You are on page 1of 13

Reaction Kinetics, Mechanisms and Catalysis (2019) 126:167–179

https://doi.org/10.1007/s11144-018-1447-4(0123456789().,-volV)(0123456789().,-volV)

Comparison of sulfonic acid loaded mesoporous silica


in transesterification of triacetin

Mahuya Bandyopadhyay1,2 • Nao Tsunoji2 • Rajib Bandyopadhyay3 •

Tsuneji Sano2

Received: 24 May 2018 / Accepted: 8 July 2018 / Published online: 13 July 2018
Ó Akadémiai Kiadó, Budapest, Hungary 2018

Abstract
Covalently linked sulfonic acid (–SO3H)-modified ordered mesoporous silicas
MCM-48, MCM-41, and SBA-15 were synthesized, characterized and their cat-
alytic activities were evaluated in the transesterification reaction of triacetin with
methanol. Acid modified materials were prepared by oxidative transformation of
immobilized functionalized unit, 3-mercaptopropyltriethoxysilane (MPTES) as a
precursor. The mesophase and porosity of the catalysts were determined by means
of X-ray diffraction and N2 adsorption techniques. No degradation of structure was
observed in the preparation process. The acid concentrations were calculated using
TG–DTA and NH3–TPD analysis. The acid modified materials were found to be
active catalysts for the transesterification of triacetin with methanol. Especially,
three-dimensional-MCM-48-SO3H showed better catalytic activity compared to its
two-dimensional counterparts MCM-41 and SBA-15.

Keywords Mesoporous silica  MCM-48  Sulfonic acid  Transesterification

Electronic supplementary material The online version of this article (https://doi.org/10.1007/s11144-


018-1447-4) contains supplementary material, which is available to authorized users.

& Mahuya Bandyopadhyay


mahuyabandyopadhyay@iitram.ac.in
1
Institute of Infrastructure, Technology, Research and Management, IITRAM, Maninagar,
Ahmedabad, Gujarat, India
2
Department of Applied Chemistry, Graduate School of Engineering, Hiroshima University,
Higashi-Hirosima 739-8527, Japan
3
School of Technology, Pandit Deendayal Petroleum University, Raisan, Gandhinagar, Gujarat,
India

123
168 Reaction Kinetics, Mechanisms and Catalysis (2019) 126:167–179

Introduction

Mesoporous silicas with high surface area and flexible pore size have been studied
extensively. Surface modifications like incorporation of organic functional groups
either by co-condensation process during synthesis or by post-synthesis grafting
[1–4] lead to improvement of physicochemical properties of these type of materials.
The different organic functional groups such as amines, amino acids, phosphoric
acid, carboxylic acids, and isocyanate etc. have been incorporated successfully on
the internal surface of SBA-15 for different applications [5–10]. Such work has been
explored further on solid acid catalysts that consist of various inorganic/organic
materials, including ion-exchange resins having sulfonic acid groups [11], metal-
containing molecular sieves [12], sulfated or mixed oxides [13], heteropoly acids
[14], metal–organic frameworks-based solid acids [15, 16], acid functionalized
silica/mesoporous silica [17, 18], zeolites [19] and so on. One of the more broadly
studied systems involves incorporation of tethered sulfonic acid groups, usually via
a propyl linkage to a silane which is condensed with surface silanol groups. For
ordered mesoporous solids, they offer an alternative, generally better, way of
emerging surface acidity to those methods that depend on isomorphous lattice
substitution (as in zeolites) to generate intrinsic acidity [20, 21], and mesoporous
materials functionalized with sulfonic acid have also shown to be active acid
catalysts in various reactions [22–24].
Organic–inorganic hybrid mesoporous silicas functionalized with sulfonic acid
groups have shown successful results for acid-catalyzed reactions [25–27]. These
materials have been derived with strong acid species by covalent attachments of
alkyl sulfonic acid groups to the silica cavities.
The transesterification reaction to form alkyl ester with short chain alcohols has
attracted immense attention in recent years for biodiesel fuels and value added
products [28–32]. Some triglycerides can be explored as a good source of bio fuel as
they have similar composition as fossil fuels [33]. A careful choice of catalyst may
improve the efficiency of these processes, and use of heterogeneous catalysts have
significant advantages over conventional homogeneous catalysts in this regard
[34–36]. There are some research reports on incorporation of alkyl sulfonic acid
groups onto the mesoporous silica framework generating acidic solid catalysts that
can catalyze esterification reactions [3, 37, 38]. There is one report on esterification
of palmitic acid using acid functionalized mesoporous silica [39]. They used
mesoporous silica synthesized using amine surfactant (denoted as HMS) and SBA-
15 as silica source and the catalytic performance of the mesoporous materials were
compared with commercially available acid catalysts. A number of reviews have
been published in recent years discussing the functionalization of mesoporous
materials and their applications in organic transformation [40–42]. We also recently
published one paper on transesterification of triacetin using amine modified
mesoporous materials, and the amine modified MCM-48 showed excellent catalytic
activity in this reaction [43].
In continuation with our previous work, where we prepared a series of amine
modified mesoporous materials and evaluated the catalytic activities in base

123
Reaction Kinetics, Mechanisms and Catalysis (2019) 126:167–179 169

catalyzed transesterification reaction, here we have extended our research on


modifying MCM-48, MCM-41 and SBA-15 with sulfonic acid groups and explored
their catalytic activities as solid acid catalyst on transesterification of triacetin to see
the difference between base modified and acid modified catalysts.
The key feature of this study was a comparison of acid functionalized MCM-48
with MCM-41 and SBA-15 in transesterification reaction by exploring reaction
parameters and conditions. Additionally, heterogeneity tests on functionalized
MCM-48 as well as regeneration and reusability of the catalyst up to fourth cycle
were also the highlights of the study.

Experimental

Synthesis of materials

MCM-48, MCM-41 and SBA-15 were synthesized following our previous report
[43], in addition to other reported processes by various research groups [44–46].

Modification of mesoporous silica by functionalization

Modifications of the materials obtained after synthesis were carried out by


functionalization with sulfonic acid [47]. It was performed by vacuum-drying the
samples at 70 °C for 4 h followed by refluxing the catalyst (1.5 g/0.025 mol) with a
mixture of 50 mL toluene and excess amount of 3-mercaptopropyltrimethoxysilane
(3.6 g/0.018 mol) at 110 °C for 12 h. The resulting solid (after filtering, washing
with toluene and air-drying) was stirred at room temperature for 24 h in excess of
30% H2O2 solution. The samples were filtered, washed with water and dispersed in
excess of 0.2 M H2SO4 solution and stirred overnight. Finally the solid function-
alized samples were recovered by filtration, washing with water and drying at
60 °C.

Catalyst characterization

For the determination of phase purity, crystallinity, Bruker AXS D8 Advance


diffractometer with graphite Cu Ka radiation working at 40 kV and 30 mA was
used. 29Si magic-angle spinning (MAS) NMR spectra were measured carried out at
Varian 600PS solid NMR spectrometer at 119.18 MHz. The catalysts were spun
around the magic angle 4 kHz at 6 mm diameter zirconia. 6.2 ls pulses, a 100 s
recycle delay, and 1000 scans were applied for the samples using 3-(trimethylsilyl)
propionic-2,2,3,3-d4 acid sodium salt as a chemical shift reference. Spinning
frequency of 4 kHz, a 90° pulse length of 5.6 ls, and a cycle delay time of 5 s were
applied for measuring of 1H-13C cross-polarized (CP) MAS NMR spectra. Here
Hexamethylbenzene was used as reference material. Scanning electron microscopy
(SEM) was performed using a Hitachi-S-4800 SEM attached with an energy-
dispersive X-ray (EDX) analyzer. SII 7300 TG–DTA (Seiko Instruments) instru-
ment was used to measure the thermal analysis. 3–4 mg of sample was heated at a

123
170 Reaction Kinetics, Mechanisms and Catalysis (2019) 126:167–179

heating rate of 10 °C/min in a flow of air (50 mL/min) from room temperature up to
700 °C. Surface area and porosity measurements of the calcined and modified
samples by means of N2 physisorption were carried out at - 196 °C on BELSORP-
max, Bel Japan instruments. For the calcined sample degassing was done at 400 °C
and for modified sample at 200 °C. The distribution of the acidity of the samples
was measured in the temperature range 100–700 °C by measuring the temperature-
programmed desorption of ammonia (NH3-TPD, CAT-B-82 NH3-TPD, Bel Japan).
Helium was used as a carrier gas. The samples (ca. 0.1 g) were pre-heated at 200 °C
under a flow of helium (50 mL/min) and then ammonia (NH3/He = 1/99 vol%) was
adsorbed at 100 °C. After removing of the excess adsorbed ammonia by flushing
with helium at 100 °C for 30 min, the samples were heated at a rate of 10 °C/min.
Infra-red spectra were recorded at room temperature on an FT-IR spectrometer
(NICOLET 6700) at a resolution of 4 cm-1.

Catalytic activity measurements

0.025 g of sulfonic acid-loaded sample was subjected to activation at 70 °C for 1 h


followed by the trransesterification reaction. The liquid phase reaction was carried
out under stirring inside a screw-capped glass tube using triacetin and methanol with
the following reaction conditions: triacetin, (TCI [ 98%) 3 g, methanol (Kanto
Chem [ 99.8%) 6.52 g); MeOH:triacetin 16:1 mol/mol; reaction temperature and
time 70 °C and 18 h, respectively. Once the reaction was complete, the products
along with the catalysts were filtered and naphthalene (0.0166 g as internal
standard) was added to the filtrate followed by analysis with GC (Shimadzu: Model
14B, carrier gas: Argon, Column: Stabilwax).

Results and discussion

Synthesis and characterization of catalyst

Figiure S1 (Supplementary Material) depicts the XRD patterns, with excellent


periodicity, of all of the mesoporous silica samples under the present study in their
as-made, calcined as well as sulfonic acid modified forms. Condensation of silanol
groups result in the decrease in unit cell volume which is supported by shifting of
the 211 peak, in the respective XRD pattern of calcined MCM-48 sample, from
2h = 2.20° to 2h = 2.70°. Successful grafting of acid group in the silicious
mesoporous materials are clearly indicated by the decrease in peak intensities after
the modification. Maintaining the overall mesostructure, the peak intensities
decreased in some region of the XRD patterns of modified MCM-48 (between 2h
4.7 and 5.3°) and MCM-41 (between 2h 4–5°) due to the grafting of SO3H group
inside the pore channels and walls, their associated scattering contrast in between
[48].
The acid amount of organic functional groups incorporated in the materials was
calculated using NH3-TPD analysis. The amount of low, medium and strong acidity
were calculated and presented in Table 1. The concentration of acid sites was

123
Reaction Kinetics, Mechanisms and Catalysis (2019) 126:167–179 171

Table 1 NH3-TPD acidity of modified mesoporous materials


Catalyst NH3 acidity mmol/g at temperature range

100–350 °C 350–550 °C 550–700 °C


(low acidity) (medium acidity) (strong acidity)

MCM-48-SO3H 0.48 0.06 0.07


MCM-41-SO3H 0.16 0.07 0.09
SBA-15-SO3H 0.26 0.04 0.03

estimated by quantity of ammonia adsorbed. We calculated total acid concentration


by considering the NH3 peak above 350 °C because the peak with lower
temperature derived from physically adsorbed NH3 on the surface of mesoporous
silica. The NH3-TPD profile is depicted in Fig. 1. Additionally thermogravimetric
analysis was also performed to estimate the total acidity of the functionalized
material.
In Fig. S2, FT-IR spectra of the non-functionalized mesoporous materials are
shown along with organic acid modified samples. The broad peak at 3400 cm-1, the
peaks at 1070 and 795 cm-1 attributed to the silanolic O–H bond, and Si–O–Si
stretching and bending vibrations respectively [49]. Deformation vibration band of
–CH2 at 1460 cm-1 are also observed for all the materials, which is attributed from
the propyl group linked to silanolic OH group. S–OH stretching vibration band at
3445 cm-1 also confirms the existence of sulfonic acid group in the sample.
Actually S=O asymmetric vibration and symmetric vibration of –SO3H groups
comes around 1000–1200 cm-1 wave length range respectively, but due to overlap
with the Si–O–Si band at 1050–1150 cm-1 this peak cannot be resolved.

c
TCD signal (a.u)

100 200 300 400 500 600 700


Temperature ( C)

Fig. 1 NH3-TPD profile of (a) MCM-41-SO3H, (b) SBA-15-SO3H and (c) MCM-48-SO3H

123
172 Reaction Kinetics, Mechanisms and Catalysis (2019) 126:167–179

Furthermore, for the modified samples, peaks at the range of 2850–3000 cm-1 were
observed, corresponding to the methylene stretching vibration of the propyl chain
indicating the incorporation in the organic moiety which is not appeared in the pure
calcined unloaded sample.
In the N2 adsorption/desorption analysis, the sulfonic acid-modified materials are
compared with pure unloaded mesoporous silica materials and are given in Fig S3.
The N2 adsorption isotherms exhibit type IV followed by capillary condensation
taking place in the mesopores, which is typical for mesoporous materials over a
range between partial pressures 0.1 and 0.8. The decrease in surface area, pore
volume, and pore diameter is observed after modified with sulfonic acid groups. For
MCM-48, the surface area decreases from 1351 to 590 m2/g and pore diameter from
2.3 to 1.5 nm. Surface area and pore volume also decreased for MCM-41 but not
much for SBA-15 and this reduction is anticipated as the organic acid group is
grafted to the silanolic surfaces. In Table 2 the surface area and porosity data are
summarized.
In the 29Si MAS NMR spectra (Fig. 2), signals at -102 due to Si(OSi)3(OH) (Q3)
and -111 ppm corresponds to Si(OSi)3(OH) (Q3) and Si(OSi)4 (Q4), respectively is
clearly seen. In 13C CP MAS NMR spectra (Fig. 3), however, three distinct peaks at
d = 11.5, 19.84, and 54 ppm, corresponding to C atoms on the Si-CH2-CH2-
CH2-SO3H group are clearly displayed in the figure. These spectra match well with
the literature data [25] and further confirm that the mercaptopropyltrimethoxysilane
precursor incorporated in the organic moiety. A signal at 36.7 ppm in the 13C CP
MAS NMR of MCM-41-SO3H (Fig. 3a) is observed, can be explained by presence
of disulphide (R–S–S–R) species, which arises due to incomplete oxidation of thiol
groups [46]. However, there is no evidence of presence of disulphide species in the
other two catalysts.
From the SEM images, shown in Figs. 4, 5, 6, the particles sizes of MCM-48-
SO3H have been found as * 250–350 nm with spherical morphology, whereas
MCM-41-SO3H shows disc shaped morphology with a diameter of * 2.5 lm and
chain-like structure of SBA-15-SO3H particles of length * 1.0 lm.

Table 2 Characteristics of various unmodified and sulfonic acid modified mesoporous materials
Catalyst Surface areaa (m2/g) Pore volumeb (cc/g) Pore diameterb (nm)

MCM-48 1351 1.1 2.3


MCM-48-SO3H 590 0.44 1.5
MCM-41 1088 0.99 2.5
MCM-41-SO3H 850 0.62 1.9
SBA-15 852 0.96 6.4
SBA-15-SO3H 661 0.85 5.3
a
Determined by the BET method
b
Determined by the BJH method

123
Reaction Kinetics, Mechanisms and Catalysis (2019) 126:167–179 173

Fig. 2 29Si MAS NMR spectra


of (a) MCM-41, (b) MCM-41-
SO3H (c) MCM-48, (d) MCM-
48-SO3H (e) SBA-15, and
(f) SBA-15-SO3H
f

c
b

-60 -80 -100 -120 -140


Chemical shift (ppm)

Fig. 3 13C CP MAS NMR


spectra of (a) MCM-41-SO3H,
(b) MCM-48-SO3H, and
SO3H
(c) SBA-15-SO3H
c
2
3 1
Intensity (a.u)

2
3
1

3 1

75 25 -25
Chemical shift (ppm)

Catalytic activity test

Comparison of various catalysts

The catalytic performances of the sulfonic acid modified materials were evaluated
in the transesterification of triacetin with methanol. The optimum temperature for
the reaction was found at 70 °C. The reaction was continued for 18 h with methanol
and triacetin molar ratio of, 16 using 0.025 g of catalyst. In this transesterification
reaction triacetin conversion was calculated and main desired product, methyl
acetate was identified. From the Fig. 7, it is clear that sulfonic acid modified-MCM-

123
174 Reaction Kinetics, Mechanisms and Catalysis (2019) 126:167–179

Fig. 4 SEM images of MCM-


48-SO3H

1 µm

Fig. 5 SEM images of MCM-


41-SO3H

5 µm

Fig. 6 SEM images of SBA-15-


SO3H

2 µm

48 displays 94% triacetin conversion and a 45% methyl acetate yield, MCM-41-
SO3H shows 60% triacetin conversion and 24% yield, and SBA-15-SO3H shows
65% conversion and 25% yield in 18 h reaction time. On the other hand, related

123
Reaction Kinetics, Mechanisms and Catalysis (2019) 126:167–179 175

Fig. 7 Transesterification of
triacetin with methyl alcohol
over MCM-48-SO3H, MCM-41-
SO3H, and SBA-15-SO3H
catalysts. Black represents
conversion of triacetin and grey
represents yield of methyl
acetate. Reaction conditions:
catalyst 0.025 g, temperature
70 °C, and time 18 h

amine modified materials using same procedure in our previous report [43], MCM-
48-NH2 exhibited much higher activity (78% conversion, 52% yield, 10 mg
catalyst, 65 °C, 4.5 h) which is in agreement with the fact that acid catalysts usually
require higher temperature and longer reaction time than their basic counterpart in
transesterification reaction [50].

Effect of reaction time

The reaction kinetics was also investigated and is summarized in Fig. 8. After 4.5 h
MCM-48-SO3H showed 35% triacetin conversion where as MCM-41-SO3H and
SBA-15-SO3H exhibited 13–15% triacetin reaction conversion and finally the
triacetin conversion reached to 60–65% for MCM-41-SO3H and SBA-15-SO3H
after 18 h reaction time. Throughout the reaction process MCM-48-SO3H came out
to be a better catalyst for this reaction than other two materials. As we have already
found and concluded in our previous publication [43] that apart from acid
concentration the structure of support is also determining factor for the better

Fig. 8 Relationship of reaction 100


time with triacetin conversion
and methyl acetate yield over
Conversion of triacetin (%)

different catalysts (a) MCM-48- 80 a


SO3H, (b) SBA-15-SO3H,
b
(c) MCM-41-SO3H. Reaction
conditions: catalyst 0.025 g and 60
temperature 70 °C

40 c

20

0
0 5 10 15 20
Reaction time (h)

123
176 Reaction Kinetics, Mechanisms and Catalysis (2019) 126:167–179

performance of any material. MCM-48 with three-dimensional pore structure offers


superior active site accessibility than the one-dimensional pore structures of MCM-
41 resulting in better catalytic activity.
It is revealed from NH3-TPD as well as TG–DTA analysis that in SBA-15-SO3H,
acid concentration is much lower than MCM-48-SO3H (Table 1) which is reflected
in its catalytic activity. However its overall acidity is similar to that of MCM-41-
SO3H and they resemble in the catalytic activity under similar reaction condition.

Effect of catalyst amount

The effect of catalyst amount onto the reaction system was also amount of catalyst
was also varied to check the minimum amount of catalyst is required to have
maximum conversion. The catalyst amount was varied from 0.01 to 0.075 g
(Fig. 9). Conversion was increased with an increase in catalyst weight. MCM-48-
SO3H gives 12% triacetin conversion with 0.01 g catalyst after 4.5 h, whereas
MCM-41-SO3H and SBA-15-SO3H do not show any catalytic activity. After
increased the catalyst amount from 0.01 to 0.025 g, the catalysts showed activities
like MCM-48-SO3H, 12-15% of triacetin conversion. Even by using 0.075 g
catalyst MCM-41-SO3H and SBA-15-SO3H shows around 39–48% triacetin
conversion, which is almost half than that on MCM-48-SO3H (86%). Average
pore size of modified MCM-48 is around 1.5 nm. The dimension of triacetin is
1.02 9 0.38 nm [51]. From this fact, it is clear that internal diffusion of triacetin
inside the mesopore is not accountable.

Leaching test

A heterogeneity test was also done to check the leaching of the sulfonic acid groups
into the reaction mixture. After 3 h of reaction the catalyst was removed and the
reaction was continued for further 1.5 h without catalyst. Another set of reaction

Fig. 9 Effect of catalyst amount 100


on triacetin conversion over
different catalysts (a) MCM-48-
SO3H, (b) SBA-15-SO3H, and 80
Conversion of triacetin (%)

(c) MCM-41-SO3H. Reaction a


conditions: temperature 70 °C,
and time 4.5 h 60
b

40
c

20

0
0 0.02 0.04 0.06 0.08
Catalyst weight (g)

123
Reaction Kinetics, Mechanisms and Catalysis (2019) 126:167–179 177

was performed with same amount of catalyst for 4.5 h without any disturbance. In
both the cases the products were analyzed by GC. The conversion remains
unchanged after withdrawal of the catalyst and the reaction is run further for another
1.5 h. This result is indicative of the true heterogeneity of the modified catalyst.

Reusability of the catalysts

One of the most important efforts for heterogeneous catalysis is the test for
reproducibility or reusability of the catalysts to justify the robustness of the prepared
modified catalyst. Hence the recyclability of the catalysts was also evaluated and is
shown in Fig. 10. For every successive run the catalyst was washed thoroughly with
methanol and dried at room temperature. The spent MCM-48-SO3H catalyst was
also characterized by XRD and FTIR (Figs. S4, S5) which revealed the restoration
of the overall structure and the functional group after reaction. The catalysts were
found to be effective for up to four consecutive runs with moderate loss in catalytic
activities. This activity loss might be possible due to catalyst deactivation by coke
formation, as supported by the weight loss and corresponding acidity measurement
of used catalyst by TGA (Fig. S6). On the other hand, this loss in activity was
significant even after just first cycle using amine modified catalyst as discussed in
our previous work [43]. The usefulness and significance of catalyst recycling is a
matter of debate as rightly discussed in full length by Molner et al. [52, 53] and
more research on the reusability of the catalyst in present study can be explored
further in future.

100
Conversion of triacetin (%)

80

60

40

20

0
MCM-48-NH2 MCM-41-NH2 SBA-15-NH2
1st 2nd 3rd 4th

Fig. 10 Recyclability test for sulfonic acid loaded mesoporous materials. Reaction conditions:
temperature 70 °C, time 18 h, and catalyst amount 0.025 g

123
178 Reaction Kinetics, Mechanisms and Catalysis (2019) 126:167–179

Conclusions

The current work highlights the preparation of sulfonic acid-modified mesoporous


silica materials via post-synthesis grafting of organic acid group where reactive -SH
group further undergoes oxidation and acidification to convert –SO3H group. The
transesterification reaction can be catalyzed by both acid/bas catalysts. We have
already shown the excellent catalytic activity of amine loaded mesoporous materials
in this reaction and the present work is the continuation of the previous one to
investigate the acid properties of different mesoporous materials. The prepared
catalysts were systematically characterized by various physico-chemical techniques.
XRD, SEM, N2 adsorption, FT-IR, 29Si MAS NMR, 13C CP MAS NMR, NH3–TPD,
and TG–DTA measurements revealed the successful anchoring of acid group onto
the surface of mesoporous materials.
The observed results from various physico-chemical techniques confirmed that
the mesoporous structures of the materials were well preserved and no indication of
structural deterioration was observed even after structural modification.
The catalytic application was performed in transesterification of triacetin with
methanol. All the three catalysts show activity in this reaction, MCM-48-SO3H
acted as a most effective for this reaction and showed the highest activity with 94%
conversion, whereas MCM-41-SO3H and SBA-15-SO3H showed 60 and 65%
conversion, respectively at 70 °C after 18 h reaction time. A moderate reusability of
the functionalized catalysts was observed and the catalyst was used up to fourth
cycle. No leaching problem was also observed. These results suggest that
functionalized MCM-48, whether it is acidic or basic, is a potential catalyst for
this model reaction, and the work may be extended further and explored for
transesterification of non-edible oils in the the formation of biodiesel.

References
1. Zhao D, Feng J, Huo Q, Melosh N, Fredrickson GH, Chmelka BF, Stucky GD (1998) Science
279:548–552
2. Stein A, Melde BJ, Schrodein RC (2000) Adv Mater 12:1403–1419
3. Van Rhijn WM, De Vos DE, Bossaert WD, Jacobs PA (1998) Chem Commun 3:317–318
4. Ide Y, Iwata M, Yagenji Y, Tsunoji N, Sohmiya M, Komaguchi K, Sano T, Sugahara Y (2016) J
Mater Chem A 4:15829–15835
5. Ziarani GM, Badiei A, Mousavi S, Lashgari N, Shahbazi A (2012) Chin J Catal 33:1832–1839
6. Tsai CT, Pan YC, Ting CC, Vetrivel S, Chiang AST, Fey GTK, Kao HM (2009) Chem Commun
33:5018–5020
7. Mondal J, Nandi M, Modak A, Bhaumik A (2012) J Mol Catal A 254:363–364
8. Wu B, Tong Z, Yuan X (2012) J Porous Mater 19:641–647
9. Canilho N, Jacoby J, Pasc A, Carteret C, Dupire F, Stébé MJ, Blin JL (2013) Colloids Surf B
112:139–145
10. Shieh FK, Hsiao CT, Wu JW, Sue YC, Bao YL, Liu YH, Wan L, Hsu MH, Deka JR, Kao HM (2013)
J Hazard Mater 260:1083–1091
11. Harmer MA, Sun Q (2001) Appl Catal A 221:45–62

123
Reaction Kinetics, Mechanisms and Catalysis (2019) 126:167–179 179

12. Timofeeva MN, Panchenko VN, Hasan Z, Khan NA, Mel’gunov MS, Abel AA, Matrosova M,
Volchod KP, Jhung SH (2014) Appl Catal A 469:427–433
13. Khan NA, Mishra DK, Ahmed I, Yoon JW, Hwang JS, Jhung SH (2013) Appl Catal A 452:34–38
14. Timofeeva MN (2003) Appl Catal A 256:19–35
15. Goestena MG, Juan-Alcañiz J, Ramos-Fernandez EV, Gupta KBSS, Stavitski E, Bekkum HV,
Gascon J, Kapteijn F (2011) J Catal 281:177–187
16. Akiyama G, Matsuda R, Sato H, Takata M, Kitagawa S (2011) Adv Mater 23:3294–3297
17. Hasan Z, Jhung SH (2014) Eur J Inorg Chem 21:3420–3426
18. Kureshy RI, Ahmad I, Pathak K, Khan NH, Abdi SHR, Jasra RV (2009) Catal Commun 10:572–575
19. Zhang G, Zhang X, Lv J, Liu H, Qiu J, Yeung KL (2012) Catal Today 193:221–225
20. Meziani MJ, Zajac J, Jones DJ, Patyka S, Roziere J, Auroux A (2000) Langmuir 16:2262–2268
21. Brunel D, Blanc AC, Galarneau A, Fajula F (2002) Catal Today 73:139–152
22. Dias AS, Pillinger M, Valente AA (2005) J Catal 229:414–423
23. Malero JA, Stucky GD, Grieken R, Morales G (2002) J Mater Chem 12:1664–1670
24. Bandyopadhyay M, Shiju NR, Brown DR (2010) Catal Commun 11:660–664
25. Diaz I, Mohino F, Perez-Pariente J, Sastre E, Wright P, Zhou W (2001) Stud Surf Sci Catal
135:1248–1253
26. Margolese D, Melero JA, Christiansen SC, Chmelka BF, Stucky GD (2000) Chem Mater
12:2448–2459
27. Bossaert WD, De Vos DE, Van Rhijn WM, Bullen J, Grobet PJ, Jacobs PA (1999) J Catal
182:156–164
28. Bender M (1999) Bioresour Technol 70:81–87
29. Diasakou M, Louloudi A, Papayannakos N (1998) Fuel 77:1297–1302
30. Ogoshi T, Miyawaki Y (1985) J Am Oil Chem Soc 62:331–335
31. Suppes GJ, Bockwinkel K, Lucas S, Botts JB, Mason MH, Heppert JA (2001) J Am Oil Chem Soc
78:139–145
32. Kildiran G, Yucel SO, Turkay S (1996) J Am Oil Chem Soc 73:225–232
33. Nam LTH, Vinh TQ, Loan NTT, Van Tho DS, Yang X, Su B (2011) Fuel 90:1069–1075
34. Hara M (2009) Chem Sus Chem 2:109–135
35. Sharma YC, Singh B (2010) Biofuels, Bioprod Biorefin 5:69–92
36. Serio MD, Tesser R, Pengmei L, Santacesaria E (2008) Energy Fuels 22:207–217
37. Diaz I, Mohino F, Perez-Pariente J, Sastre E (2003) Appl Catal A 242:161–169
38. Alvaro M, Corma A, Das D, Fornes V, Garcia H (2005) J Catal 231:48–55
39. Mbaraka IK, Radu DR (2003) Y Lin VS, Shanks BH. J Catal 219:329–336
40. Sayari A, Hamoudi S (2001) Chem Mater 13:3151
41. Lee AF, Bennett JA, Manayil JC, Wilson K (2014) Chem Soc Rev 43:7887–7916
42. Shagufta Ahmad I, Dhar R (2017) Catal. Surv. Asia 21:53–69
43. Bandyopadhyay M, Tsunoji N, Sano T (2017) Catal Lett 147:1040–1050
44. Gies H, Grabowski S, Bandyopadhyay M, Grunert W, Tkachenko OP, Klementiev KV, Birkner A
(2003) Micropor Mesopor Mater 60:31–42
45. Lesaint C, Lebeau B, Marichal C, Patarin J (2005) Micropor Mesopor Mater 83:76–84
46. Wang X, Tseng YH, Chan JCC (2007) J Phy Chem C 111:2156–2164
47. Siril PF, Davison AD, Randhawa JK, Brown DR (2007) J Mol Catal A 267:72–78
48. Yoshitake H, Yokoi T, Tatsumi T (2002) Chem Mater 14:4603–4610
49. Li Y, Zhou G, Li C, Qin D, Qiao W, Chu B (2009) Colloids Surf A 341:79–85
50. Fredman B, Pryde EH, Mounts TL (1984) J Am Oil Chem Soc 61:1638–1643
51. Silveira JQ, Vargas MD, Ronconi CM (2011) J Mater Chem 21:6034–6039
52. Molnar A (2011) Chem Rev 111:2251–2320
53. Molnar A, Papp A (2017) Coord Chem Rev 349:1–65

123

You might also like