate production via catalytic transesterification: effects of influ- Yadav, G.D., Chandan, P.A., 2014. A green process for glycerol valorization to glycerol
encing parameters. Energy Convers. Manag. 88, 484e497. carbonate over heterogeneous hydrotalcite catalyst. Catal. Today 237, 47e53.
Wang, J.X., Chen, K.T., Wu, J.S., Wang, P.H., Huang, S.T., Chen, C.C., 2012. Production Yang, G., Deng, Y., Ding, H., Lin, Z., Shao, Y., Wang, Y., 2015. A facile approach to
of biodiesel through transesterification of soybean oil using lithium orthosili- synthesize MCM-41 mesoporous materials from iron ore tailing: influence of
cate solid catalyst. Fuel Process. Technol. 104, 167e173. the synthesis conditions on the structural properties. Appl. Clay Sci. 111, 61e66.
Wang, X., Chen, L., Guo, Q., 2015. Development of hybrid amine-functionalized Zheng, L., Xia, S., Lu, X., Hou, Z., 2015. Transesterification of glycerol with dimethyl
MCM-41 sorbents for CO2 capture. Chem. Eng. J. 260, 573e581. carbonate over calcined Ca-Al hydrocalumite. Chin. J. Catal. 36, 1759e1765.
Wu, G., Gao, Y., Ma, F., Zheng, B., Liu, L., Sun, H., Wu, W., 2015. Catalytic oxidation of Zhou, C., Gao, Q., Luo, W., Zhou, Q., Wang, H., Yan, C., Duan, P., 2015. Preparation,
benzyl alcohol over manganese oxide supported on MCM-41 zeolite. Chem. characterization and adsorption evaluation of spherical mesoporous Al-MCM-
Eng. J. 271, 14e22. 41 from coal fly ash. J. Taiwan Inst. Chem. Eng. 52, 147e157.

13

You might also like