You are on page 1of 12

Chemical Engineering Journal 346 (2018) 477–488

Contents lists available at ScienceDirect

Chemical Engineering Journal


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

Production of glycerol carbonate using a novel Ti-SBA-15 catalyst T



Parmila Devi, Umashankar Das, Ajay K. Dalai
Department of Chemical and Biological Engineering, College of Engineering, University of Saskatchewan, 57 Campus Drive, Saskatoon, SK S7N 5A9, Saskatoon, Canada

H I GH L IG H T S G R A P H I C A L A B S T R A C T

• Synthesis of Ti-SBA-15 catalysts for


conversion of glycerol to glycerol
carbonate.
• Effects of Ti loading on structural
properties and catalytic activity of Ti-
SBA-15.
• Ti-SBA-15 (A) demonstrated more
than 10-times greater activity than
SBA-15.
• Lewis acidic nature of Ti was the main
driving force in facilitating the reac-
tion.
• Glycerol carbonate was formed via the
formation of O-methoxy carbonyl in-
termediate.

A R T I C LE I N FO A B S T R A C T

Keywords: The present study was focused on the development of a green process for the production of glycerol carbonate
Glycerol carbonate (GYC) from glycerol and dimethyl carbonate using a Ti-SBA-15 catalyst. Ti-SBA-15 catalysts with varying Si/Ti
Transesterification ratio were synthesized in-situ using sol gel method and characterized using various chemical and spectroscopic
Catalyst techniques to study the effects of Ti incorporation on surface and catalytic properties of Ti-SBA-15 catalysts. The
Mechanism
process parameters were optimized to obtain high glycerol conversion and GYC selectivity. Ti-SBA-15 catalysts
Reusability
with lower Si/Ti ratio demonstrated higher glycerol conversion and GYC selectivity as compared to the catalysts
with higher Si/Ti ratio. A regression model was developed to analyze the correlation between reaction para-
meters and reaction outcomes which suggest that the reaction temperature has the most significant effect on the
glycerol conversion and GYC selectivity. A reaction mechanism portraying the role of Ti-SBA-15 in facilitating
the formation of GYC was presented. GYC was formed via the formation of O-methoxy carbonyl intermediate and
the reaction was catalyzed by the Lewis acidic nature of Ti-SBA-15 catalyst. A kinetic model was proposed based
on the results obtained from this study. In addition, reusability study of the production GYC from glycerol using
Ti-SBA-15 suggests that the process has the commercialization potential.

1. Introduction emissions (GHG) and carbon footprints, the demand for bio-based fuel
alternatives have increased significantly worldwide. The high biodiesel
In view of increasing environmental concerns from the use of fossil production in North America and European countries generates a sur-
fuels and stringent government regulations on greenhouse gas plus of glycerol as a byproduct. Biodiesel production generates around


Corresponding author.
E-mail address: ajay.dalai@usask.ca (A.K. Dalai).

https://doi.org/10.1016/j.cej.2018.04.030
Received 2 November 2017; Received in revised form 3 April 2018; Accepted 6 April 2018
Available online 07 April 2018
1385-8947/ © 2018 Elsevier B.V. All rights reserved.
P. Devi et al. Chemical Engineering Journal 346 (2018) 477–488

100 g of glycerol glut per kg of biodiesel produced [1,2]. The increased 2. Experimental
biodiesel production, broad availability of glycerol at low prices and
need of utilization of chemicals generated from petrochemical sources 2.1. Materials
has created a huge interest in glycerol as a building block [2]. Due to
low oil prices over past few years, there is a mounting pressure on All the chemicals used in the present study were analytical reagent
biodiesel industries to utilize glycerol glut to produce high value pro- grade and used without any purification. Pluronic P123 (Aldrich, MW
ducts to support economic sustainability of biodiesel production. Gly- 5800) was used as a structure-directing agent, tetraethylorthosilicate
cerol has been used as a resource for synthesis of various chemicals (TEOS 98%, Aldrich) was used as a silicon source, and titanium iso-
using different processes like esterification, transesterification, acet- propoxide (TIP 97%, Aldrich) was used as a titanium precursor.
alization, dehydrogenation and polymerization etc. [3–6]. This study Hydrochloric acid (HCl 37%) was used to maintain the solution pH.
focuses on the development of a green process for the conversion gly- Glycerol, DMC, GYC were purchased from Sigma Aldrich. Hydrogen,
cerol to glycerol carbonate (GYC). helium, nitrogen and air with high purity grade (99.99%) were pur-
GYC is a high value product associated with positive physico- chased from Praxair, Saskatoon, Canada.
chemical properties such as low toxicity, low flammability and low
vapour pressure. GYC has wider applications in paint and pharmaceu- 2.2. Catalyst preparation
tical industries. It is used as a monomer for polyurethane formation, a
component of detergents and coatings and a novel component of gas- SBA-15 and Ti-SBA-15 catalysts were synthesized by sol–gel
separation membranes [7–9]. GYC has been synthesized from glycerol method. In a typical sol–gel synthesis, pluronic P123 (9.28 g, 0.016 M)
by transesterification alkyl carbonates [10], carbonylation with urea was dissolved in deionized water (228.6 g) and stirred at 40 °C for 2 h.
[11] and carbonation of glycerol with carbon dioxide (CO2) [12]. The After 2 h, HCl (4.54 g, 0.46 M) was added to the above mixture and
carbonation of glycerol with CO2 provides a low GYC yield and the use stirred till a clear homogeneous solution was obtained (∼2 h). Then, a
of elevated temperature and pressure conditions in the presence of a premixed solution of TEOS (20.83 g, 0.979 M) and TIP (5.58 g, 0.19 M)
catalyst limits wider application of this process [13]. The reaction of was added dropwise to above homogeneous solution and the mixture
glycerol with urea is another method but it gives side products such as was stirred for 24 h at 40 °C. The resulting mixture was subjected to
isocyanic acid and biuret and also the reaction requires the use of va- hydrothermal treatment at 100 °C for an additional 24 h to ensure
cuum to remove ammonia from the reaction mixture to initiate the further framework condensation. The solid products were recovered by
reaction [11]. Alkyl carbonates have been used as alternative transes- filtration and dried in air at 100 °C for 24 h. Finally, the dried products
terification agents to produce GYC at much milder conditions [14]. were calcined at 550 °C for 6 h to obtain Ti-SBA-15 catalyst (Si/Ti
Basic oxide catalysts have been extensively studied for these reac- ratio = 4). The calcined catalyst was designated as Ti-SBA-15 (A).
tions due to their high stability in liquid-phase reactions and presence Ti-SBA-15 (B–E) catalysts with varying Si/Ti ratio were prepared by
of relatively strong basic sites. However, these basic catalysts exhibit changing molar quantities of TEOS and TIP following the above dis-
low activity during recycling [15]. Titanium-silica based catalysts have cussed synthesis procedure. The following gel compositions were used:
attracted considerable attention due to their excellent properties as 0.98 M TEOS/0.024–0.19 M TIP/0.46 M HCl/0.016 M P123/127 M
mildly acidic catalysts for selective catalytic reactions [16]. A study H2O. Ti-SBA-15 catalysts with Si/Ti ratios 8, 16, 24 and 32 were de-
from our research group demonstrates that solid acid catalysts can ef- signated as Ti-SBA-15 (B), Ti-SBA-15 (C), Ti-SBA-15 (D), Ti-SBA-15 (E),
fectively catalyze both esterification and transesterification reactions to respectively. SBA-15 was prepared following the above procedure ex-
produce biodiesel from canola oil [17]. The present work is designed to cept that titanium precursor TIP was not added in the reaction.
examine titanium-SBA-15 (Ti-SBA-15) based solid acidic catalysts for
the transesterification reaction of dimethyl carbonate (DMC) and gly- 2.3. Catalyst characterization
cerol to produce GYC. The incorporation of Ti into the silica matrix can
be performed using two methods (i) post-grafting of Ti on pre-synthe- The BET surface area, the pore volume and the pore diameter of the
sized silicate material (SBA-15) and (ii) in-situ synthesis by incorpora- SBA-15 and Ti-SBA-15 catalysts (A–E) were determined using the N2
tion Ti and Si precursors mixture into acidic solution [18]. Generally, adsorption–desorption isotherm at -196 °C using a Micromeritics ASAP
in-situ method is preferred compared to post-grafting method as it al- apparatus (Model ASAP 2000). The catalysts were degassed under va-
lows the uniform distribution of Ti and prevent the formation of TiO2 cuum at 300 °C for 2 h before the adsorption measurements. The spe-
clusters on SBA-15 surface. However, it is very difficult to control the cific surface area was determined by the standard Brunauer-Emmett-
retention and dispersion behavior of titanium as Ti species are very Teller (BET) method at the relative pressure range of 0.05–0.2. The pore
sensitive to change in synthesis conditions and thermal profile size was evaluated using the Barrett-Joyner-Halenda (BJH) method.
[11,19,20]. Therefore, it is important to study the effects of Ti loading Crystal structure of the Ti-SBA-15 catalysts was analyzed by Bruker
on chemical nature of Ti species in Ti-SBA-15. The aim of the present D8 Advance diffractometer (using Cu-Ka radiation (κ = 1.54 Å) at
study is to investigate the effects of Ti loading on surface properties and 40 kV/40 mA. Low-angle XRD patterns were recorded from 2θ = 0 to
chemical nature of Ti-SBA-15 catalyst and evaluate the catalytic ac- 10° at a scanning rate of 2θ = 3° per min. Wide-angle XRD patterns
tivity of Ti-SBA-15 on the conversion of glycerol to GYC. This study was were recorded by scanning the samples from 2θ = 10 to 60 at a scan-
designed with the following objectives: (i) synthesis of Ti-SBA-15 cat- ning rate of 2θ = 3° per min. The reference patterns of the different
alysts and to study the effects of Ti integration in SBA-15 framework on single metal oxides were obtained from the Powder Diffraction File 2
transesterification of glycerol to GYC; (iii) to establish a correlation (PDF-2) database licensed by the International Center for Diffraction
between Ti loading and surface properties and chemical nature of the Data (ICDD).
catalysts; (iv) optimization of reaction conditions for transesterification Attenuated total reflectance infrared (ATR-IR) spectra were re-
of glycerol to GYC with high yield and selectivity; (v) development of a corded using a Nicolet Magna 850 Fourier transform spectrometer
kinetic model and devise a reaction mechanism for the formation of equipped with a liquid nitrogen cooled narrow band MCT detector.
GYC and (vi) catalyst reusability study. Each spectrum was obtained from the acquisition of 32 scans at 4 cm−1
resolution from 4000 to 400 cm−1. Raman spectra of powdered catalyst
samples (TiO2, SBA-15 and Ti-SBA-15) was recorded using Renishaw
system 2000 spectrometer equipped with argon laser. The 514 nm line
from an argon ion laser operated at a power of 15 mW was used as the
excitation source.

478
P. Devi et al. Chemical Engineering Journal 346 (2018) 477–488

The surface morphology of the SBA-15 and Ti-SBA-15 catalysts GYC yield were calculated using Eqs. (2)–(4), respectively.
(A–E) was determined by scanning electron micrograph (SEM). Co−Ci
Scanning electron microscopy (SEM) images were recorded on a Carl Conversion (%) = × 100
Co (2)
Zeiss Evo-40 SEM operating at 10 kV. Samples were mounted on alu-
minum stubs using adhesive carbon tape and gold splutter coated to CGYC
Selectivity (%) = × 100
reduce charging. CTP (3)
The composition of Si and Ti in Ti-SBA-15 (A and B) catalysts was
Conversion × Selectivity
determined using an inductively coupled plasma optical emission Estimated yield (%) =
100 (4)
spectroscopy (Thermo Fisher Scientific iCAP 7000 series) in the fused
catalysts samples. The whole rock analysis method was used for the where Co is the initial amount of glycerol in the reaction mixture, Ci is
fusion of fresh Ti-SBA-15 (A–E) and recycled Ti-SBA-15 (E). Briefly, the final amount of the glycerol in the reaction mixture, CGYC is the
0.1 g of the catalyst was fused with lithium metaborate in Claisse Ox amount of GYC formed and CTP is the amount of all the products
Automatic Fusion Machine at 1000 °C. After fusion and adequate formed.
cooling of residues, solutions were made up to 100 mL by adding 0.04 N
HNO3 acid and the samples were analyzed using ICP-OES. 2.6. Analytical method

2.4. Experimental design The amount of glycerol and GYC was determined by GC (Agilent
7890A) equipped with a flame ionizing detector (FID). The stabilwax
The experiments were designed by a central composite design (CCD) capillary column with 30 m length and 0.25 mm inner diameter with
defined under response surface methodology (RSM) using Design 0.5 μm film thickness was used. The oven temperature was pro-
Expert software (version 9). Three independent variables (reaction grammed to begin at 40 °C for 2 min and ramped to 280 °C at 10 °C/min
temperature (A), glycerol to DMC molar ratio (B) and catalyst loading with a final hold time of 5 min. Sample (1 μL) was injected into the
(C) were considered to design the experiments in order to investigate column with a split ratio of 10:1. Helium was used as a carrier gas. The
the effects of operational parameters on the conversion of glycerol to results have an error of ± 5% throughout the experiments. The for-
GYC. The range of the independent variables was selected based on mation of GYC was confirmed by GC, GC–MS and 1H NMR. Bruker
preliminary studies and literature review. The variables were set at five- AM500 1H NMR spectra of the samples was recorded in deuterated
level and the number of experiments were set to 20 (Table S1). The chloroform (CDCl3) solvent using a 500 MHz NMR spectrometer with a
individual and interaction effects of independent variables such as re- BBO probe.
action temperature (65–110 °C), glycerol to DMC molar ratio (1:1–1:5)
and catalyst loading (1–10 wt%) to the predicted response parameter 2.7. Catalyst reusability study
(GC conversion and GYC selectivity) were determined. The mathema-
tical relationship between independent variables and response para- The reusability of optimized catalyst Ti-SBA-15 (A) was studied by
meters was analyzed by a second-order polynomial equation (Eq. (1)) conducting three runs in duplets at the optimized reaction conditions
and optimized. obtained from the RSM optimization study. After each run, the catalyst
was filtered and washed with methanol (3 × 10 ml) and dried at 100 °C
3 3 3 3
βii Xi2 for 12 h. In a batch reaction, there was around 5% loss of catalyst
Y = βo + ∑ βi Xi + ∑ ∑ βij Xi Xj + ∑
i=1 i=1 j=i+1 i=1 (1) during filtration, drying and handling. Therefore, the loss of catalyst
was made up with reused catalyst from another batch.
where Y is the predicted response (glycerol conversion or GYC se-
lectivity); Xi and Xj are independent variables and βo is the model 3. Results and discussion
constant. βi and βii are the linear and quadratic effect of the in-
dependent variables, respectively. βij is the linear interaction effect 3.1. Catalyst characterization
between the independent variables. The significance level of model and
interaction among variables were estimated by analysis of variance The effects of Ti loadings on surface properties and chemical nature
(ANOVA). of Ti species in Ti-SBA-15 catalysts were studied using different che-
mical and spectroscopic characterization techniques such as BET sur-
2.5. Experimental set-up and reaction procedure face area, elemental analysis, XRD, FTIR, Raman spectroscopy and
SEM.
The experiments for conversion of glycerol to GYC were performed The N2 adsorption–desorption isotherms of SBA-15 and Ti-SBA-15
according to the design of experiments (Table S1). The reactions were (A–E) catalysts with varying Si/Ti ratios are presented in Fig. 1. The
performed under nitrogen atmosphere in a 50 mL three-neck round catalysts A–E and SBA-15 showed a type IV isotherm with type H1
bottle flask equipped with a reflux condenser and a thermometer. hysteresis loop at the relative pressure P/Po ranging from 0.5 to 0.9.
Uniform heating and stirring of the reaction mixture was maintained This indicates that all the catalysts are mesoporous in nature. In Ti-SBA-
using temperature controlled oil bath equipped with a magnetic stirrer. 15 catalysts (A–E), the hysteresis loop is wide open beyond the P/Po
In a typical procedure, a mixture of glycerol (5 g, 0.054 M, 1 mol region of 0.8, indicating the presence of intra-aggregate voids of TiO2
equiv.), DMC (14.68 g, 0.163 M, 3 mol equiv.) and catalyst (500 g, particles. The H1 hysteresis loop ends with horizontal plateau in-
10 wt%) was heated at 65 °C and stirred at 600 rpm for 7 h. After 7 h, dicating narrow pore size distribution and the presence of ordered
the sample was collected from the reaction mixture and the conversion mesoporous structure.
of glycerol to GLY was monitored by gas chromatograph (GC). The The textural properties of the catalysts including BET surface area,
round bottom flask was cooled to room temperature and catalyst was total pore volume and pore diameter were calculated from desorption
separated from reaction mixture by filtration. The catalyst was re- isotherms and are summarized in Table 1. The surface area of SBA-15
peatedly washed with methanol (3 × 10 mL) and the filtrate was eva- was found to be 812 m2/g which is within the range of 500–1300 m2/g
porated using a rotary evaporator to obtain crude GYC. High vacuum reported in the literature for good quality SBA-15 material [21]. The
distillation of crude GYC was carried out to obtain pure product. The data in Table 1 shows that the Ti-SBA-15 catalysts A–E have higher
purity of GYC sample was ensured by GC and 1H NMR. surface area than SBA-15 which suggests that titanium integration into
The conversion of GC to GYC, the selectivity of GYC formation and the silica framework has contributed to an increase in the surface area

479
P. Devi et al. Chemical Engineering Journal 346 (2018) 477–488

results indicate that Ti incorporation in SBA-15 framework was well


accomplished up to a Si/Ti ratio of 16. Further Ti loading beyond this
Si/Ti ratio led to the formation of a layer in the catalyst surface. This
phenomena was further evidenced from XRD analysis that is discussed
later.
The total pore volume of the catalysts A–E was found in the range of
0.82–1.69 cm3/g whereas SBA-15 had a total pore volume of 1.03 cm3/
g (Table 1). The results indicate that the integration of Ti into the silica
frame-work in SBA-15 led to an increase in the pore volume when
compared against SBA-15 except for the catalyst E. The lower total pore
volume of catalyst E may be due to the aggregation of excess Ti mi-
crocrystals in the pores. The micropore volume of SBA-15 and the
catalysts A–E was calculated using t-method. The results in Table 1
indicate that the integration of Ti into the silica frame-work reduced the
micropore volume of the catalysts A, B and D (0.016–0.02 cm3/g) as
compared to SBA-15 (0.058 cm3/g) expect for the catalysts E (0.12 cm3/
g) and C (0.07 cm3/g). The decrease in the micropore volume of a
catalyst is attributed due to the incorporation of metal in micropores
[22]. The pore diameter of the catalysts A–E was found lower than SBA-
15. The above results suggest that Ti is present in Ti-SBA-15 catalysts in
Fig. 1. N2 adsorption–desorption isotherm of SBA-15 and Ti-SBA-15 (A–E) two phases. Initially Ti is incorporated into the silica frame to form a Si-
catalysts. O-Ti network and the excess Ti aggregates as anatase microcrystals
outside channels. This inference was also supported from the XRD data
Table 1 of the catalysts. The existence of a dual phase of Ti could be responsible
Textural properties of SBA-15 and Ti-SBA-15 (A–E) catalysts. for variation in the BET surface area, pore volume and pore diameter of
the catalysts.
Catalyst Si/Ti BET surface Total pore Micropore Pore
ratio area (m2/g) volume volume (cm3/ diameter The composition of Si and Ti content in two catalysts Ti-SBA-15 (A)
(cm3/g) g) (nm) and Ti-SBA-15 (B) was determined by elemental analysis. The Si/Ti
ratio in Ti-SBA-15 (A) and Ti-SBA-15 (B) was found to be 3.81 and 7.89,
SBA-15 – 812 1.03 0.06 5.1
respectively (Table S2) against expected values of 4 and 8, respectively.
Ti-SBA-15 4 1000 1.24 0.02 5.0
(A) The low-angle XRD plot of SBA-15 and Ti-SBA-15 (A–E) catalysts
Ti-SBA-15 8 1245 1.69 0.02 4.5 are shown in Fig. 2a. All the samples showed a diffraction peak of high
(B) intensity at 2θ = 1.1 corresponding to an ordered mesoporous structure
Ti-SBA-15 16 1319 1.49 0.07 4.5 and one broad diffraction peaks of low intensity 2θ = 1.9–2.3. These
(C)
Ti-SBA-15 24 1212 1.64 0.019 4.4
three peaks represent d100 (h), d110 (l) and d200 (k) planes, respectively
(D) and are characteristics of a hexagonal ordered mesoporous structure
Ti-SBA-15 32 846 0.82 0.12 3.8 (p6mm symmetry) [22]. These results suggest that incorporation of Ti
(E) into SBA-15 framework did not affect the standard structure of SBA-15
and the P6mm hexagonal symmetry structure was maintained in Ti-
SBA-15 samples irrespective of variation in Ti loading [21]. Typical
of Ti-SBA-15 catalysts. There is a gradual increase in the BET surface
wide-angle X-ray diffraction (XRD) patterns of TiO2, SBA-15 and Ti-
area of the catalysts (EBET(s) < DBET(s) < CBET(s)) with an increase in
SBA-15 with different Ti loadings are shown in Fig. 2b. Wide angle XRD
the wt% integration of Ti in silica framework in SBA-15 (decrease in Si/
patterns show the characteristic diffraction peak of TiO2 at 2θ value of
Ti ratio). However with further increase in Ti loading, the BET surface
25, 38, 48, and 54° and a broad diffraction peak around 2θ = 23° due to
area was reduced gradually CBET(s) > BBET(s) > ABET(s) which could be
presence of amorphous silica phase in SBA-15 and Ti-SBA-15 catalysts.
due to the aggregation of Ti microcrystals on the catalyst surface. The

Fig. 2. XRD plot of SBA-15 and Ti-SBA-15 (A–E) catalysts (a) low angle; (b) wide angle.

480
P. Devi et al. Chemical Engineering Journal 346 (2018) 477–488

Fig. 4. FTIR spectra of TiO2, SBA-15 and Ti-SBA-15 (A–E) catalysts.


Fig. 3. Raman spectra of TiO2, SBA-15 and Ti-SBA-15 (A–E) catalysts.
Supplementary information). It can be seen from Fig. S1 that very
These results suggest that the incorporation of Ti in silica matrix lead to strong Brønsted (1545 and 1640 cm−1) and strong Lewis (1620 and
the formation of bulk anatase phase and resulted in the appearance of 1455 cm−1) acid sites are absent in Ti-SBA-15 (A–E) catalysts. How-
different lines associated to crystalline anatase TiO2 [19]. The intensity ever, weak Lewis acid sites were observed around 1484 and 1580 cm−1,
of these peaks was found to decrease with an increase in Ti loadings which might be associated with the presence of Ti4+ ions [26]. In ad-
ratio due to difference in the mass of Ti present in different samples. It dition, hydrogen-bonded pyridine peaks around 1595 and 1445 cm−1
was confirmed from the XRD plot that excess amount of Ti tends to form were observed in SBA-15 and Ti-SBA-15 (A–E) catalysts.
an extra-framework of anatase TiO2 crystals that was further identified The surface morphology of SBA-15 and Ti-SBA-15 catalysts was
from Raman spectroscopy. analyzed using SEM and the results are presented in Fig. 5. Fig. 5 shows
Raman spectra of SBA-15 (Fig. 3) showed three peaks in the region that both SBA-15 and Ti-SBA-15 contain short rod-like morphologies,
of 480, 790 and 925 cm−1 which were due to asymmetric stretching indicating that the original shape of pure SBA-15 is reserved in Ti-SBA-
(SieOeSi) vibration, symmetric stretching (SieOeSi) vibration and 15. After incorporation of Ti, the formation of small bead of Ti species
framework defects due to surface silanol groups, respectively [23–24]. was observed on short rod-like structures while the formation of large
It was observed that the silanol peak in SBA-15 at 925 cm−1 shifted to aggregated particles was rarely observed indicating the uniform dis-
990 cm−1 in Ti-SBA-15 (A–E) catalysts due to resonance Raman effects tribution of Ti species.
of framework Ti species. This peak can be assigned to asymmetric
stretching vibration of SieOeTi bond. Interestingly, the intensity of this 3.2. Catalyst activity study
peak decreased with an increase in Ti content in Ti-SBA-15 (A–E) cat-
alysts due to change in amount of extra-framework of Ti species. These 3.2.1. Screening of Ti-SBA-15 catalysts
results can be explained based on the consideration that titanium ions Ti-SBA-15 catalysts (A–E) with varying Si/Ti molar ratio (4–32)
should be located in the framework in isolated form in Ti-SBA-15 cat- were screened for the conversion of glycerol to GYC and their catalytic
alysts but at higher loadings some of titanium ions exist as extra-fra- activity was compared against TiO2 and SBA-15 (Fig. 6). Ti-SBA-15
mework Ti species. These extra-framework Ti species absorb the Raman catalysts A–E showed much higher catalytic activity as compared to
scattering and diminish the Raman signal intensity due to Ti framework TiO2 and SBA-15 which indicate that Ti plays an important role in fa-
as the formation of bulk anatase TiO2 crystal phase was increased at cilitating the conversion of glycerol to GYC. The conversion of glycerol
higher Ti loadings [24]. In Ti-SBA-15 (A–E) catalysts, a strong signal at to GYC was increased with decrease in Si/Ti molar ratio of the Ti-SBA-
144 cm−1 and peaks at 395, 515, and 638 cm−1 confirm the presence of 15 catalysts. The conversion of glycerol to GYC was decreased from
anatase phase TiO2 crystal form. 81% to 41% with an increase in Si/Ti molar ratio from 4 to 32, while no
FTIR spectra of SBA-15 and Ti-SBA-15 (A–E) catalysts are shown in significant change in selectivity of GYC was observed with an increase
Fig. 4. The broad band in the region of 800 to 1080 cm−1 attributed to in Si/Ti molar ratio. Among the catalysts A–E, Ti-SBA-15 (A) demon-
SieO bonding. The peak at 960 cm−1 belongs to symmetric SieOH strated high glycerol conversion and high selectivity, therefore it was
vibrations and peak at 1072 cm−1 belongs to symmetric SieOeTi vi- chosen for further experiments to optimize the reaction parameters.
brations [25]. The overlapping of vibrations of TieOeSi and SieOH
bonds led to the appearance of a band centered at 1050 cm−1. It was 3.2.2. Optimization of reaction parameters
observed from Fig. 4 that intensity of the peak at 1050 cm−1 increased A total of 20 experiments were designed using CCD in Design Expert
with decrease in Ti loadings in the catalysts Ti-SBA-15 (A) to Ti-SBA-15 software (version 9) and the detailed experimental plan is presented in
(E). Increase in the intensity of peak indicates the formation of TieOeSi Table S1. The effects of variables such as reaction temperature
bond due to the interaction of Ti precursor with silanol group in silica (65–110 °C), glycerol to DMC ratio (1:1–1:5) and catalyst loading (1 wt
matrix. Interestingly, an increase in the intensity of this band was %–10 wt%) on the glycerol conversion and GYC selectivity were studied
pronounced at lower Ti loading (higher Si/Ti ratio) whereas a very and the results were fitted to a second-order polynomial model. The
slight change was observed at higher titanium loading. Pyridine ad- fitting of model was expressed by comparing the experimental and
sorbed FTIR spectra of SBA-15 and Ti-SBA-15 (A–E) was recorded to predicted responses.
characterize the nature of the acid sites of the catalysts (Fig. S1, This model predicts glycerol conversion (%) and GYC selectivity

481
P. Devi et al. Chemical Engineering Journal 346 (2018) 477–488

Fig. 5. SEM images of SBA-15 and Ti-SBA-15 catalysts.

9. The glycerol conversion was found to be increased with an increase


in catalyst loading and glycerol to DMC molar ratio at temperature
65 °C and 87.5 °C, respectively. However at 110 °C, glycerol conversion
increased with an increase in catalyst loading (1–5 wt%) and glycerol to
DMC molar ratio (1–3), while further increase in catalyst loading
(> 5 wt%) and glycerol to DMC molar ratio (> 3) lead to a decrease in
glycerol conversion. This behaviour indicates that an optimum combi-
nation of reaction variable (i.e. temperature, catalyst loading and gly-
cerol to DMC molar) is required to achieve the desired response and an
increase in variable value above optimum negatively affect the desired
response.
The interaction plot of catalyst loading at different temperature and
glycerol to DMC molar ratio is shown in Figs. 8 and 9. It was observed
that glycerol conversion significantly increased with an increase in
catalyst loading up to 5 wt% at different temperatures and glycerol to
DMC molar ratio while further increase in catalyst loading does not
have much effect on glycerol conversion. The interaction of glycerol to
DMC molar ratio with reaction temperature and catalyst loading shows
Fig. 6. Screening of catalysts for conversion of glycerol to GYC. Reaction con- an enhancement in the glycerol conversion as the DMC to glycerol
dition: glycerol: DMC molar ratio – 1:3; catalyst loading – 10 wt%; temperature molar ratio increases. The conversion of glycerol increased with an
– 65 °C; reaction time – 7 h. increase in the reaction temperature at different catalyst loadings and
approached to constant at higher catalyst loadings.
versus three main effects (linear terms), three two-factor effects (in- Quadratic interaction reflects the cumulative effects of varying
teraction terms) and two curvature effects (quadratic terms). The effects temperature, glycerol to DMC molar ratio and catalyst loading on re-
of the reaction parameters (reaction temperature, glycerol to DMC sponse variables, i.e. glycerol conversion and GYC selectivity.
molar ratio and catalyst loading) on glycerol conversion (%) and GYC According to this quadratic interactions the following conditions were
selectivity (%) are shown in Fig. 7. The results indicate that glycerol found optimum: reaction temperature (87.5 °C), catalyst loading (5.5 wt
conversion and GYC selectively increased with an increase in tem- %) and glycerol to DMC molar ratio (1:5). Therefore, these conditions
perature up to 87.5 °C, however further increase in temperature had led were considered optimal for GYC production in high yield and se-
to a reduction in GYC selectivity. The reaction was accelerated rapidly lectivity. The effect of temperature was found to be most important
above 65◦C which could be due to co-produced methanol from the parameter on glycerol conversion and GYC selectivity.
reaction system (above boiling point of methanol) and shifting the re-
action toward the formation of GYC.
3.2.3. Statistical studies
The effect of glycerol to DMC molar ratio varying from 1 to 5 was
The statistical significance of the model was analyzed by ANOVA
investigated (Fig. 7). The glycerol conversion and GYC selectivity in-
which gives useful information about the interaction effect of reaction
creased with an increase in glycerol to DMC molar ratio from 1 to 3.
variables (reaction temperature, glycerol to DMC molar ratio and cat-
Further increase in glycerol to DMC molar ratio > 3 does not have
alyst loading) with response variables (glycerol conversion and GYC
much effect on glycerol conversion, however a significant increase in
selectivity) as well as their significance. An F-value (analysis factor) of
GLY selectivity was increased.
101.04 and very low p-values (0.00011 for glycerol conversion and <
The effects of catalyst loading on glycerol conversion and GYC se-
0.0001 for GYC selectivity) implies that the model was highly sig-
lectivity were investigated by varying Ti-SBA-15 (A) loading from 1 to
nificant within 95% confidence level. The “Lack of Fit F-value” of the
10 wt%. Glycerol conversion and selectivity were increased with an
model of 8.79 and 197.77 were observed for glycerol conversion and
increase in the catalyst amount from 1 to 5.5 wt%. However further
GYC selectivity, respectively indicating that there is only 0.11 and
increase in catalyst loading from 5.5 to 10% had led to a decrease in
0.01% chances for occurrence of “Lack of Fit F-value” due to noise
GYC selectivity due to the formation of side products.
(Tables S3 and S4). The values of the determination coefficient R2
The interaction plots of the effects of different operational para-
(0.98), the adjusted R2 (0.97) and predicted R2 (0.90) indicate that the
meters on glycerol conversion and selectivity are shown in Figs. 8 and
quadratic response surface model is appropriate for predicting the

482
P. Devi et al. Chemical Engineering Journal 346 (2018) 477–488

Fig. 7. Effect of different reaction parameters on glycerol conversion and GYC selectivity.

performance of glycerol conversion. 3.2.4. Reaction mechanism and model development


Quadratic model was used to study the surface response for glycerol The reaction products formed at optimized reaction conditions
conversion and GYC selectivity using factors A, B, C, AB, BC, AC, A2, B2, using a Ti-SBA-15 (A) catalyst were quantified using GC and the
C2 and intercept, which were analyzed as function of model where A, B structural identities of the products were confirmed by GC–MS and 1H
and C represent reaction temperature, glycerol to DMC molar ratio and NMR spectroscopy. The reaction products were identified as GYC,
catalyst loading respectively. The quadratic model prediction for the methanol and 1-(o-methoxy-carbonyl) glycerol. It was observed that the
response variables (glycerol conversion and GYC selectivity) were conversion of glycerol to GYC progressed via 1-(o-methoxy-carbonyl)
found statistically valid. Final quadratic equation for glycerol conver- glycerol intermediate. Fig. 10 shows that GYC obtained from this pro-
sion and GYC selectivity was established, where positive sign before cess was similar to the GYC standard obtained from Sigma-Aldrich. A
coefficient suggests synergistic effect of factor towards response and plausible reaction mechanism for the conversion of glycerol to GYC has
negative sign shows an antagonistic effect. been presented in Fig. 11.
The catalytic reaction was progressed via adsorption, surface reac-
Conversion = +91.25 + 7.01∗ A+ 13.14∗ B+ 12.73∗C−1.37∗AB tion and desorption mechanism on Ti-SBA-15 surface. The first step
−0.20∗AC−2.95∗BC−0.94∗A2−9.44∗B2−7.94∗C2 (5) involves the adsorption of glycerol and DMC on catalyst surface
through hydrogen bonding interactions. It is speculated that Lewis acid
sites of Ti+4 activates the carbonyl group of DMC as evidenced from Py-
Selectivity = +76.81−3.54∗A + 6.03∗B−4.68∗C −1.84∗AB + 0.50∗AC
FT-IR (Fig. S1) and oxygen framework of silica (SieOeTi) (weak con-
+ 0.59∗BC + 0.30∗A2 + 2.45∗B2 + 1.62∗C 2 (6)
jugate base) activates the hydroxyl group of glycerol. The second step
involves the surface reaction between glycerol and DMC in which hy-
The observed and predicted values of glycerol conversion and GYC
droxyl group of glycerol reacts at the activated carbonyl carbon of DMC
selectivity are presented in Table S1. It was found that the experimental
forming a 1-(o-methoxy-carbonyl)glycerol complex with a loss me-
values for glycerol conversion and GYC selectivity were in good
thanol molecule. The next step involves the cyclization and re-
agreement with the predicted values obtained by the model. The sig-
arrangement reactions that lead to the formation of GYC and subse-
nificance of the model was also confirmed by linearity between normal
quently losing another molecule of methanol. Finally, GYC is desorbed
probability graphs i.e. the predicted data is in good agreement with
from the catalyst surface and the regenerated catalyst further
observed (Table S1).

483
P. Devi et al. Chemical Engineering Journal 346 (2018) 477–488

Fig. 8. Interaction plot of different reaction parameters and their effect on glycerol conversion.

participates in the catalytic reaction. Various steps discussed above for dCG
−rG = − ′ CM . S . CGYC . S
= kSR CG . S ·CDMS . S−kSR
the catalytic conversion of glycerol to GYC on Ti-SBA-15 surface can be dt (12)
explained by following equations:
dCG kSR {K1 K2 CG. CDMS−(K3 K 4 CM . CGYC / K SR)} Ct2
Adsorption of glycerol on catalyst vacant sites can be given by Eq. − =
(7) dt (1 + K1 CG + K2 CDMC + K3 CM + K 4 CGYC )2 (13)

K1 When reaction is away from equilibrium


G + S ⇔ G. S (7)
dCG kr wCG CDMS
Adsorption of DMC on catalyst vacant sites can be given by Eq. (8) =− =
dt (1 + K1 CG + K2 CDMC + K3 CM + K 4 CGYC )2 (14)
K2
DMC + S ⇔ DMC . S (8) where kr w = kSR K1 K2 Ct2 ; w is weight of catalyst
If the adsorption and desorption constants are very small, then Eq.
The surface reaction of glycerol and DMC on catalyst surface leads (7) reduces to
to the formation of GYC and methanol.
dCG
K SR − = kr wCG CDMC
G. S + DMC . S ⇐⇒ M . S + GYC . S (9) dt (15)

The desorption reaction of methanol and glycerol and GYC from A large amount of DMC is used in the reaction, therefore
catalyst surface can be given by Eqs. (10) and (11). CDMC ≈ CDMC.O can be assumed. The above equation can be written in
terms of fractional conversion.
K3
MS ⇔ M + S (10) dXG
− = k ′ (1−XG )
K4
dt (16)
GYC . S ⇔ GYC + S (11)
where k ′ = kr wCDMC . O
The model was developed based on the following assumptions: (i) By integrating the above equation and constant initial glycerol
All the sites on the catalyst surface have similar physical and chemical concentration, the final expression becomes.
properties; (ii) The rate of non-catalyzed reactions is slower in com-
−ln(1−XG ) = k ′t (17)
parison to the catalyzed reactions, thus considered negligible; (iii)
Surface reaction controls the rate of reaction; and (iv) The adsorption of The reaction kinetics of the conversion of glycerol to GYC was
reactants and desorption of products occur very fast. studied at different reaction temperatures (65–110 °C) and a reaction
As, it is assumed that the rate of reaction is controlled by surface time ranging between 1 h and 7 h using Ti-SBA-15 (A) as a catalyst at
reactions only, then the rate of reaction can be given by the following optimized reaction conditions (catalyst loading (5.5 wt%) and glycerol
equation: to DMC molar ratio (1:5)) and results are shown in Fig. 12. It can be

484
P. Devi et al. Chemical Engineering Journal 346 (2018) 477–488

Fig. 9. Interaction plot of different reaction parameters and their effect on GYC selectivity.

observed from Fig. 12 that glycerol conversion increased rapidly up to against 1/T (K) was plotted (Fig. 12, inset) to obtain the apparent en-
2 h and then the conversion was sluggish with further increase in re- ergy of activation. The apparent energy of activation of 39.2 kJ/mol
action time. While the selectivity of GYC was initially constant till 5 h shows that the reaction is intrinsically kinetically controlled.
and then decreased with an increase with reaction time due to the
formation of side products. A plot of ln (1-XA) versus time was made at 3.2.5. Catalyst reusability study
different temperatures (Fig. 12). It is seen that the data fit well (% fit The catalyst reusability study was performed to confirm its stability
error ± 1.5%) and hence validate the model. Arrhenius plot of ln k and the economic feasibility of the catalyst. The spent catalysts were

Fig. 10. Comparison of NMR plot of GYC obtained in the present study with GYC standard (Sigma Aldrich).

485
P. Devi et al. Chemical Engineering Journal 346 (2018) 477–488

Fig. 11. A proposed reaction mechanism for the reaction of glycerol and DMC to produce GYC using Ti-SBA-15 catalyst.

Fig. 12. Kinetics fit of experimental data obtained at different temperatures.


Reaction condition: Temperature – 65–110 °C; glycerol to DMC molar ratio –
1:3; catalyst loading – 5.5 wt%; reaction time – 7h. Fig. 13. Reusability study of Ti-SBA-15 (A) catalyst.

used up to three cycles and the results are presented in Fig. 13. After reused catalyst which supports the assumption that side products pro-
each reaction, the catalyst was separated by filtration, washed with duced during the reaction might be adsorbed on the active catalytic
methanol and dried at 110 °C for using in next cycle. The results in- sites resulting in decrease in activity and selectivity of the catalyst.
dicate that the activity of the catalyst was reduced only by 10% after As BET and FTIR results suggest pore blocking in reused catalysts is
three cycles, however GYC selectivity was significantly affected after either due to coke formation or amorphous silica species. To find out
3rd cycle. To find out the reason behind the gradual loss of GYC se- the exact reason, the catalyst (after 3rd cycle) was regenerated by
lectivity, the catalyst sample obtained after third cycle was tested by calcination (500 °C for 4 h) and reused for further experiments. The
ATR-FTIR and BET surface area analyzer. The N2 physiosorption data selectivity of the calcined catalyst was found much improved compared
indicated that the surface area of Ti-SBA-15 (5) catalyst decreased from to the previous reaction (non-calcined, 3rd cycle) and glycerol con-
845 to 106 m2/g for the regenerated catalyst. In addition, the micropore version of 91% and GYC selectivity of 73% were achieved. It was found
volume of the catalyst was also dropped from 0.016 cm3/g (fresh cat- that glycerol conversion and GYC selectivity of the spent catalyst was
alyst) to 0.009 cm3/g (regenerated catalyst). These results suggest that reduced only ∼3% and 9%, respectively after 4th cycle compared to
after regeneration cycles, the catalytic active sites are probably closed the 1st cycle. After 4th cycle, the structural identities of the products
by condensed silica frameworks and not fully accessible to the reactant were confirmed by GC–MS and 1H NMR spectroscopy and the reaction
molecules [27]. Fig. S2 shows the FTIR spectra for the fresh Ti-SBA-15 products were identified as GYC, methanol and 1-(o-methoxy-carbonyl)
(A) and recycled catalyst (recycle-1, recycle-2). The change in intensity glycerol. Apart from formation of 1-(o-methoxy-carbonyl) glycerol and
of peak at 3800 cm−1, 1550 cm−1 and 545 cm−1 was observed for the methanol, no other by-products were observed experimentally. XRD

486
P. Devi et al. Chemical Engineering Journal 346 (2018) 477–488

Table 2
Comparison of Ti-SBA-15 (A) catalyst against reported solid catalysts.
Catalyst Reaction temperature (°C) Reaction time (h) Glycerol conversion (%) GYC selectivity (%) GYC yield (%) References

Ni doped hydrotalcite 100 2 100 55 55 [28]


Mg–La oxide 85 1 81.3 90.0 75.71 [29]
SW21 140 4 52.1 95.3 49.7 [11]
Au/Fe2O3 150 4 80 48 39 [30]
Au/Nb2O5 140 4 49.5 85.4 42.3 [30]
Zn/TPA 140 4 69.2 99.4 68.7 [31]
Zr-P 140 3 80 100 76 [13]
CaO 65 0.08 93.4 n.a.* 85 [32]
CH3OK 70 4 n.a. n.a. 96 [10]
Co3O4/ZnO 145 4 n.a. 100 35 [33]
Mg/Al/Zr catalysts 75 1.5 94 n.a. n.a. [34]
La2O3 140 1 67.6 98.1 68.9 [35]
Cu/La2O3 150 12 15.2 45.4 33.4 [36]
CeO2 150 5 78.9 n.a. n.a. [12]
Ti-SBA-15 (A) 87.5 4 94 87.17 82 Present study

* n.a. – not available.

analysis of the reused catalyst (before and after calcination) was per- Ti was the main driving force in facilitating the reaction. A kinetic
formed and the results are given in the Fig. S3. Fig. S3a,b shows that model was developed which suggests that temperature was a critical
there is no change in XRD plot before and after reuse, which indicates parameter for the conversion of glycerol to GYC. Ti-SBA-15 displayed
that decrease in selectivity of catalysts was not due to the condensed better performance compared to various solid catalysts reported in the
silica frameworks. Based on these results it can be concluded that the literature. Overall the process has the great economic potential which
decrease in catalyst selectivity was due to the temporary pore blockage warrants further research for the commercialization of the process.
and activity of catalyst can be restored by calcination. It shows that this
catalyst is promising for practical applications. Acknowledgement

Authors are thankful for the financial support provided by the


3.3. Comparison of Ti-SBA-15 (A) against reported solid catalysts
Agriculture Development Fund (ADF), Canada.
Various solid catalysts have been reported in the literature for the
Appendix A. Supplementary data
conversion of glycerol to GYC using DMC as the source of carbonylation
(Table 2). The catalytic activity of Ti-SBA-15 (A) was compared against
Supplementary data associated with this article can be found, in the
some of these solid catalysts in relation to glycerol conversion, GYC
online version, at http://dx.doi.org/10.1016/j.cej.2018.04.030.
yield and selectivity. Some of these catalysts showed very high catalytic
conversion but low selectivity and some showed high selectivity but
References
low glycerol conversion. The Ti-SBA-15 (A) catalyst demonstrated high
glycerol conversion of 94%, GYC selectivity of 87% and yield 82%. The
[1] M.G. Álvarez, M. Plíšková, A.M. Segarra, F. Medina, F. Figueras, Synthesis of gly-
industrial application of this catalyst is warranted in view of the fol- cerol carbonates by transesterification of glycerol in a continuous system using
lowing reasons: (i) the performance of Ti-SBA-15 (A) is comparable to supported hydrotalcites as catalysts, Appl. Catal., B 113–114 (2012) 212–220.
the reported catalysts and commercial catalysts as discussed in Table 2, [2] J.R. Ochoa-Gómez, O. Gómez-Jiménez-Aberasturi, B. Maestro-Madurga,
A. Pesquera-Rodríguez, C. Ramírez-López, L. Lorenzo-Ibarreta, et al., Synthesis of
and (ii) Ti-SBA-15 is non-toxic having eco-friendly silica framework and glycerol carbonate from glycerol and dimethyl carbonate by transesterification:
easily recoverable from the reaction mixture. catalyst screening and reaction optimization, Appl. Catal., A 366 (2009) 315–324.
[3] S. Bagheri, N.M. Julkapli, W.A. Yehye, Catalytic conversion of biodiesel derived raw
glycerol to value added products, Renewable Sustainable Energy Rev. 41 (2014)
4. Conclusions 113–127.
[4] R. Bai, S. Wang, F. Mei, T. Li, G. Li, Synthesis of glycerol carbonate from glycerol
and dimethyl carbonate catalyzed by KF modified hydroxyapatite, J. Ind. Eng.
Ti-SBA-15 is disclosed as a highly active green catalyst for the Chem. 17 (2011) 777–781.
production of GYC from glycerol in high yield (82%) and selectivity [5] J.S. Chang, D.H. Chen, Optimization on the etherification of glycerol with tert-butyl
alcohol, J. Taiwan Inst. Chem. Eng. 42 (2011) 760–767.
(87%). A sol gel method was employed for in situ incorporation of Ti
[6] Z. Gholami, A.Z. Abdullah, K.T. Lee, Dealing with the surplus of glycerol production
into the silica framework of SBA-15. Ti-SBA-15 catalysts with varying from biodiesel industry through catalytic upgrading to polyglycerols and other
Si/Ti ratio were synthesized and the influence of Ti on textural prop- value-added products, Renewable Sustainable Energy Rev. 39 (2014) 327–341.
erties and surface morphology of Ti-SBA-15 catalysts and its catalytic [7] M.G. Álvarez, A.M. Frey, J.H. Bitter, A.M. Segarra, K.P. de Jong, F. Medina, On the
role of the activation procedure of supported hydrotalcites for base catalyzed re-
activity were investigated. The activity of Ti-SBA-15 catalysts was actions: glycerol to glycerol carbonate and self-condensation of acetone, Appl.
found to increase with an increase in Ti loadings (lower Si/Ti ratio) and Catal., B 134–135 (2013) 231–237.
the catalyst with higher Ti loading (Ti-SBA-15 (A)) showed greater [8] H. Li, X. Jiao, L. Li, N. Zhao, F. Xiao, W. Wei, et al., Synthesis of glycerol carbonate
by direct carbonylation of glycerol with CO2 over solid catalysts derived from Zn/
catalytic activity as compared to Ti-SBA-15 catalysts with higher Si/Ti Al/La and Zn/Al/La/M (M = Li, Mg and Zr) hydrotalcites, Catal. Sci. Technol. 5
ratio. Ti-SBA-15 (A) demonstrated more than 10-times greater activity (2015) 989–1005.
than SBA-15 which indicates the important role of Ti in exhibiting [9] F.S.H. Simanjuntak, J.S. Choi, G. Lee, H.J. Lee, S.D. Lee, M. Cheong, et al., Synthesis
of glycerol carbonate from the transesterification of dimethyl carbonate with gly-
higher catalytic properties. Investigation of surface and chemical cerol using DABCO and DABCO-anchored Merrifield resin, Appl. Catal., B 165
properties of catalysts suggests that higher Ti loadings led to the for- (2015) 642–650.
mation of an extra layer of anatase TiO2 crystals on the catalyst surface [10] J. Esteban, E. Domínguez, M. Ladero, F. Garcia-Ochoa, Kinetics of the production of
glycerol carbonate by transesterification of glycerol with dimethyl and ethylene
which accounts for their enhanced catalyst activity. A reaction me-
carbonate using potassium methoxide, a highly active catalyst, Fuel Process.
chanism based on the products identified from the reaction mixture Technol. 138 (2015) 243–251.
suggests that the reaction proceeded via the formation of an O- [11] M. Aresta, A. Dibenedetto, F. Nocito, C. Ferragina, Valorization of bio-glycerol: New
methoxy-carbonyl glycerol intermediate, and the Lewis acidic nature of catalytic materials for the synthesis of glycerol carbonate via glycerolysis of urea, J.

487
P. Devi et al. Chemical Engineering Journal 346 (2018) 477–488

Catal. 268 (2009) 106–114. acid/base and structural properties of titanate-loaded mesoporous silica by grafting
[12] J. Liu, Y. Li, J. Zhang, D. He, Glycerol carbonylation with CO2 to glycerol carbonate of zinc oxide, J. Phys. Chem. C. 116 (2012) 14318–14327.
over CeO2 catalyst and the influence of CeO2 preparation methods and reaction [25] A.I. Acatrinei, M.A. Hartl, J. Eckert, E.H.L. Falcao, G. Chertkov, L.L. Daemen,
parameters, Appl. Catal., A 513 (2016) 9–18. Hydrogen adsorption in the Ti-doped mesoporous silicate SBA-15, J. Phys. Chem. C.
[13] M. Aresta, A. Dibenedetto, F. Nocito, C. Pastore, A study on the carboxylation of 113 (2009) 15634–15638.
glycerol to glycerol carbonate with carbon dioxide: the role of the catalyst, solvent [26] R. Srivastava, D. Srinivas, P. Ratnasamy, Sites for CO2 activation over amine-
and reaction conditions, Atmos. Environ. 41 (2007) 407–416. functionalized mesoporous Ti(Al)-SBA-15 catalysts, Microporous Mesoporous
[14] F.S.H. Simanjuntak, T.K. Kim, S.D. Lee, B.S. Ahn, H.S. Kim, H. Lee, CaO-catalyzed Mater. 90 (2006) 314–326.
synthesis of glycerol carbonate from glycerol and dimethyl carbonate: isolation and [27] S.Y. Chen, T. Mochizuki, Y. Abe, M. Toba, Y. Yoshimura, Ti-incorporated SBA-15
characterization of an active Ca species, Appl. Catal., A 401 (2011) 220–225. mesoporous silica as an efficient and robust Lewis solid acid catalyst for the pro-
[15] K. Jagadeeswaraiah, C.R. Kumar, P.S.S. Prasad, S. Loridant, N. Lingaiah, Synthesis duction of high-quality biodiesel fuels, Appl. Catal., B 148–149 (2014) 344–356.
of glycerol carbonate from glycerol and urea over tin-tungsten mixed oxide cata- [28] P. Liu, M. Derchi, E.J.M. Hensen, Promotional effect of transition metal doping on
lysts, Appl. Catal., A 469 (2014) 165–172. the basicity and activity of calcined hydrotalcite catalysts for glycerol carbonate
[16] F. Bérubé, F. Kleitz, S. Kaliaguine, A comprehensive study of titanium-substituted synthesis, Appl. Catal., B 144 (2014) 135–143.
SBA-15 mesoporous materials prepared by direct synthesis, J. Phys. Chem. C. 112 [29] F.S.H. Simanjuntak, V.T. Widyaya, C.S. Kim, B.S. Ahn, Y.J. Kim, H. Lee, Synthesis of
(2008) 14403–14411. glycerol carbonate from glycerol and dimethyl carbonate using magnesium-lan-
[17] R.V. Sharma, C. Baroi, A.K. Dalai, Production of biodiesel from unrefined canola oil thanum mixed oxide catalyst, Chem. Eng. Sci. 94 (2013) 265–270.
using mesoporous sulfated Ti-SBA-15 catalyst, Catal. Today 237 (2014) 3–12. [30] C. Vieville, J.W. Yoo, S. Pelet, Z. Mouloungui, Synthesis of glycerol carbonate by
[18] J.Y. Kim, J.Y. Kim, H.J. Kang, W.Y. Kim, Y.H. Lee, J.S. Lee, Isomorphic Ti sub- direct carbonatation of glycerol in supercritical CO2 in the presence of zeolites and
stitution into SBA-15 without Ti loss and with lower TiO2 segregation, Inorg. Chem. ion exchange resins, Catal. Lett. 56 (1998) 245–247.
53 (2014) 5884–5886. [31] C.R. Kumar, K. Jagadeeswaraiah, P.S.S. Prasad, N. Lingaiah, Samarium-exchanged
[19] F. Bérubé, A. Khadhraoui, M.T. Janicke, F. Kleitz, S. Kaliaguine, Optimizing silica heteropoly tungstate: An efficient solid acid catalyst for the synthesis of glycerol
synthesis for the preparation of mesoporous Ti-SBA-15 epoxidation catalysts, Ind. carbonate from glycerol and benzylation of anisole, ChemCatChem 4 (2012)
Eng. Chem. Res. 49 (2010) 6977–6985. 1360–1367.
[20] M.M. Araújo, L.K.R. Silva, J.C. Sczancoski, M.O. Orlandi, E. Longo, A.G.D. Santos, [32] W.K. Teng, G.C. Ngoh, R. Yusoff, M.K. Aroua, Microwave-assisted transesterifica-
et al., Anatase TiO2 nanocrystals anchored at inside of SBA-15 mesopores and their tion of industrial grade crude glycerol for the production of glycerol carbonate,
optical behavior, Appl. Surf. Sci. 389 (2016) 1137–1147. Chem. Eng. J. 284 (2016) 469–477.
[21] W. Zhan, J. Yao, Z. Xiao, Y. Guo, Y. Wang, Y. Guo, et al., Catalytic performance of [33] V. Calvino-Casilda, G. Mul, J.F. Fernandez, F. Rubio-Marcos, M.A. Banares,
Ti-SBA-15 prepared by chemical vapor deposition for propylene epoxidation: The Monitoring the catalytic synthesis of glycerol carbonate by real-time attenuated
effects of SBA-15 support and silylation, Microporous Mesoporous Mater. 183 total reflection FTIR spectroscopy, Appl. Catal., A 409–410 (2011) 106–112.
(2014) 150–155. [34] P. Liu, M. Derchi, E.J.M. Hensen, Synthesis of glycerol carbonate by transester-
[22] X.F. Qian, T. Kamegawa, K. Mori, H.X. Li, H. Yamashita, Calcium phosphate coat- ification of glycerol with dimethyl carbonate over MgAl mixed oxide catalysts,
ings incorporated in mesoporous TiO2/SBA- 15 by a facile inner-pore sol−gel Appl. Catal., A 467 (2013) 124–131.
process toward enhanced adsorption-photocatalysis performances, J. Phys. Chem. C [35] L. Wang, Y. Ma, Y. Wang, S. Liu, Y. Deng, Efficient synthesis of glycerol carbonate
117 (2013) 19544–19551. from glycerol and urea with lanthanum oxide as a solid base catalyst, Catal.
[23] W.H. Zhang, J.Q. Lu, B. Han, M.J. Li, J.H. Xiu, P.L. Ying, et al., Direct synthesis and Commun. 12 (2011) 1459–1462.
characterization of titanium-substituted mesoporous molecular sieve SBA-15, [36] J. Zhang, D. He, Surface properties of Cu/La2O3 and its catalytic performance in the
Chem. Mater. 14 (2002) 3413–3421. synthesis of glycerol carbonate and monoacetin from glycerol and carbon dioxide,
[24] B. Mei, A. Becerikli, A. Pougin, D. Heeskens, I. Sinev, W. Grünert, et al., Tuning the J. Colloid Interface Sci. 419 (2014) 31–38.

488

You might also like