Vous êtes sur la page 1sur 8

Microporous and Mesoporous Materials 164 (2012) 182189

Contents lists available at SciVerse ScienceDirect

Microporous and Mesoporous Materials


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

Epoxidation of biodiesel with hydrogen peroxide over Ti-containing


silicate catalysts
Nicole Wilde, Christian Worch, Wladimir Suprun, Roger Glser
Institute of Chemical Technology, Universitt Leipzig, Linnstr. 3, 04103 Leipzig, Germany

a r t i c l e

i n f o

Article history:
Available online 6 July 2012
Dedicated to Prof. Dr. Ing. Jens Weitkamp on
the occasion of his 70th birthday
Keywords:
FAME
Epoxidation
Hydrogen peroxide
TS-1
Microwave-assisted synthesis

a b s t r a c t
The heterogeneously catalyzed epoxidation of methyl oleate with hydrogen peroxide in the liquid phase
is reported using different catalysts such as TS-1, Ti-MCM-41, TiOxSiO2, and MOx and WOx supported on
Al2O3 or SiO2. At 323 K, in acetonitrile as the solvent and with an industrial TS-1 catalyst, an epoxide
selectivity of 87% at 93% conversion is achieved after 24 h. Variation of the catalyst mass, particles size
and the reaction temperature proove that the conversion is limited by mass transport to the outer crystal
surface of the catalyst. With TS-1 as the catalyst, also the unsaturated fatty acid methyl esters (FAME) in
biodiesel can be epoxidized with a conversion of 90% and selectivity of 76%. A higher TON than over TiMCM-41 or even the industrial TS-1 catalysts is reached, when TS-1 with nanoscaled particles and
stacked morphology is applied as the epoxidation catalyst. The latter was synthesized by microwaveassisted synthesis and possesses Ti sites in similar coordination geometry as conventional TS-1 as shown
by DR-UVVis spectroscopy.
2012 Elsevier Inc. All rights reserved.

1. Introduction
The current chemical industry has to respond to a steadily
increasing demand of synthetic products on the one hand and
the shortage of fossil and mineral raw materials as well as their
decreasing quality on the other hand [1,2]. Therefore, renewable
feedstocks have recently gained considerable interest as raw materials for chemical production. By far the largest share of utilized
renewables in the chemical industry is held by fats and oils [3].
Among the applications of fats and oils, their conversion to fatty
acid methyl esters (FAME) for biodiesel production is one of most
prominent. FAME can be further converted to epoxidized fatty acid
esters which play an important role for a broad range of large-scale
industrial synthesis of chemicals and intermediates such as plasticizers and stabilizers in PVC, intermediates in the production of
polyurethane polyols, components for lubricants, cosmetics or
pharmaceuticals [46].
Currently, epoxy fatty acid compounds are mainly obtained on
the industrial scale by the Prileshajew reaction, in which the unsaturated oils are converted with percarboxylic acids, such as peracetic or performic acid. This route suffers from several drawbacks: (I)
in the acidic reaction media, the selectivity for epoxides is relatively low due to oxirane ring opening, (II) the handling of peracids
and highly concentrated hydrogen peroxide solutions is strongly

Corresponding author.
E-mail address: roger.glaeser@uni-leipzig.de (R. Glser).
1387-1811/$ - see front matter 2012 Elsevier Inc. All rights reserved.
http://dx.doi.org/10.1016/j.micromeso.2012.06.047

hazardous, and (III) the aqueous solutions of C1C3 carboxylic


acids formed as by products are strongly corrosive [7,8].
In view of the principles of green chemistry, catalyzed epoxidations are the preferred alternatives to conventional stoichiometric
epoxidation reactions such as the Prileshajew reaction. Catalysts
based on titanium, tungsten and molybdenum were investigated
for epoxidations of unsaturated olens, with high yields and/or
selectivity of the target products [912]. Molybdenum complexes
or oxides on silica and alumina supports were extensively studied
in the epoxidation of cyclooctene and limonene as well as allyl
alcohol [1315]. For the epoxidation of fatty acids esters, the tungsten-based active compound tetrakis, i.e., ([(C8H17)3NCH3]3+
[PO4[W(O)(O2)2]4]3 ) was reported [16]. In the presence of H2O2,
acid reaction conditions result favoring ring opening of the epoxide
formed.
In several recent studies, catalytic epoxidations of FAME were
reported using organic hydroperoxides, such as tert-butyl- or cumene hydroperoxide, as epoxidizing agents [1719]. Typically,
TiO2SiO2 and Ti-MCM-41 were used as catalysts. However, hydrogen peroxide would be an economically and environmentally largely preferred oxidant. In fact, MCM-41 and MCM-48 with Ti
sites grafted onto the surface were found to be efcient catalysts
in the epoxidation of methyl oleate with aqueous hydrogen peroxide [20]. Conversions up to 96% and yields of the methyl epoxystearate up to 91% were obtained over Ti/MCM-41 after 24 h at 358 K
in liquid acetonitrile. Suarez et al. [21] and Sepulveda et al. [22] reported on the performance of different aluminas as catalysts for
the epoxidation of fatty acid methyl esters with anhydrous and

183

N. Wilde et al. / Microporous and Mesoporous Materials 164 (2012) 182189

aqueous hydrogen peroxide. Over a solgel derived alumina,


methyl oleate and soy bean oil methyl ester conversion of 95%
and epoxide selectivities >95% were achieved after 24 h at 353 K
in ethyl acetate as a solvent [22]. Titanosilicate molecular sieves,
especially titanium silicalite-1 (TS-1), were widely studied for the
selective epoxidation of a variety of organic substrates using aqueous H2O2 as the sole oxidant [2326]. Often high selectivities up to
94% at essentially complete conversion are achieved. The present
work was, thus, devoted to investigate the heterogeneously catalyzed epoxidation of unsaturated fatty acid methyl esters with
aqueous hydrogen peroxide as the sole oxidant over solid catalysts.
In particular, the catalytic activity, selectivity and stability of different supported transition metal oxide catalysts, e.g., TiOx, MoOx
or WOx on SiO2 or Al2O3, were studied and compared to that over
a commercial titanium silicalite-1 (TS-1) and Ti-MCM-41 with
methyl oleate as a model substrate. After a screening of the catalysts, the inuence of the solvent and the reaction temperature
was studied. It was also of interest whether the catalysts can be reused and if a deactivation takes place. An additional goal was to
clarify, whether the conversion of the rather bulky FAME molecules is limited by mass transport processes and, if yes, whether
strategies for an improved catalyst design can be derived. Finally,
it was investigated whether the knowledge obtained for methyl
oleate can be transferred to the conversion of FAME from real
biodiesel with aqueous hydrogen peroxide over Ti-containing silica
catalysts.

heating rate of 10 K min 1, i.e., 2 h at 473 K, 2 h at 673 K and


15 h at 813 K.
Titanium silicalite-1 with stacked morphology (TS-1_s) was
prepared according to [23] by microwave-assisted synthesis using
tetraethyl orthosilicate (TEOS, >99%, Merck), tetrapropylammonium hydroxide (TPAOH, 10 wt.% aqueous solution, Sigma Aldrich),
titanium(IV) isopropoxide (TIP, 97%, Sigma Aldrich), isopropanol
(99%, BDH Prolabo) and deionized water. Typically, TEOS (36.5 g),
TPAOH solution (35.7 g) and deionized water (39.2 g) were mixed
under stirring for 1 h. TIP (0.71 g) and Isopropanol (8.15 g) were
mixed under stirring in a separate beaker for 45 min. Subsequently
the solutions were added together and further stirred for 2 h at
room temperature and then stirred at 353 K for 1 h to remove isopropanol. The resulting mixture was partitioned to a 100 cm3 PTFE
autoclave and transferred to a microwave oven (MLS, Start 1500)
for crystallization. Therefore, the synthesis mixture were heated
for 5 min with the irradiation power of 600 W to 438 K and hold
at this temperature under autogeneous pressure for 20 min in case
of TS-1_s_20 and 60 min for TS-1_s_60. The obtained solids were
recovered by centrifugation, washed ve times with 25 cm3 deionized water, dried at 393 K for 6 h and calcined in air at 823 K for 6 h
as well.
An industrial sample of titanium silicalite-1 (TS-1, powder) was
supplied by Evonik.
Finally, all catalysts were pressed (6  108 N m 2), crushed and
sieved to obtain a fraction with a grain size between 50 and
150 lm for use in the catalytic experiments.

2. Experimental section

2.2. Catalyst characterization

2.1. Catalyst synthesis

The catalysts were characterized by powder X-ray diffraction


(XRD, Siemens D 5000) using CuKa radiation (k = 1.5418 ) in the
2h-range of 5 and 80 with a step size of 0.05. The specic surface
area ABET were determined from N2-sorption isotherms at 77 K
using an ASAP 2000 (micromeritics) apparatus. Table 1 reports
the values for ABET of the samples. The metal loading of the samples
was determined by elemental analysis via optical emission spectrometry with inductively coupled plasma (ICP-OES) after dissolving the solids in a mixture of HNO3, HF, and H3BO3 by chemical
extraction under pressure.
Diffuse reectance UVVis (DR-UVVis) spectra were measured
at room temperature on Perkin Elmer Lambda 650 S equipped with
a 150 mm integrating sphere using spectralon (PTFE, reective
value 99%) as a reference.

Supported MOxAl2O3 (M: W, Mo) as well as MoOxSiO2 catalysts were prepared by incipient wetness impregnation of
c-Al2O3 (bimodal, 99.7%, Alfa Aesar, specic surface area: 80
120 m2 g 1) or SiO2 (Aerosil 380, Degussa) with aqueous solutions
of (NH4)6H2W12O40H2O (85% WO3 basis, Aldrich) and
(NH4)2MoO4 (99%, Aldrich) under constant stirring at room temperature. Typically, 5 g of c-Al2O3 or SiO2 were used. The molar ratio of metal to aluminum or silicon, respectively, for impregnation
was kept constant at nM/nAl or Si = 1/10. After removal of water under UV irradiation for 180 min, the impregnated catalysts were further dried at 428 K for 16 h and calcined in air for 4 h at 823 K.
TiOxSiO2 was obtained according to [27] by grafting of titanium sites on SiO2 (Aerosil 380, Degussa) using a solution of titanium(IV) isopropoxide (TIP, 97%, Aldrich) in cyclohexanol (99%,
SigmaAldrich). Typically, 5 g SiO2 were treated with 150 cm3 of
the TIP solution (cTIP = 0.02 mol l 1) under reux for 2 h at 433 K.
After evaporation of the solvent in vacuum, the obtained solid
was calcined in air for 5 h at 823 K.
Ti-containing MCM-41 (Ti-MCM-41) was synthesized according
to [28] starting from silica gel beads (Licrospher Si 60, Merck) as
the silica source, Na2Ti3O7 as the titanium source (Aldrich) and
cetyltrimethylammonium hydroxide (CTMAOH) as the structuredirecting agent. CTMAOH was prepared by ion exchange of 1 g
CTMABr (99%, Acros) dissolved in 42 cm3 demineralized water
with 10 g Ampersep (900 OH, Fluka). The silica gel beads, a dened
amount of Na2Ti3O7 (Aldrich) and a 0.08 M CTMAOH aqueous solution were mixed under stirring for 10 min to obtain the synthesis
gel of the molar composition 0.012 TiO2 : SiO2 : 0.004 Na2O :
0.202 CTMAOH : 140.2 H2O. The resulting suspension was transferred to a polypropylene ask (Nalgene, 60 cm3 volume) and
held at 383 K for 1 day. Then, the resulting solid was removed by
ltration, washed ve times with 25 cm3 deionized water and once
with 20 cm3 anhydrous ethanol, and dried in air at 363 K for 15 h.
The obtained materials were calcined stepwise in air with a

2.3. Catalytic Experiments


Catalytic experiments were carried out batchwise in a twonecked round-bottom glass reactor (V = 25 cm3) with a septum
and a reux condenser in the liquid phase at 323 K with magnetic
stirring (400 rpm). Methyl oleate (MO, P99%, SigmaAldrich) or
biodiesel (FAME, methyl oleate: 72 wt.%, methyl linoleate + methyl
Table 1
Specic surface area ABET, metal content of the catalysts as well as turnover number
TON and turnover frequency TOF in the epoxidation of methyl oleate with hydrogen
peroxide after 24 h.
Sample

Specic
surface
area ABET
(m2 g 1)

Metal
content
(wt.%)

TON

TOF
(h 1)

TS-1 (ind.)
Ti-MCM-41
TiOxSiO2
WOx-Al2O3
MoOx-Al2O3
MoOx-SiO2

448
994
330
131
223
154

1.5
1.2
1.8
12.5
7.8
9.1

5.9
0.6
1.4
1.4
0.7
1.0

0.24
0.02
0.06
0.06
0.03
0.04

N. Wilde et al. / Microporous and Mesoporous Materials 164 (2012) 182189

(a)
conversion XMO / %

100
80
60
40
20
0

(b)
selectivity SME/ %

100
80
60
40
20
0

(c)
100
2

80

linolenate: 19 wt.%, rest: not determined ; JCN Neckermann-Biodiesel GmbH Halle) were used as substrates, hydrogen peroxide
(73 wt.% aqueous solution, Solvay-Wolfen) as oxidant, acetonitrile
(99,9%, BDH Prolabo) as a solvent and chlorobenzene (STD, 99.8%,
Aldrich) as internal standard. In a typical experiment, 10 cm3 solvent was loaded into the reactor, followed by the addition of the
substrate (MO or FAME, 90 mg, 0.30 mmol), hydrogen peroxide
(70 mg, 1.44 mmol) and chlorobenzene (67 mg). The loaded reactor was immersed in a heated oil bath and allowed to equilibrate
for 10 min to the reaction temperature of 353 1 K. The reaction
was started by addition of the catalyst (150 mg). A pre-drying of
the catalysts (in air at 373 K) did not have an inuence on the catalytic results as the catalysts are hydrophobic (vide infra). Thus, the
catalysts were used without pre-treatment. Samples (0.5 cm3)
were taken from the reaction mixture through to the septum via
syringe after 0, 1, 3, 5 and 24 h of reaction. The catalyst was removed from the samples by centrifugation. A dened amount of
the samples (0.1 cm3) was diluted in 0.5 cm3 acetonitrile and analyzed by capillary gas chromatography (Shimadzu GC 2010
equipped with a ame ionization detector) using nitrogen as the
carrier gas. Product separation was achieved on a capillary
column (95% dimethylpolysiloxane cross-linked with 5% diphenylpolysiloxane, Restek Rtx-5 MS, length 30 m, inner diameter
0.25 mm, coating thickness 0.25 lm). Reaction products were
identied by co-injection of authentic samples (see ESI) and by
GCmass spectrometry (GC Varian 3800).
The turnover number TON in MO conversion was calculated as
moles of methyl oleate converted per mole of metal sites present in
the catalysts. Accordingly, the turnover frequency TOF was obtained by dividing TON by the reaction time in hours. For the epoxidation of biodiesel (FAME), the conversion was calculated as the
ratio of the mass of converted unsaturated substrates, i.e., methyl
oleate, methyl linolate, and methyl linolenate, and their initial
mass in the reactant mixture. The epoxide selectivity in FAME conversions is given for the cumulative mass of all epoxides (mono-,
di- and tri-epoxides) formed relative to converted mass of the substrates. For calculation of the H2O2 conversion, the concentration of
H2O2 in the initial reactant solution and in the product samples
was determined by iodometric titration. The initial rate of the
MO conversion was calculated from tangential slopes of the time
dependence of the MO concentration at t = 0 using linear curve tting between 0 and 1 h of reaction time.
For reusability tests in three consecutive reactions, the catalyst
was ltered off, dried at room temperature and added to a fresh
reactant solution. For recycling experiments, the catalyst was removed from the reaction mixture by ltration after the third run,
calcined in air 673 K for 24 h and, again, added to fresh reactant
solution.

conversion XH O / %

184

60
40
20
0

3
3
iO 2
iO 2
d.)
lO
lO
-41
-S
in
-xS
-A 2
M
-A 2
Ox
O
1(
C
x
i
x
O
o
T
oO
-M
W
M
TS
M
Ti

Fig. 1. Conversion of methyl oleate XMO (a), epoxide selectivity SME (b) and
conversion of hydrogen peroxide XH2 O2 (c) in the epoxidation of methyl oleate over
different catalysts in acetonitrile at 323 K after 24 h.

3. Results and discussion


3.1. Conversion of methyl oleate
3.1.1. Catalyst screening
For a screening of catalysts for the conversion of methyl oleate
with aqueous hydrogen peroxide solution in the liquid phase, Ticontaining silicates as well as supported Mo- and W-oxides were
selected as catalysts. On the Al2O3-supported WOx and MoOx catalysts, crystalline phases were not observed by XRD (see ESI) indicating a high metal oxide dispersion on the support with overall
metal contents between 7.8 and 12.5 wt.% (Table 1). Among the
Mo- and W-based catalysts, the highest conversion of methyl oleate (57%) is achieved over WOxAl2O3 probably due to the high
metal loading (Fig. 1). The Mo-based catalysts are less active,

although somewhat more selective for epoxidation than the


W-based catalyst. The low activity of MoOxSiO2 could be a result
of a lower metal oxide dispersion. This is supported by the presence of reections of crystalline MoO3 in the orthorhombic phase
[29] in the XRD pattern (see ESI). Interestingly, however, H2O2 is
completely converted on all three W- or Mo-based catalysts. This
is consistent with earlier literature reports on oxidation reactions
with H2O2 over these type of catalysts [30].
Clearly the highest MO conversion and epoxide selectivity are
reached over the industrial TS-1 catalyst (Fig. 1). With >90% and
87%, respectively, these are much higher than those of the supported W- and Mo-oxides as well as TiOxSiO2. The higher activity
of the TS-1 (ind.) catalyst is also apparent when comparing
the TON and TOF (Table 1). Note, however, the similar TON for

185

N. Wilde et al. / Microporous and Mesoporous Materials 164 (2012) 182189

Fig. 2. Proposed reaction scheme for the epoxidation of methyl oleate with H2O2.

TiOxSiO2 and WOxAl2O3 despite the different metal loading and


the higher selectivity of Ti-based catalyst for the epoxide formation. Overall, the TON over the industrial TS-1 catalysts is larger
by a factor of 48 compared to the other catalysts studied.
In order to supply a high accessible surface, Ti-MCM-41
(ABET = 994 m2 g 1, Table 1, dP,DFT=3.4 nm, VP,BJH = 0.92 cm3 g 1)
was also included in the study. However, both MO conversion
and ME selectivity were below 10% on this catalyst. Either, the
rather bulky reactant molecules of MO do not have access to the
mesopores or the Ti sites are less active than those on TS-1 (ind.)
or TiOxSiO2. The lower TON for the Ti-MCM-41 with respect to
the latter two Ti-containing catalysts points towards a difference
in catalytically active Ti-sites rather than in the accessibility of
the pore system. This is in accordance with the results reported
by Guidotti et al. [20] for the epoxidation of methyl oleate with
hydrogenperoxide over Titanium-grafted MCM-41, MCM-48 and
SiO2 and Rios et al. [31] for the epoxidation of methyl oleate with
TBHP over Ti-MCM-41 and TiOxSiO2. The latter authors found that
conversion and selectivity are similar for the two catalyst and,
thus, largely independent of the geometry of the pore system. It
is also consistent with earlier reports describing the presence of
two active Ti sites, i.e., the tetrahedral (SiO)4Ti (species I) and the
tripodal (SiO)3Ti(OH) (species II), in Ti-containing silicates
[32,33]. While species I is predominant in TS-1, species II prevails
in amorphous TiSiO2 and Ti-MCM-41, respectively. It was found
for styrene epoxidation that species I exhibits superior selectivity
for the epoxide (as compared to species II) [32]. We may, therefore,
assume that a larger concentration of species I in TS-1(ind.) compared to TiOxSiO2 and Ti-MCM-41, also accounts for the observed
differences in conversion, TON, and selectivity for the epoxidation
of methyl oleate. The question of Ti coordination will be referred to
again in a later section.
Besides the major product methyl 9,10-epoxy stearate (ME),
several by-products were formed, predominantly in consecutive
reactions (Fig. 2). The hydrolysis product methyl 9,10-dihydroxy
stearate and the products from oxidative cleavage of the epoxide
were present in about equal amounts. The ketone from rearrangement is formed in small amounts only (<1%). The amount
of detected cleavage products increased with reaction time. However, the exact amount of the by-products was not further
quantied.

3.1.2. Inuence of the solvent


As opposed to organic hydroperoxides such as tert.-butyl or cumene hydroperoxide, the solubility of FAME in aqueous H2O2 solution is very limited. To obtain a homogeneous solution of the
FAME/H2O2-reactant mixture, the utilization of a liquid solvent is
required. The nature of the solvent is known to have a major inuence on reaction kinetics and product selectivity during the oxidative conversions over TS-1 as a catalyst [34,35].
For the epoxidation of methyl oleate with H2O2 over TS-1 (ind.),
the effect of polarity and the protic/aprotic nature of different solvents on initial reaction rate, conversion of MO and H2O2 and epoxide selectivity is summarized in Table 2. Generally, the initial
reaction rate (determined within the rst h of conversion) is well
reected in the conversion of MO after 24 h of reaction. The only
exceptions are the conversions in methanol and diisopropylether
where the rate was lowest and experimental error in rate determination becomes important.
The activity of TS-1 is lower in methanol than in acetonitrile.
This result cannot be explained only in terms of polarity, since both
solvents have similar dielectric constants (emethanol = 32.7,
eacetonitrile = 37.5). Likewise, it is in contrast with the reported positive effect of protic solvents on the reactivity of TS-1 [34,36]. This
might be an indication that the conversion does not occur inside
the micropores of the TS-1 catalysts, but at the external surface
where the solvent effect might be different.

Table 2
Inuence of the solvent on initial rate of reaction, conversion of methyl oleate XMO,
selectivity SME and conversion of hydrogen peroxide XH2 O2 in the epoxidation of
methyl oleate over TS-1 (ind.) at 323 K after 24 h.
Solvent

Initial Rate
r0
(mol l 1 h 1)

Conversion
XMO (%)

Selectivity
SME (%)

Conversion
XH2 O2 (%)

Ethylacetate
Acetone
Acetonitrile
Acetonitrile/
methanol
Diglyme
Methanol
Diisopropylether

0.025
0.024
0.021
0.016

98
92
93
79

36
52
87
85

72
36
97
62

0.015
0.002
0.005

79
20
4

84
17
31

66
57
67

186

N. Wilde et al. / Microporous and Mesoporous Materials 164 (2012) 182189

Scheme 1.

In general, MO conversion over 90% was achieved by using


aprotic-polar solvents like ethyl acetate rather than in aprotic nonpolar solvents like diisopropylether (Table 2). The MO conversion
in aprotic-polar solvents falls in the following order: ethylacetate > acetonitrile > acetone > acetonitrile/methanol > diglyme.
However, epoxide selectivity in ethyl acetate and acetone were below 50%. For ethyl acetate, conversion by acid catalyzed hydrolysis
leads to acetic acid and ethanol, the latter catalyzing epoxide ring
opening and formation of different diols. In the case of acetone, a
cyclic ketal (scheme 1) formed by the acid catalyzed reaction of
aceton with the diol from hydrolysis of the epoxide ring may explain the low selectivity.
The same effect, i.e., the formation of a cyclic ketal, was observed by Corma et al. [35] for epoxidation of 1-hexene with
H2O2 in the presence of Ti-BETA as the catalyst. Diol formation is
also the reason for reduced epoxide selectivity when methanol is
added to acetonitile as the solvent. Diglyme offers lipophilic properties and has the ability to dissolve high amounts of fatty acid
methyl esters [37], but, again, does not result in conversions of
MO or H2O2 above 80%. Since the clearly highest values for both
conversion of MO (93%) and epoxide selectivity (87%) were found
for acetonitrile, this solvent was chosen for the further studies.
Note that in acetonitrile, both MO and the aqueous H2O2 solution
are soluble at the concentrations applied here and that, thus, a
homogeneous liquid phase is in contact with the solid catalyst.

conversion X or selectivity S / %

100

XMO
SME
60
0
310

320

330

340

350

T/K
Fig. 3. Effect of reaction temperature on conversion of methyl oleate XMO and
epoxide selectivity SME in the epoxidation of methyl oleate over TS-1 (ind.) in
acetonitrile after 3 h.

but is lower at higher temperature. At these higher temperatures,


the formation of by-products, predominantly the cleavage products (Fig. 2), is favored. From the initial rate data for 313333 K,
an activation energy of 21.8 kJ mol 1 is calculated (see ESI). This
is considerably lower than that of typical epoxidations over TS-1.
For instance, the activation energy of ethylene epoxidation with
H2O2 was calclulated by DFT methods to be 98 kJ mol 1 [39]. The
low activation energy observed here points to a mass-transport
limitation of the MO conversion under the present reaction conditions. For further studies, the temperature was kept at 323 K as this
provides the most suitable compromise between selectivity and
conversion rate.

3.1.3. Inuence of the temperature


Generally, it is advantageous to carry out the epoxidation at
lower temperatures and, thus, limit the occurrence of side reactions such as hydrolysis, ethercation and deactivation of the catalyst [38]. Here, the reaction temperature was varied in the range
of 313 to 353 K. With increasing the temperature from 313 to
333 K, the conversion of MO increases almost linearly from 66%
to 91%, but levels off at higher temperatures (Fig. 3). The epoxide
selectivity is only slightly affected by temperature up to 333 K,

0.020

0.01

Thiele Modulus

6
0.015

0.010

r0

inital rate r0 / mol l-1 h-1

0.025

inital rate r0 / mol l-1 h-1

80

0.005

0.000
0

100

mcat. / mg

200

300

0
0.1

max. particle size / mm

Fig. 4. Initial rate r0 as a function of catalyst mass (left part) and as a function of maximum catalyst particle size (right part) for the epoxidation of methyl oleate over TS-1
(ind.) in acetonitrile at 323 K. Also the Thiele Modulus U is shown in the right part.

187

N. Wilde et al. / Microporous and Mesoporous Materials 164 (2012) 182189

conversion X or selectivity S / %

100

80

60

40

XMO

XH O

20

XMO

SME
0

10

15

SME
run 1

20

run 2

run 3

RC

t/h
Fig. 5. Conversion of methyl oleate XMO, epoxide selectivity SME and conversion of hydrogen peroxide XH2 O2 as a function of reaction time in the epoxidation of methyl oleate
over TS-1 (ind.) in acetonitrile at 323 K (left part) as well as after separation and reuse or regeneration (RC: regenerated catalyst, calcination in air ow for 24 h at 643 K) of the
catalyst (right part, reaction time 24 h).

-5

-5

-10

-10

-15

-15

1.4x10

-7

1.2x10

-7

1.0x10

-7

8.0x10

-8

m/%

DTA / V

-6

1.2x10

273

473

673

873
T/K

1073

1.1x10

-6

1.1x10

-6

Intensity (me = 18) / A

3.1.5. Catalyst reusability and regeneration


The conversion of MO and H2O2 as well as the selectivity for the
epoxide (ME) are shown as a function of reaction time for the chosen conditions, i.e., in liquid acetonitrile and at 323 K over TS-1
(ind.) as the catalyst in Fig. 5, left part. Both conversion of MO
and of H2O2 rise rapidly within the rst 5 h of the experiment
where after the conversion increases only slightly to reach values
of 87% and 93%, respectively (see also Table 2). Note, however, that
H2O2 is present in a vefold excess with respect to MO and, thus,
the vast majority of H2O2 is converted by unproductive decomposition. Concomitantly, the epoxide selectivity decreases steadily
from 92% to 87% at 24 h. This supports that the by-products are indeed formed via consecutive reactions from ME (see Fig. 2). Since,
however, the conversion slows down strongly after the rst 5 h of

reaction and since a complete conversion of MO was not reached in


any of the experiment, the question of catalyst deactivation arose.
In fact, it was reported that a deactivation of Ti-MCM-41 catalysts
occurs during oxidation of cyclohexene with aqueous H2O2 due to
adsorbed reaction residues hindering the access to the active centers and/or due to Ti leaching [41]. In the former case, the catalyst
could be regenerated by calcination to eliminate adsorbed reaction
residues on the active sites.
If the catalyst TS-1 (ind.) is reused three times (after removal by
ltration and drying in air at room temperature), the MO conversion steadily drops while the epoxide selectivity slightly increases
(Fig. 5, right part). By calcination of the catalyst in air at 673 K for
24 h after the third reuse, the initial activity and selectivity of the
catalyst can be completely recovered. This regeneration was further characterized by TG-DTA-MS analysis of the catalyst after
the third reuse (Fig. 6). The small weight loss (2 wt.%) up to
373 K is due to the loss of physisorbed water, consistent with
hydrophobic nature of the TS-1 surface. A more pronounced
weight loss of 11 wt.% together with a strong exothermicity in

Intensity (me = 44) / A

3.1.4. Inuence of catalyst particle size and mass


To more deeply investigate whether or not the MO conversion
with H2O2 over the industrial TS-1 catalyst is limited by mass
transport effects, the mass and the particle size of the catalyst were
varied. Upon increasing the catalyst mass from 10 to 300 mg, the
inital rates increase (Fig. 4, left part). In the absence of mass transport effects, the initial rates would be expected to increase linearly
with catalyst mass. However, a deviation of this expected linear increase for catalyst mass above 150 mg indicates the presence of
mass-transport limitations. Additionally, the initial rates decrease
with increasing catalyst particle size (Fig. 4, right part), corresponding to a decrease in MO conversion of 55% for particles
<0.2 mm to 13% for particles >1.6 mm. From the decrease of the
initial rates with particle size, the effective diffusion coefcient
of the reactant MO was calculated applying the Thiele Modulus
U (Fig. 4, right part, for calculation see ESI). The value found here,
i.e., Deff = 5.7  10 11 m2 s 1 is somewhat lower than that reported
by Zieverink et al. [40] for methyl oleate in the hydrogenation and
isomerization over an alumina supported palladium catalyst
(1.8  10 10 m2 s 1). The lower temperature and the difference in
size, geometry and tortuosity of the catalyst particle may account
for this lower effective diffusivity.
These ndings lead to the unambiguous conclusion that the
epoxidation of methyl oleate with H2O2 over TS-1 is strongly limited by diffusion into the catalyst particle. As the size of the TS-1
crystallites was not altered, it is the diffusion of the reactants to
the outer surface of the TS-1 crystallites that limits the conversion
rate.

1273

Fig. 6. TG-DTA-proles (top part) and MS analysis of the off-gas during thermogravimetric analysis of the catalyst TS-1 (ind.) after run 3 (see Fig. 5).

188

N. Wilde et al. / Microporous and Mesoporous Materials 164 (2012) 182189

80

K-M-units

conversion X or selectivity S / %

100

60

XMO
40

XFAME
SME

20

TS-1_s_60
TS-1_s_20

SeFAME

Ti-MCM-41
0
0

10

15

20

25

200

t/h

300

400

500

Fig. 7. Conversion of the substrates and selectivity for epoxidized products in the
epoxidation of methyl oleate (MO) or biodiesel (FAME) with hydrogen peroxide
over TS-1 (ind.) in acetonitrile at 323 K as a function of reaction time.

700

800

Fig. 8. DR-UVVis spectra of TS-1 with stacked morphology (TS-1_s_60, and TS1_s_20 from microwave irradiation for different times during synthesis) and of TiMCM-41.

DTA curve up to 1273 K is accompanied by the formation of carbon


dioxide and water. This supports that during heating the catalyst in
air, organic residues on the catalyst are removed. The deactivation
of TS-1 (ind.) in the epoxidation of MO is therefore attributed to
blocking of the active Ti-sites on the outer crystal surface by organic deposits, presumable formed by consecutive reactions of target
products, such as dimerization or oligomerization. These can, however, be completely removed by calcination in air.

(overall fraction of mono-, di- and triunsaturated compounds:


91 wt.%). As shown in Fig. 7, the conversion of biodiesel (FAME)
is only lower by ca. 10% in the rst 5 h than that of pure MO as
the reactant, but reaches a similar value (90% for FAME vs. 93%
for MO) after 24 h. The selectivity for the formation of epoxides
is also a little lower than with MO as the only reactant and
amounts to 76% (vs. 87% for MO) after 24 h (Fig. 7).
3.2.2. Epoxidation of biodiesel over TS-1 with stacked morphology
As an attempt to nd a more active catalyst for the biodiesel
epoxidation, TS-1 with stacked morphology was prepared according to [23]. These catalysts are comprised of particles with sizes
in the sub-micrometer scale and, thus, a high outer crystal surface.
They were previously reported to be superior catalysts for the
epoxidation of linear C6C12-olens with the C@C-bond in terminal
position with H2O2 with respect to conventional TS-1.
In this study, two samples were prepared using different irradiation times in microwave-assisted synthesis. The samples show
the typical XRD patterns for TS-1 (see ESI) and possess Ti contents

3.2. Conversion of biodiesel (FAME)


3.2.1. Epoxidation of biodiesel over TS-1 (ind.)
In order to evaluate whether commercial biodiesel can also be
epoxidized using aqueous H2O2 solution, the conversion was again
carried out over the catalyst TS-1 (ind.) under the same reaction
conditions as for the MO epoxidation, but with a FAME mixture derived from rapseed oil. This biodiesel consists predominantly of
methyl oleate (72 wt.%) and smaller fractions of di- and triunsaturated compounds such as methyl linolate and methyl linolenate

100

10

XFAM E
S eFA M E

80

60

40

20

TON

conversion X or selectivity S / %

600

Wavelength / nm

-M
Ti

-4

1
TS

-1

d
(in

.)
TS

-1

2
s_

0
TS

-1

6
s_

0
-M
Ti

-4

1
TS

-1

d
(in

.)
TS

-1

2
s_

0
TS

-1

6
s_

Fig. 9. Conversion of biodiesel XFAME and epoxide selectivity SeFAME (left part) turnover number TON (right part) in the epoxidation of biodiesel over different Ti-containing
catalysts in acetonitrile at 323 K after 24 h.

N. Wilde et al. / Microporous and Mesoporous Materials 164 (2012) 182189

of 0.5 and 0.7 wt.% and average crystallite sizes of 60 and 100 nm
(see ESI for SEM micrograph) for 20 (sample TS-1_s_20) and
60 min (sample TS-1_s_60) microwave irradiation, respectively.
The DR-UVVis spectra of these samples are typical for TS-1
and show an absorption band around 215 nm characteristic for Ti
in tetrahedral coordination (Fig. 8) [24]. In sharp contrast,
Ti-MCM-41 shows a broader, less intense band with maximum at
230 nm, a shoulder at 250 nm and another band with maximum
at 330 nm. This indicates, probably, the presence of higher coordinated Ti species [24,41] in this sample.
The conversion of biodiesel on the catalysts with the stacked
morphology (TS-1_s_20 and TS-1_s_60) is clearly lower than that
over commercial catalyst TS-1 (ind.), whereas the selectivity is only
a little lower (Fig. 9, left part). The activity of these catalysts is
much higher than that of Ti-MCM-41, as expected for the smaller
fraction of active sites with high epoxidation activity, i.e., framework incorporated Ti. If, however, the TON is compared, the catalysts from microwave-assisted synthesis are clearly superior to
the commercial TS-1 (ind.) (Fig. 9, right part). Especially, the sample obtained after 60 min microwave irradiation TS-1_s_60 exhibits a TON of 8.7 vs. 5.9 for TS-1 (ind.). One reason for this higher
activity is certainly the higher amount of active Ti sites on the outer surface of the small crystals of TS-1_s_60. Note, however, that on
the sample with smaller crystallites (TS-1_s_20), TON is lower than
for TS-1_s_60. This might be explained by a higher surface hydrophobicity for the sample from microwave treatment at longer time
[23]. The hydrophobicity of TS-1_s_60 might even be higher than
that of TS-1 (ind) contributing to the higher TON with respect to
the commercial catalyst.
4. Conclusions
The heterogeneously catalyzed epoxidation of fatty acid methyl
esters with aqueous hydrogen peroxide solution provides an
attractive route to chemicals and intermediates for a wide range
of applications such as polymer production. Using an industrial
TS-1 as the catalyst, methyl oleate conversion with hydrogen peroxide yields the epoxide with a selectivity of 87% at 93% conversion
after 24 h of reaction in liquid acetonitrile. Likewise, commercial
biodiesel can be epoxidized over TS-1 with 76% selectivity at 90%
conversion. TS-1 is superior to other Ti-containing catalysts such
as Ti-MCM-41 or TiOxSiO2 due to its high fraction of tetrahedrally
coordinated Ti on framework positions as the active sites. The TS-1
catalyst is subject to a deactivation by deposition of organic compounds. The activity can, however, be completely be restored by
calcination in air. Indeed, it was found out that diffusion limits
the conversion of FAME over TS-1. Therefore, catalysts with small
crystallite size and high outer surface area are benecial for high
activity. Here, TS-1 with crystallite sizes in the sub-micrometer
scale and with stacked morphology as obtained from microwaveassisted synthesis are shown to exhibit a signicantly higher turnover number than the industrial TS-1 catalyst. Further attempts to
obtain even more active catalysts for the epoxidation of biodiesel
with aqueous hydrogen peroxide solution should, thus, focus on
the preparation of TS-1 with nanstructured crystallites and a
highly accessible outer surface area, e.g., within hierarchically
structured pore systems.

189

Appendix A. Supplementary data


Supplementary data associated with this article can be found, in
the online version, at http://dx.doi.org/10.1016/j.micromeso.2012.
06.047.
References
[1] G. Knothe, R.O. Dunn, M.O. Bagby, Biomass 666 (1997) 172.
[2] J.O. Metzger, U. Biermann, W. Friedt, S. Lang, W. Luhs, G. Machmuller, M.R.
Klaas, H.J. Schafer, M.P. Schneider, Ang. Chem. Int. Ed. 39 (2000) 2206.
[3] U. Biermann, U. Bornscheuer, M.A.R. Meier, J.O. Metzger, H.J. Schafer, Ang.
Chem. Int. Ed. 50 (2011) 3854.
[4] DE 4201343A1, 22.07.1993, Henkel KGaA, B. Gruber.
[5] F.D. Gunstone, F.B. Padley, in: F.D. Gunstone (Ed.), Lipid Technologies and
Applications, Marcel Dekker, New York, 1997, p. 759.
[6] DE 4202758C2, 05.08.1993, Henkel KGaA, H. Kluth, P. Daute, J. Klein, R.
Grtzmacher, W. Klauck.
[7] A. Debal, G. Rafaralahitsimba, E. Ucciani, Fett Wissenschaft Technologie-Fat Sci.
Technol. 95 (1993) 236.
[8] M.A. Camblor, A. Corma, P. Esteve, A. Martinez, S. Valencia, Chem. Commun.
(1997) 795.
[9] G.A. Eimer, V.R. Elias, M.E. Crivello, E.R. Herrero, S.G. Casuscelli, Ind. Eng. Chem.
Res. 48 (2009) 9076.
[10] K. Burgess, B.S. Lane, Chem. Rev. 103 (2003) 2457.
[11] A.L.V. de P, B.F. Sels, D.E. De Vos, P.A. Jacobs, J. Org. Chem. 64 (1999) 7267.
[12] J.Y. Tang, L. Wang, G. Liu, Y. Liu, Y.Z. Hou, W.X. Zhang, M.J. Jia, W.R. Thiel, J. Mol.
Catal. A: Chem. 313 (2009) 31.
[13] S. Imamura, H. Sasaki, M. Shono, H. Kanai, J. Catal. 177 (1998) 72.
[14] A.C. Gomes, S.M. Bruno, S. Gago, R.P. Lopes, D.A. Machado, A.P. Carminatti, A.A.
Valente, M. Pillinger, I.S. Goncalves, J. Organomet. Chem. 696 (2011) 3543.
[15] M. Abrantes, P. Neves, M.M. Antunes, S. Gago, F.A.A. Paz, A.E. Rodrigues, M.
Pillinger, I.S. Goncalves, C.M. Silva, A.A. Valente, J. Mol. Catal. A: Chem. 320
(2010) 19.
[16] E. Poli, J.M. Clacens, J. Barrault, Y. Pouilloux, Catal. Today 140 (2009) 19.
[17] P.T. Anastas, J.C. Warner (Eds.), Green Chemistry Theory and Practice, Oxford
University Press, New York, 1998.
[18] M. Guidotti, N. Ravasio, R. Psaro, E. Gianotti, S. Coluccia, L. Marchese, J. Mol.
Catal. A: Chem. 250 (2006) 218.
[19] M. Guidotti, R. Psaro, N. Ravasio, M. Sgobba, E. Gianotti, S. Grinberg, Catal. Lett.
122 (2008) 53.
[20] M. Guidotti, E. Gavrilova, A. Galarneau, B. Coq, R. Psaroa, N. Ravasio, Green
Chem. 13 (2011) 1806.
[21] P.A.Z. Suarez, M.S.C. Pereira, K.M. Doll, B.K. Sharma, S.Z. Erhan, Ind. Eng. Chem.
Res. 48 (2009) 3268.
[22] J. Sepulveda, S. Teixeira, U. Schuchardt, Appl. Catal. A 318 (2007) 213.
[23] H.L. Jin, N.Z. Jiang, S.M. Oh, S.E. Park, Top. Catal. 52 (2009) 169.
[24] G.N. Vayssilov, Catal. Rev. Sci. Eng. 39 (1997) 209.
[25] A. Wroblewska, J. Mol. Catal. A: Chem. 229 (2005) 207.
[26] R.A. Sheldon, J. Chem. Technol. Biotechnol. 68 (1997) 381.
[27] A. Campanella, M.A. Baltanas, M.C. Capel-Sanchez, J.M. Campos-Martin, J.L.G.
Fierro, Green Chem. 6 (2004) 330.
[28] L. Wang, J.J. Zou, M.Y. Zhang, B. Zhu, X.W. Zhang, Z.T. Mi, Catal. Lett. 124 (2008)
139.
[29] Y.K. Park, S.J. Kim, N. You, J. Cho, S.J. Lee, J.H. Lee, J.K. Jeon, J. Ind. Eng. Chem. 17
(2011) 186.
[30] Y. Su, Y.M. Liu, L.C. Wang, M. Chen, Y. Cao, W.L. Dai, H.Y. He, K.N. Fan, Appl.
Catal. A 315 (2006) 91.
[31] L.A. Rios, P. Weckes, H. Schuster, W.F. Hoelderich, J. Catal. 232 (2005) 19.
[32] D. Srinivas, P. Manikandan, S.C. Laha, R. Kumar, P. Ratnasamy, J. Catal. 217
(2003) 160.
[33] P. Ratnasamy, V.N. Shetti, P. Manikandan, D. Srinivas, J. Catal. 216 (2003) 461.
[34] M.G. Clerici, G. Bellussi, U. Romano, J. Catal. 129 (1991) 159.
[35] A. Corma, P. Esteve, A. Martinez, J. Catal. 161 (1996) 11.
[36] M.G. Clerici, P. Ingallina, J. Catal. 140 (1993) 71.
[37] C.Y. Lin, K.H. Wang, Fuel 83 (2004) 507.
[38] A. Wroblewska, E. Lawro, E. Milchert, Ind. Eng. Chem. Res. 45 (2006) 7365.
[39] E. Karlsen, K. Schoffel, Catal. Today 32 (1996) 107.
[40] M.M.P. Zieverink, M.T. Kreutzer, F. Kapteijn, J.A. Moulijn, Ind. Eng. Chem. Res.
44 (2005) 9668.
[41] E.R. Herrero, G.A. Eimer, S.G. Casuscelli, G.E. Ghione, M.E. Crivello, Appl. Catal.
A 298 (2006) 232.

Vous aimerez peut-être aussi