You are on page 1of 10

Ind. Eng. Chem. Res.

2010, 49, 8125–8134 8125

Kinetics of Heterogeneous Catalytic Epoxidation of Propene with Hydrogen


Peroxide over Titanium Silicalite (TS-1)
Sang Baek Shin and David Chadwick*
Department of Chemical Engineering, Imperial College London, London SW7 2AZ, United Kingdom

The kinetics of the heterogeneous catalytic epoxidation of propene to propene oxide with hydrogen peroxide and
titanium silicalite (TS-1) as the catalyst was studied under mild conditions in an autoclave reactor using methanol/
water mixtures as the solvent. The effects of reactant concentration, solvent composition, catalyst loading, stirring
speed, and reaction temperature on the propene oxidation were investigated. The initial rates increased almost
linearly with catalyst loading. The apparent orders of reaction with respect to hydrogen peroxide and propene
were found to be 0.67 (0.17-0.70 wt %) and 0.63 (0.3-0.8 bar), respectively. The apparent activation energy was
found to be 25.8 kJ mol-1. Methanol content in the solvent was shown to have a dual influence on the reaction
rate, enhancing the solubility of propene and inhibiting the epoxidation reaction by competitive adsorption.

1. Introduction only a limited number of works have been devoted to detailed


kinetics studies.48,49
Propene oxide is an important chemical intermediate in the
chemical industry that is mainly used for the production of In this article, we report the kinetics of the epoxidation of
polyether polyols, propene glycols, and propene glycol ethers.1,2 propene to propene oxide using hydrogen peroxide as the oxidant,
The propene oxide industry has employed two different types of methanol/water as the solvent, and titanium silicalite as the catalyst
commercial processes for several decades: the chlorohydrin process under operating conditions different from those used in previous
and the hydroperoxidation process. However, both of these studies. In particular, the reaction conditions were extended to lower
processes have multiple reaction steps and several essential concentrations of reactants and milder pressures. Previous studies
disadvantages: The chlorohydrin process suffers from environ- have examined methanol/water and isopropanol solvents,48,49 which
mental liabilities caused by a toxic oxidant and waste. On the other were found to increase the reaction rate as a result of the enhanced
hand, the economy of the hydroperoxidation process primarily solubility of propene. In the present study, unexpected results were
depends on the marketability of coproducts. Therefore, direct obtained concerning the effect of the methanol content where
epoxidation of propene with ex situ or in situ produced hydrogen methanol was found to inhibit the initial rates. It is shown that this
peroxide has recently attracted much attention as a more environ- effect can be explained by a reaction model that includes a term
mentally benign and profitable chemical process.1-10 In this for methanol competitive adsorption and allows for the increased
approach, the reaction is carried out under mild conditions, and solubility of propene as a function of methanol concentration.
theoretically, only water is generated as byproduct. Even though Furthermore, the data are demonstrated to favor a Langmuir-
the first commercial-scale propene oxide plant based on hydrogen Hinshelwood-type formulation of the rate equation.
peroxide recently started its operation,11 a number of studies are
still being conducted to gain a better understanding of catalysis 2. Experimental Section
and develop more efficient reactor concepts. It seems clear that 2.1. Catalyst Preparation. TS-1 catalyst was prepared ac-
the next generation after the hydrogen peroxide-to-propene oxide cording to the method described in the literature.20,28 A solution
(HPPO) process would be the direct gas-phase epoxidation of of 1.5 g of tetraethyl orthotitanate (TEOT, g95.0%, Fluka) was
propene by oxygen or a mixture of oxygen and hydrogen, but gradually added to 45 g of tetraethyl orthosilicate (TEOS, g99.0%,
currently available catalysts have difficulty in attaining sufficient Fluka) with magnetic stirring for 30 min. An aliquot of 100 g of
conversion for commercial application.1,2,7,12-19 20 wt % tetrapropylammonium hydroxide solution (TPAOH,
Titanium silicalite (TS-1) catalyst has been widely investigated Fluka) was then gradually added to the mixed solution under stirring
because it can promote the formation of propene oxide with a high for an additional 30 min. After the mixture had been kept at 60 °C
selectivity of greater than 95% and a high hydrogen peroxide for 3 h, 56 g of deionized water was added to the solution. The
conversion.3,4,9,10,20-27 Clerici and co-workers first demonstrated composition of the final solution was the same as the molar reagent
that TS-1 can be used efficiently as a catalyst for the epoxidation ratios employed by Clerici et al.:20 SiO2/TiO2 ) 32.7, TPAOH/
of various olefins, including propene, using hydrogen peroxide in SiO2 ) 0.46, and H2O/SiO2) 35. The resulting solution was
methanol under mild conditions.20 Most TS-1 catalysts have been transferred to an autoclave in an oven and crystallized at 175 °C
prepared following the procedure described in the original patent,28 for 48 h without stirring. After the solution had been cooled to
but a variety of attempts have also been made to find easier and room temperature, the crystalline product was separated by
less expensive methods for preparing TS-1 catalyst.29-37 Several centrifugation. Washing with deionized water and separation were
catalysts involving Pd and Pt supported on TS-1 have also been repeated five times. The solid product was dried at 100 °C for 2 h
studied for the liquid-phase epoxidation of propene and olefins.4,38-47 and then finally calcined at 550 °C for 5 h in an air-atmosphere
Although the liquid-phase epoxidation of propene using hydrogen furnace. The resulting catalyst was ground into a powder with a
peroxide and TS-1 catalyst has been studied for several decades, pestle and mortar.
* To whom correspondence should be addressed. Tel.: +44 (0)20 2.2. Characterization. The crystallographic structure of TS-1
7594 5579. Fax: +44 (0)20 7594 5637. E-mail: d.chadwick@ was identified on a PANalytical X’Pert PRO X-ray powder
imperial.ac.uk. diffractometer equipped with a nickel filter and Cu KR radiation
10.1021/ie100083u  2010 American Chemical Society
Published on Web 07/27/2010
8126 Ind. Eng. Chem. Res., Vol. 49, No. 17, 2010

(λ ) 1.541 Å). Bragg angles between 5° and 50° were scanned


with a 0.0042° step size. Fourier transform infrared (FT-IR) spectra
were measured on a Perkin-Elmer Spectrum 1 spectrometer using
the KBr wafer technique. UV-visible measurements were per-
formed on a Perkin-Elmer Lambda 25 spectrometer. Scanning
electron microscopy (SEM) images were recorded on a Leo Gemini
1525 instrument to inspect the morphology of the catalyst. The
lattice structure of TS-1 was confirmed by transmission electron
microscopy (TEM) image taken on a JEOL JEM 2010 instrument.
Thermogravimetric analysis (TGA) was performed on a TA Q500
thermogravimetric analyzer by measuring weight changes of
catalyst under nitrogen gas. The temperature was raised from 30
to 1000 °C at a rate of 5 °C min-1. Brunauer-Emmett-Teller
(BET) analysis data was obtained by nitrogen adsorption and
desorption in a Micromeritics TriStar 3000 apparatus.
2.3. Epoxidation of Propene. A 200 mL thick-walled glass
Büchi autoclave system constructed by Strata Technology Ltd. was
used for batch experiments. The catalytic oxidations were carried Figure 1. X-ray diffraction pattern of TS-1 catalyst.
out in a magnetic-drive-stirred autoclave reactor that was immersed
in a water bath to control the reaction temperature. In a typical
run, a mixture consisting of 0.68 g of 35 wt % hydrogen peroxide,
35 g of methanol, and 34.32 g of deionized water was placed in
the autoclave reactor. A small sample was taken to determine the
initial concentration of hydrogen peroxide before the introduction
of the catalyst. Then, 0.28 g of TS-1 catalyst was added, and the
reactor was closed. The rotational speed of the gas entrainment
stirrer was set at 1000 rpm and maintained throughout experiment.
The reactor was heated to a predetermined temperature (40 °C).
When the set temperature was reached, the reactor was charged
with 10% propene in nitrogen at a predetermined total pressure (7
bar) to initiate the reaction. The reaction products were sampled
at regular time intervals through a dip tube. It is noted that a range
of low concentrations of hydrogen peroxide and propene was
employed in this work, whereas most previous studies conducted
in a batch reactor have used pure propene.9,10,20,23,26,49,50
The concentrations of hydrogen peroxide were determined by
iodometric titration. Liquid samples were analyzed on a gas
chromatograph (Shimadzu GC-2014) equipped with a flame Figure 2. FT-IR spectrum of TS-1 catalyst.
ionization detector and a capillary column (30 m × 0.32 mm) with 3. Results and Discussion
nitroterephthalic acid-modified polyethylene glycol (0.25 µm) as
the stationary phase. Propene oxide was the main product; propene 3.1. Catalyst Characterization. The analysis results showed
glycol and propene glycol monomethyl ether were the byproducts. good agreement with the data indicated in the original patent.28
Conversion of hydrogen peroxide, utilization of hydrogen peroxide, Figure 1 shows the XRD pattern of TS-1 catalyst in which single
selectivity to propene oxide, selectivity to byproduct, and yield of peaks at 24.4° and 29.3° demonstrate a conversion from a
propene oxide are defined by the following equations monoclinic symmetry (silicalite) to an orthorhombic symmetry
(titanium silicalite). The peak intensity magnitudes are known
nH0 2O2 - nH2O2 to be related to the crystallinity of TS-1, which strongly affects
XH2O2 ) (1) the catalyst activity, but it was not easy to predict the degree of
nH0 2O2 activity only from the peak intensity magnitudes.51 Figure 2
shows the FT-IR spectrum of TS-1 catalyst. There is an
nPO + nbyproducts absorption band at 971.4 cm-1 that is not present in the spectrum
UH2O2 ) (2) of either pure silicalites or titanium oxides. In general, the
nH0 2O2XH2O2 intensity of this band increases with the amount of titanium that
substitutes the silicon in the framework of silicalites.
nPO Figure 3 shows the UV-vis spectrum of TS-1. An absorption
SPO ) (3) band near 215 nm indicates four-coordinated tetrahedral Ti
nPO + nbyproducts
species in TS-1. No other peaks are observed, which demon-
nbyproducts strates that there is no octahedral Ti or anatase TiO2.9,52
Sbyproducts ) (4) In the SEM images shown in Figure 4, well-defined, uniform
nPO + nbyproducts cubic crystallites with rounding edges are observed to be
distributed homogeneously. It is known that TS-1 crystallizes
YPO ) XH2O2SPO (5) in the form of slightly rounded cubes when crystallization occurs
in a static autoclave and that this shape is more active for the
The initial rates of the chemical reaction were calculated from epoxidation of propene than other shapes such as spheres and
tangential slopes at t ) 0 using exponential curve fitting. twinned coffins.51,53 The size of the particles was in the range
Ind. Eng. Chem. Res., Vol. 49, No. 17, 2010 8127

Figure 3. UV-vis spectrum of TS-1 catalyst.

Figure 5. TEM micrograph of TS-1 catalyst.

Figure 6. TGA profile of TS-1 catalyst.

up to 1000 °C indicates the good thermal stability of TS-1. The


weight loss up to 100 °C should be attributed to the loss of
physisorbed water. The small weight loss up to 100 °C suggests
that the prepared TS-1 catalyst had a relatively high hydrophobic
property, which is known to enhance the activity of TS-1.3,9,10
BET analysis gave a surface area of 421 m2 g-1 and a pore
volume of 0.255 cm3 g-1, which are comparable to the values
reported in the literature.3,7,9,10,54,55
3.2. TS-1 Performance at Normal Operating Conditions.
Figure 4. SEM micrographs of TS-1 catalyst. Prior to the propene epoxidation experiments in the Büchi
autoclave system, the effects of the reactor metal and catalyst
of 200-300 nm. The activity of TS-1 is also strongly influenced on the decomposition of hydrogen peroxide were examined. The
by the size of the catalyst particles: the smaller the particle size, decomposition of hydrogen peroxide by TS-1 catalyst was
the more effective the catalyst.51,53 The TS-1 prepared in this observed to be 3.3% after 5 h at normal operating conditions,
study therefore had a suitable shape and particle size to exhibit and the effect of the reactor metal was negligible (only 0.3%
sufficient activity for the epoxidation of propene largely free after 5 h).
of intraparticle mass-transfer resistance. The TEM micrograph Figure 7 shows the performance of TS-1 at normal operating
in Figure 5 exhibits well-ordered lattice fringes of the MFI conditions in the autoclave reactor. TS-1 catalyst prepared in
structure of TS-1, which is indicative of high crystallinity. This this study showed a high catalytic activity comparable to the
also confirms that the TS-1 catalyst had uniformly sized results reported in the previous literature.20,49 After 5 h, 92%
nanopores. of the hydrogen peroxide had been consumed. The selectivity
Figure 6 shows the TGA profile of the TS-1 catalyst. The to propene oxide gradually decreased with time to 93% after
weight loss up to 100 °C and the total weight loss up to 1000 5 h. TS-1 is a weak acid catalyst, which can lead to the solvoysis
°C were 1.2% and 2.9%, respectively. The small weight loss of propene oxide. All of the observed byproducts were generated
8128 Ind. Eng. Chem. Res., Vol. 49, No. 17, 2010

Figure 8. Effect of agitation speed on the initial rate of propene epoxidation.


Reaction conditions: 40 °C, 7 bar (10% C3H6/N2), 0.35 wt % H2O2, 50 wt
% CH3OH, 0.28 g of TS-1, 100-1500 rpm stirring.

As can be seen in Figure 8, the rate of reaction was independent


of agitation speed when the agitation speed was greater than
500 rpm. This is clear evidence for the absence of gas-liquid
mass-transfer resistance at and above a stirring speed of 500
rpm. Thus, 1000 rpm was employed for the speed of agitation
in all further experiments.
The Mears criterion allows for the estimation of possible
mass-transfer limitations from the bulk fluid to the surface of
the catalyst.57 The Mears criterion calculated for the reactor
system was less than 0.15, which shows that external diffusion
resistance can be neglected.
To check the possibility of mass-transfer limitations, the
effectiveness factor was calculated from the Thiele modulus and
the effective diffusivity. The effectiveness factor was very close
Figure 7. Propene epoxidation in an autoclave reactor at normal operating to 1, which demonstrates that internal mass diffusion inside the
conditions: (a) concentrations of hydrogen peroxide, propene oxide, and catalyst particle had negligible resistance.57 These findings
byproduct and (b) conversion of hydrogen peroxide, utilization of hydrogen confirm that the chemical reaction on the surface of the catalyst
peroxide, selectivity to propene oxide, selectivity to byproduct, and yield
of propene oxide. Reaction conditions: 40 °C, 7 bar (10% C3H6/N2), 0.35 was rate limiting. The equations and calculated parameter values
wt % H2O2, 50 wt % CH3OH, 0.28 g of TS-1, 1000 rpm stirring. used to check for mass-transfer limitations are shown in the
Appendix.
from propene oxide by the cleavage of the oxirane ring by 3.4. Catalyst Loading. The effects of catalyst loading on
nucleophiles.21 In the hydrolysis of propene oxide, propene the initial rate and selectivity of propene epoxidation were
glycol is formed from water, and propene glycol monomethyl studied in a range of catalyst concentrations (Figure 9). The
ether is formed from reaction with methanol. The utilization of initial rate was found to be proportional to the catalyst loading,
hydrogen peroxide was kept between 88% and 96% for 5 h. and this linearity reconfirms the absence of mass-transfer
The main byproducts were 1-methoxy-2-propanol and 2-meth- resistance. In Figure 9a, it seems that the data point at the largest
oxy-1-propanol, and small amounts of 3-methoxy-1-propanol reciprocal of catalyst loading has a significant upward deviation
and 1,2-propanediol were also detected. from a regression line. This is because, at a low level of catalyst
3.3. Mass-Transfer Limitations. The epoxidation of propene loading, the measured rate of reaction in a batch reactor can be
is a heterogeneous gas-liquid-solid-catalyzed reaction. To affected easily by several factors such as slight catalyst loss
investigate the true kinetics, mass-transfer resistances between from splashing.
the phases must be removed. In this study, a self-aspirating The initial rate increased linearly with the amount of catalyst,
stirrer was used to intensify the contact between the phases by but the selectivity to propene oxide decreased with an increase
recycling gas inside the autoclave. The mass-transfer resistances in the catalyst loading, particularly at higher catalyst loadings
were controlled by the level of turbulence induced by stirring. (Figure 9b). The selectivity to propene oxide after 5 h was as
A Rushton turbine type of impeller with six vertical blades was low as 86% when 0.5 g of TS-1 was used.
employed to generate large vortices. The Reynolds number (Re) 3.5. Initial Concentration of Hydrogen Peroxide. The
was 33000 at an agitation speed of 1000 rpm. Therefore, the effect of the concentration of hydrogen peroxide on the initial
system was operated in a completely turbulent flow regime (Re rate is shown in Figure 10. The results indicate that the order
> 1000056). The system was maintained in turbulent flow even of reaction is 0.67 with respect to hydrogen peroxide. This
at 500 rpm (Re ) 16500) but fell into laminar flow below 300 finding is consistent with that of Clerici et al., who reported
rpm. that the reaction order with respect to hydrogen peroxide is
The effect of the agitation speed of the stirrer on the initial between zero and unity.20 Also, a similar fractional order can
rate of reaction was examined in the range of 100-1500 rpm. be indirectly confirmed when a power-law model is applied to
Ind. Eng. Chem. Res., Vol. 49, No. 17, 2010 8129

Figure 11. Effect of initial propene partial pressure on the initial rate of
propene epoxidation. Reaction conditions: 40 °C, 3-8 bar (10% C3H6/N2),
0.35 wt % H2O2, 50 wt % CH3OH, 0.28 g of TS-1.

As the initial concentration of hydrogen peroxide increased,


the selectivity to propene oxide decreased. The selectivity to
propene oxide after 5 h was 84% when the initial concentration
of H2O2 was 0.2 mol kg-1, which is twice the hydrogen peroxide
concentration at normal operating conditions. These results
suggest that the initial concentration of hydrogen peroxide
should be adjusted according to the propene feed concentration
in order to minimize side reactions of propene oxide in
commercial applications.
3.6. Initial Propene Partial Pressure. The results obtained
at various initial propene partial pressures (0.3-0.8 bar) are
presented in Figure 11, which indicates that the order with
respect to propene was 0.63. This result differs from those
reported in some published studies. Almost first order can be
obtained with respect to propene when a power-law model is
Figure 9. Effect of catalyst loading on (a) initial rate of propene epoxidation
and (b) selectivity to propene oxide. Reaction conditions: 40 °C, 7 bar (10% applied to the results presented by Liang et al.,49 and Lee et
C3H6/N2), 0.35 wt % H2O2, 50 wt % CH3OH, 0.14-0.50 g of TS-1, 1000 al.58 reported that propene follows pseudo-first-order kinetics
rpm stirring. over methyltrioxorhenium catalyst. This discrepancy might be
due to different solvents and higher concentrations of reactants
employed in the previous studies. The effect of the solvent is
discussed in the next section.
The production rate of propene oxide increased with increas-
ing propene partial pressure, as expected, but the selectivity to
propene oxide was highest when the propene partial pressure
was 0.7 bar. A possible explanation for this is that a propene
partial pressure of 0.7 bar provides the stoichiometric amount
of propene for the initial concentration of hydrogen peroxide,
which is favorable for maintaining high selectivity.
3.7. Concentration of Solvent. Good catalytic performance
can be achieved in solvents that are able to dissolve both propene
and hydrogen peroxide. It is known that methanol is the best
among many kinds of protic or aprotic solvents.20,21,25,26,50
Therefore, methanol/water mixtures were used as the solvent
in this study.
The variations in the propene oxide concentration and
selectivity with varying weight percentage of methanol are
presented in Figure 12. As shown in Figure 12b, there was only
Figure 10. Effect of initial hydrogen peroxide concentration on the initial
rate of propene epoxidation. Reaction conditions: 40 °C, 7 bar (10% C3H6/ a slight difference in the selectivity to propene oxide when the
N2), 0.17-0.70 wt % H2O2, 50 wt % CH3OH, 0.28 g of TS-1. water content was below 50%, but the selectivity decreased
dramatically as the water content increased beyond this con-
the results presented by Liang and colleagues.49 However, Thiele centration. In particular, the formation of 1,2-propanediol was
et al. reported that hydrogen peroxide was consumed according greatly promoted by water. This result is consistent with those
to pseudo-first-order kinetics.21 of previous researchers.20,25,48
8130 Ind. Eng. Chem. Res., Vol. 49, No. 17, 2010

Table 1. Henry’s Law Constants of Propene in Water and Methanol


at 40 °C59-61
solvent HE (atm) HE [mol/(L · atm)] HE,MeOH/HE,H2O
-3
H2O 6667 8.3 × 10 31.3
MeOH 90.5 2.6 × 10-1

Table 2. Additional Measurement of Initial Rates


H2O2 (wt %) PC3H6 (bar) CH3OH (wt %) r0 (mol L-1 gcat-1 min-1)
0.35 2.8 50 0.0165
1.05 0.7 50 0.0103
1.05 2.8 10 0.0167

Previous kinetic studies have sought to fit the rate of this


reaction to Eley-Rideal or Langmuir-Hinshelwood model
equations. For example, Liang et al.49 concluded that an
Eley-Rideal model with adsorption of hydrogen peroxide and
free propene was the best fit to their experimental data. The
proposed rate equation also allowed for competitive adsorption
by propene and the product propene oxide.49 To gain better
insight into the effects of the reactants, product, and methanol
concentration on the initial reaction rate, Eley-Rideal and
Langmuir-Hinshelwood model equations incorporating a term
for competitive adsorption of methanol and allowance for the
enhancement of propene solubility were used to fit to the present
experimental data. For example, a Langmuir-Hinshelwood
(L-H) model can be expressed by eq 6, where HE,H2O and
HE,MeOH are Henry’s law constants for water and methanol,
respectively. The denominator of eq 6 can be modified in the
case of a dual-site L-H model to give eq 7.

r0 )
kK1K2CH2O2(φH2OHE,H2O + φMeOHHE,MeOH)PC3H6
(6)
[1 + K1CH2O2 + K2(φH2OHE,H2O + φMeOHHE,MeOH)PC3H6 +

K3CMeOH]2
Figure 12. Effect of methanol concentration on (a) propene oxide production
and (b) selectivity to propene oxide. Reaction conditions: 40 °C, 7 bar (10% kK1K2CH2O2(φH2OHE,H2O + φMeOHHE,MeOH)PC3H6
C3H6/N2), 0.35 wt % H2O2, 0-70 wt % CH3OH, 0.28 g of TS-1. r0 ) (7)
(1 + K1CH2O2 + K3CMeOH)[1 + K2(φH2OHE,H2O +
It is known that the reaction rate of propene epoxidation
generally increases with increasing methanol content because φMeOHHE,MeOH)PC3H6 + K4CMeOH]
the solubility of propene in methanol is higher than that in
water.3,20,25,48 Contrary to expectations, in the present study, An Eley-Rideal model with free propene can be represented
the initial rate of propene oxide production was found to by eq 8, where allowance is also made for competitive
decrease with increasing amount of methanol with a high degree adsorption of propene following Liang et al.49 A similar equation
of linearity, as shown in Figure 12a. Interestingly, the initial can be derived for the case where hydrogen peroxide is free.
rate showed a maximum at low methanol concentration and was r0 )
fairly high even when there was no methanol (Figure 12a inset). kK1CH2O2(φH2OHE,H2O + φMeOHHE,MeOH)PC3H6
However, in water, the productivity to propene oxide for 5 h
was very low when compared with that in methanol/water 1 + K1CH2O2 + K2(φH2OHE,H2O + φMeOHHE,MeOH)PC3H6 + K3CMeOH
mixtures. The final concentration of propene oxide after 5 h (8)
was highest at 50% methanol content because side reactions
Table 1 summarizes the Henry’s law constants (HE) of
were promoted at less than 50% methanol and the reaction rate
propene with respect to water and methanol that were used to
was low at greater than 50% methanol.
consider the relative amounts of dissolved propene in the
These results showing a decreasing rate with increasing different solvent mixtures.59-61
methanol content are quite different from previously reported To augment the analysis, three additional measurements of
work on the influence of methanol and indicate that methanol initial rates were made (Table 2). Eley-Rideal and L-H model
can act as an inhibitor through competitive adsorption. The equations were evaluated, and statistical analysis was carried
observation of a slight maximum at low methanol content also out using all of the measured initial rates and a nonlinear least-
points to the more conventional effect of enhancing the solubility squares function of software R. Converged solutions were not
of propene. It seems likely that this inhibiting effect of methanol found for the Eley-Rideal models with competitive adsorption
was magnified by the mild conditions used in the present study of methanol as in eq 8. The nonlinear regression analysis showed
and masked in earlier works that used higher concentrations of that the initial rate data could be fitted best by a form of L-H
hydrogen peroxide and higher propene pressures. model equation (Figure 13). There is good agreement between
Ind. Eng. Chem. Res., Vol. 49, No. 17, 2010 8131

Figure 13. Experimental initial rates and fitting by the Langmuir-Hinshelwood models: (a) parity plot and (b) comparison with respect to methanol content.

Table 3. Kinetic Parameters and Statistical Results


t test (p value)
model kK1K2 K1 K2 K3 K4 F test (p value) RSS σ2
(I) L-H single-site 3.70 × 10-3 8.35 × 10-4 4.96 × 10-2 2.02 × 10-5 - 6.14 × 10-9 7.38 × 10-6 4.92 × 10-7
(II) L-H dual-site 1.53 × 10-5 2.83 × 10-3 1.90 × 10-3 1.11 × 10-2 1.18 × 10-5 3.11 × 10-10 2.21 × 10-6 1.47 × 10-7

Table 4. Parameters with 95% Confidence Interval for L-H Single- Table 5. Comparison of Conditions Used for Investigations into the
and Dual-Site Models at 40 °C Effect of Methanol Content
L-H single-site L-H dual-site ref(s) CH2O2 (mol/L) PC3H6 (atm) catalyst (g/L) temp (°C)
parameter units model model
this study 0.1 0.7 4 40
k mol L-1 gcat-1 min-1 1.66 ( 0.98 0.20 ( 0.056 Wang et al.,48 0.45 4 12 60
K1 L mol-1 9.74 ( 4.63 9.66 ( 5.31 Li et al.25
K2 L mol-1 4.15 ( 4.07 20.57 ( 10.65 Chen et al.3 1 1 9 25
K3 L mol-1 0.57 ( 0.17 0.12 ( 0.08 Clerici et al.20 0.5-1.75 4-8 4-8 20-60
K4 L mol-1 - 0.96 ( 0.26
propene and hydrogen peroxide concentrations spanning the
the experimental and predicted initial rates. The kinetic param- present and previously reported conditions, using eq 6 and the
eters and the results of t and F tests are given in Table 3. parameters of Table 4 (similar set of curves can be generated
According to the F test, the L-H dual-site model (model II) from eq 7). From this simulation, it can be seen that the
is statistically more significant than the single-site model (model dependence of the initial rates of propene epoxidation on the
I). However, model I also gives a small residual sum of squares methanol content can follow various trends, either increasing
(RSS) and variance (σ2), which leads to good fitting to the initial or decreasing, according to the levels of reactant concentrations.
rate data. Thus, it can be concluded that methanol competes This means that the increases in reaction rate with increasing
with propene and hydrogen peroxide for adsorption on single amount of methanol observed in the previous studies were a
sites or on dual sites of the catalyst surface. Based solely on consequence of the higher reactant concentrations used, which
the present results, it is not possible to differentiate unambigu- overcame the effect of competitive adsorption by methanol.
ously between the two L-H models. The confidence intervals
for model I and model II are presented in Table 4.
It is known that the high activity of TS-1 in propene
epoxidation is attributable in part to its hydrophobicity as it
preferentially adsorbs the less polar molecules propene and
H2O2. Methanol would be expected to be preferentially absorbed
on the active sites of TS-1 from water. The values of the
adsorption constants in Table 4 are in the order propene ≈ H2O2
> methanol, which is consistent with the polarities of these
species. Although the adsorption constant of methanol is
relatively small, the methanol concentration is large, giving rise
to a significant effect on the reaction rate. The TS-1 catalyst
prepared in this study contained only tetrahedral Ti and no
anatase TiO2 phase or extraframework Ti. On balance, the dual-
site L-H model (II) would seem to be more consistent with
the known structure of TS-1.
In Table 5, the operating conditions used in this work are
compared with those of previous studies that investigated the
effect of methanol content. Presented in Figure 14 are initial Figure 14. Initial rates of propene epoxidation at various reactant and methanol
rates with respect to methanol content calculated at various concentrations predicted by the Langmuir-Hinshelwood model (eq 6).
8132 Ind. Eng. Chem. Res., Vol. 49, No. 17, 2010

Figure 16. Arrhenius plot of propene epoxidation. Reaction conditions:


21-50 °C, 7 bar (10% C3H6/N2), 0.35 wt % H2O2, 50 wt % CH3OH, 0.28 g
of TS-1.

to be 25.8 kJ mol-1. The concentration of solvent including


methanol and water had a tradeoff effect on the concentration
of propene oxide because the relative amounts of methanol and
water affected the reaction rate and selectivity to propene oxide
differently, that is, the reaction rate decreased but the selectivity
increased with increasing methanol content.
The initial rates of propene oxide production were observed
to decrease with increasing methanol content from a maximum
at low methanol concentrations. This effect was shown to be
due to the enhancement of propene solubility by methanol and
the opposing effect of competitive adsorption by methanol.
Langmuir-Hinshelwood single-site and dual-site rate equations
were found to give a good fit to the initial rate data.

Figure 15. Effect of reaction temperature on (a) propene oxide production Acknowledgment
and (b) selectivity to propene oxide. Reaction conditions: 21-50 °C, 7 bar
(10% C3H6/N2), 0.35 wt % H2O2, 50 wt % CH3OH, 0.28 g of TS-1. Financial support by EPSRC (U.K.) is gratefully acknowledged.
The authors also thank Haseong Kim at Imperial College London
3.8. Reaction Temperature. Figure 15 shows the effect of
for the help with and discussion of the statistical analysis.
the reaction temperature on propene epoxidation. The initial rate
increased with an increase in temperature, but the selectivity to
propene oxide decreased, which led to low production rate of Nomenclature
propene oxide as the reaction proceeded. It is noteworthy that Ci ) concentration of species i (mol kg-1 or mol L-1)
there was a significant difference in the selectivity between 40 HE,i ) Henry’s law constant of propene for solvent i (mol L-1
and 50 °C. In particular, a large amount of 1,2-propanediol was atm-1)
produced at 50 °C. k ) reaction rate constant (mol L-1 gcat-1 min-1)
The Arrhenius plot shown in Figure 16 gives an activation Ki ) adsorption equilibrium constant of species i (L mol-1)
energy 25.8 kJ mol-1 for propene epoxidation over TS-1 ni ) number of moles of species i (mol)
catalyst. This apparent activation energy is close to that reported ni0 ) initial number of moles of species i (mol)
over TS-1 in isopropanol by Liang et al.49 Pi ) pressure of species i
r0 ) initial reaction rate (mol kg-1 h-1 or mol L-1 gcat-1 min-1)
4. Conclusions Si ) selectivity to product i
Ui ) utilization of species i
TS-1 catalyst prepared according to the original patent was
Xi ) conversion of species i
characterized by several analysis techniques and showed good
Yi ) yield of species i
performance for the selective oxidation of propene to propene
oxide in an autoclave reactor. Greek Letter
The change in operating conditions had a profound effect on φi ) volume fraction of solvent i in a solvent mixture
the propene oxide concentration and selectivity. The selectivity
to propene oxide decreased with increasing catalyst loading, Appendix
reactant concentration, water content, and reaction temperature,
which, in turn, led to a decrease of the propene oxide
A1. Liquid-Solid Mass-Transfer Limitations
concentration. The apparent orders of reaction with respect to
hydrogen peroxide and propene were found to be 0.67 and 0.63, The Wilke-Chang equation62 was used to calculate the diffusion
respectively. The apparent activation energy was determined coefficient of propene in a mixture of water and methanol
Ind. Eng. Chem. Res., Vol. 49, No. 17, 2010 8133
-8
7.4 × 10 (φMB) T 1/2 (2) Cheng, W. G.; Wang, X. S.; Li, G.; Guo, X. W.; Zhang, S. J. Highly
DAB ) Efficient Epoxidation of Propylene to Propylene Oxide over TS-1 Using
ηBVA 0.6 Urea + Hydrogen Peroxide as Oxidizing Agent. J. Catal. 2008, 255 (2),
343–346.
The dimensionless numbers, such as the Schmidt number, (3) Chen, L. Y.; Chuah, G. K.; Jaenicke, S. Propylene Epoxidation with
Reynolds number, and Sherwood number, and the mass-transfer Hydrogen Peroxide Catalyzed by Molecular Sieves Containing Framework
Titanium. J. Mol. Catal. A: Chem. 1998, 132 (2-3), 281–292.
coefficient were calculated according to the following equations57
(4) Laufer, W.; Meiers, R.; Holderich, W. Propylene Epoxidation with
ν Hydrogen Peroxide over Palladium Containing Titanium Silicalite. J. Mol.
Sc ) Catal. A: Chem. 1999, 141 (1-3), 215–221.
DAB (5) Beckman, E. J. Production of H2O2 in CO2 and Its Use in the Direct
Synthesis of Propylene Oxide. Green Chem. 2003, 5 (3), 332–336.
Nd2
Re ) (6) Liu, X. W.; Wang, X. S.; Guo, X. W.; Li, G.; Yan, H. S. Regeneration
ν of Lamina TS-1 Catalyst in the Epoxidation of Propylene with Hydrogen
Peroxide. Catal. Lett. 2004, 97 (3-4), 223–229.
Sh ) 2 + 0.6Re1/2Sc1/3 (7) Wang, Q. F.; Wang, L.; Chen, J. X.; Wu, W.; Mi, Z. T. Deactivation
kcdp and Regeneration of Titanium Silicalite Catalyst for Epoxidation of
Sh ) Propylene. J. Mol. Catal. A: Chem. 2007, 273 (1-2), 73–80.
DAB (8) Zhao, J. L.; Zhou, J. C.; Su, J.; Guo, H. C.; Wang, X. S.; Gong,
W. M. Propene Epoxidation with In-site H2O2 Produced by H2/O2 Non-
A Mears criterion of less than 0.15 indicates that there are equilibrium Plasma. AIChE J. 2007, 53, 3204–3209.
no liquid-solid mass-transfer limitations57 (9) Park, S.; Cho, K. M.; Youn, M. H.; Seo, J. G.; Baeck, S. H.; Kim,
T. J.; Chung, Y. M.; Oh, S. H.; Song, I. K. Epoxidation of Propylene with
-rA′ FpRn Hydrogen Peroxide over TS-1 Catalyst Synthesized in the Presence of
Cm ) Polystyrene. Catal. Lett. 2008, 122 (3-4), 349–353.
kcCA (10) Park, S.; Cho, K. M.; Youn, M. H.; Seo, J. G.; Jung, J. C.; Baeck,
S. H.; Kim, T. J.; Chung, Y. M.; Oh, S. H.; Song, I. K. Direct Epoxidation
The values of these parameters are collected in Table A1. of Propylene with Hydrogen Peroxide over TS-1 Catalysts: Effect of
Table A1. Calculation Results for Liquid-Solid Mass-Transfer Hydrophobicity of the Catalysts. Catal. Commun. 2008, 9 (15), 2485–2488.
Limitations (11) Short, P. L. BASF, Dow Open Novel Propylene Oxide Plant. Chem.
Eng. News 2009, 87 (11), 21.
parameter description units value (12) Klemm, E.; Dietzsch, E.; Schwarz, T.; Kruppa, T.; de Oliveira,
-1 A. L.; Becker, F.; Markowz, G.; Schirrmeister, S.; Schutte, R.; Caspary,
DAB diffusion coefficient cm s 2
2.5 × 10-5 K. J.; Schuth, F.; Honicke, D. Direct Gas-phase Epoxidation of Propene
Sc Schmidt number - 248 with Hydrogen Peroxide on TS-1 Zeolite in a Microstructured Reactor. Ind.
Re Reynolds number - 33029 Eng. Chem. Res. 2008, 47 (6), 2086–2090.
Sh Sherwood number - 687 (13) Oyama, S. T.; Zhang, X. M.; Lu, J. Q.; Gu, Y. F.; Fujitani, T.
kc mass-transfer coefficient cm s-1 685.9 Epoxidation of Propylene with H2 and O2 in the Explosive Regime in a
-rA′ rate of reaction kmol kgcat-1 s-1 6.684 × 10-5 Packed-bed Catalytic Membrane Reactor. J. Catal. 2008, 257 (1), 1–4.
Cm Mears criterion - 5.71 × 10-12 (14) Nijhuis, T. A.; Visser, T.; Weckhuysen, B. M. Mechanistic Study
A2. Internal Mass-Transfer Limitations into the Direct Epoxidation of Propene over Gold/Titania Catalysts. J. Phys.
Chem. B 2005, 109 (41), 19309–19319.
To calculate the effectiveness factor, the effective diffusivity (15) Zwijnenburg, A.; Makkee, M.; Moulijn, J. A. Increasing the Low
Propene Epoxidation Product Yield of Gold/titania-based Catalysts. Appl.
and Thiele modulus were calculated using the following Catal. A: Gen. 2004, 270 (1-2), 49–56.
equations,57 with approximate values for the porosity, constric- (16) Sinha, A. K.; Seelan, S.; Tsubota, S.; Haruta, M. Catalysis by Gold
tion factor, and tortuosity Nanoparticles: Epoxidation of Propene. Top. Catal. 2004, 29 (3-4), 95–
102.
DABεpσ (17) Suo, Z. H.; Jin, M. S.; Lu, J. Q.; Wei, Z. B.; Li, C. Direct Gas-
De ) Phase Epoxidation of Propylene to Propylene Oxide Using Air as Oxidant
τ
on Supported Gold Catalyst. J. Nat. Gas Chem. 2008, 17 (2), 184–190.


k1FcSa (18) Bravo-Suarez, J. J.; Lu, J.; Dallos, C. G.; Fujitani, T.; Oyama, S. T.
φ1 ) R Kinetic Study of Propylene Epoxidation with H2 and O2 over a Gold/
De Mesoporous Titanosilicate Catalyst. J. Phys. Chem. C 2007, 111, 17427–
17436.
3
η) (φ1 coth φ1 - 1) (19) Yap, N.; Andres, R. P.; Delgass, W. N. Reactivity and Stability of
φ12 Au in and on TS-1 for Epoxidation of Propylene with H2 and O2. J. Catal.
2004, 226 (1), 156–170.
The values of these parameters are collected in Table A2. (20) Clerici, M. G.; Bellussi, G.; Romano, U. Synthesis of Propylene
Oxide from Propylene and Hydrogen Peroxide Catalyzed by Titanium
Table A2. Calculation Results for Internal Mass-Transfer Silicalite. J. Catal. 1991, 129 (1), 159–167.
Limitations (21) Thiele, G. F.; Roland, E. Propylene Epoxidation with Hydrogen
parameter description units value Peroxide and Titanium Silicalite Catalyst: Activity, Deactivation and
Regeneration of the Catalyst. J. Mol. Catal. A: Chem. 1997, 117 (1-3),
εp catalyst porosity - 0.4 351–356.
σ constriction factor - 0.8 (22) Mantegazza, M. A.; Petrini, G.; Spanò, G.; Bagatin, R.; Rivetti, F.
τ tortuosity - 3 Selective Oxidations with Hydrogen Peroxide and Titanium Silicalite
De effective diffusivity m2 s-1 2.66 × 10-10 Catalyst. J. Mol. Catal. A: Chem. 1999, 146, 223–228.
k1 rate constant m3 m-2 s-1 2.11 × 10-10 (23) Guo, X. W.; Wang, X. S.; Liu, M.; Li, G.; Chen, Y. Y.; Xiu, J. H.;
Fc density of catalyst kg m-3 2370 Zhuang, J. Q.; Zhang, W. P.; Bao, X. H. Epoxidation of Propylene with
Sa surface area of catalyst m2 kg-1 421000 Dilute H2O2 over Titanium Silicalite Containing Trace Aluminum. Catal.
φ1 Thiele modulus - 0.003488
Lett. 2002, 81 (1-2), 125–130.
η internal effectiveness factor - 0.9999
(24) Li, G.; Wang, X. S.; Yan, H. S.; Liu, Y. H.; Liu, X. W. Epoxidation
Literature Cited of Propylene Using Supported Titanium Silicalite Catalysts. Appl. Catal.
A: Gen. 2002, 236 (1-2), 1–7.
(1) Nijhuis, T. A.; Makkee, M.; Moulijn, J. A.; Weckhuysen, B. M. (25) Li, G.; Meng, J. W.; Wang, X. S.; Guo, X. W. Effect of Solvents
The Production of Propene Oxide: Catalytic Processes and Recent Develop- on Propene Epoxidation Catalyzed by Titanium Silicalite. React. Kinet.
ments. Ind. Eng. Chem. Res. 2006, 45 (10), 3447–3459. Catal. Lett. 2004, 82 (1), 73–80.
8134 Ind. Eng. Chem. Res., Vol. 49, No. 17, 2010

(26) Liu, X. W.; Wang, X. S.; Guo, X. W.; Li, G. Effect of Solvent on (45) Danciu, T.; Beckman, E. J.; Hancu, D.; Cochran, R.; Grey, R.;
the Propylene Epoxidation over TS-1 Catalyst. Catal. Today 2004, 93-95, Hajnik, D.; Jewson, J. Direct Synthesis of Propylene Oxide with CO2 as
505–509. the Solvent. Angew. Chem., Int. Ed. 2003, 42, 1140–1142.
(27) Clerici, M. G. TS-1 and Propylene Oxide, 20 Years Later. Oil Gas (46) Wang, Q. F.; Wang, L.; Mi, Z. T. Influence of Pt-Pd/TS-1 Catalyst
Eur. Mag. 2006, 32 (2), 77–82. Preparation on Epoxidation of Olefins with Hydrogen Peroxide. Catal. Lett.
(28) Taramasso, M.; Perego, G.; Notari, B. Preparation of porous 2005, 103 (1-2), 161–164.
crystalline synthetic material comprised of silicon and titanium oxides. U.S. (47) Chen, Q. L.; Beckman, E. J. One-pot Green Synthesis of Propylene
Patent 4,410,501, 1983. Oxide Using in Situ Generated Hydrogen Peroxide in Carbon Dioxide.
(29) Kraushaar, B.; van Hooff, J. H. C. A New Method for the Green Chem. 2008, 10 (9), 934–938.
Preparation of Titanium-Silicalite (TS-1). Catal. Lett. 1988, 1 (4), 81–84. (48) Wang, X. S.; Guo, X. W.; Li, G. Synthesis of Titanium Silicalite
(30) Gao, H. X.; Suo, J. S.; Li, S. B. An Easy Way to Prepare Titanium (TS-1) from the TPABr System and Its Catalytic Properties for Epoxidation
Silicalite-1 (TS-1). J. Chem. Soc., Chem. Commun. 1995, (8), 835–835. of Propylene. Catal. Today 2002, 74 (1-2), 65–75.
(49) Liang, X. H.; Mi, Z. T.; Wu, Y. L.; Wang, L.; Xing, E. H. Kinetics
(31) Serrano, D. P.; Uguina, M. A.; Ovejero, G.; Vangrieken, R.; of Epoxidation of Propylene over TS-1 in Isopropanol. React. Kinet. Catal.
Camacho, M. Synthesis of TS-1 by Wetness Impregnation of Amorphous Lett. 2003, 80 (2), 207–215.
SiO2-TiO2 Solids Prepared by the Sol-Gel Method. Microporous Mater. (50) Zhang, Z. G.; Kang, J. N.; Wang, Y. Q. Effects of Organic Solvent
1995, 4 (4), 273–282. Addition on the Epoxidation of Propene Catalyzed by TS-1. React. Kinet.
(32) Uguina, M. A.; Serrano, D. P.; Ovejero, G.; Vangrieken, R.; Catal. Lett. 2007, 92, 49–54.
Camacho, M. Preparation of TS-1 by Wetness Impregnation of Amorphous (51) van der Pol, A. J. H. P.; van Hooff, J. H. C. Parameters Affecting
SiO2-TiO2 Solids: Influence of the Synthesis Variables. Appl. Catal. A: the Synthesis of Titanium Silicalite-1. Appl. Catal. A: Gen. 1992, 92 (2),
Gen. 1995, 124 (2), 391–408. 93–111.
(33) Gontier, S.; Tuel, A. Synthesis of Titanium Silicalite-1 Using (52) Fan, W. B.; Duan, R. G.; Yokoi, T.; Wu, P.; Kubota, Y.; Tatsumi,
Amorphous SiO2 as Silicon Source. Zeolites 1996, 16 (2-3), 184–195. T. Synthesis, Crystallization Mechanism, and Catalytic Properties of
(34) Li, G.; Guo, X. W.; Wang, X. S.; Zhao, Q.; Bao, X. H.; Han, X. W.; Titanium-Rich TS-1 Free of Extraframework Titanium Species. J. Am.
Lin, L. W. Synthesis of Titanium Silicalites in Different Template Systems Chem. Soc. 2008, 130 (31), 10150–10164.
and Their Catalytic Performance. Appl. Catal. A: Gen. 1999, 185 (1), 11– (53) van der Pol, A. J. H. P.; Verduyn, A. J.; van Hooff, J. H. C. Why
18. Are Some Titanium Silicalite-1 Samples Active and Others Not. Appl. Catal.
(35) Kim, K. Y.; Ahn, W. S.; Park, D. W.; Oh, J. H.; Lee, C. M.; Tai, A: Gen. 1992, 92 (2), 113–130.
W. P. Microwave Synthesis of Titanium Silicalite-1 Using Solid Phase (54) Keshavaraja, A.; Ramaswamy, V.; Soni, H. S.; Ramaswamy, A. V.;
Precursors. Bull. Korean Chem. Soc. 2004, 25 (5), 634–638. Ratnasamy, P. Synthesis, Characterization, and Catalytic Properties of
(36) Ahn, W. S.; Lee, K. Y. Extensions in the Synthesis and Catalytic Micro-Mesoporous, Amorphous Titanosilicate Catalysts. J. Catal. 1995,
Application of Titanium Silicalite-1. Catal. SurV. Asia 2005, 9 (1), 51–60. 157 (2), 501–511.
(37) Chen, P.; Chen, X. S.; Tanaka, K.; Kita, H. A Novel and Less- (55) Li, Y. G.; Lee, Y. M.; Porter, J. F. The Synthesis and Characteriza-
Expensive Preparation of Titanium Silicalite-1 Membrane. Chem. Lett. 2007, tion of Titanium Silicalite-1. J. Mater. Sci. 2002, 37 (10), 1959–1965.
36 (8), 1078–1079. (56) Sinnott, R. K. Coulson & Richardson’s Chemical Enginering, Vol.
6: Chemical Engineering Design; Butterworth-Heinemann: Oxford, U.K.,
(38) Meiers, R.; Dingerdissen, U.; Holderich, W. F. Synthesis of 2005.
Propylene Oxide from Propylene, Oxygen, and Hydrogen Catalyzed by (57) Fogler, H. S. Elements of Chemical Reaction Engineering; Prentice-
Palladium-Platinum-Containing Titanium Silicalite. J. Catal. 1998, 176 Hall: Upper Saddle River, NJ, 1992.
(2), 376–386. (58) Lee, H.-J.; Shi, T.-P.; Busch, D. H.; Subramaniam, B. A Greener,
(39) Meiers, R.; Holderich, W. F. Epoxidation of Propylene and Direct Pressure Intensified Propylene Epoxidation Process with Facile Product
Synthesis of Hydrogen Peroxide by Hydrogen and Oxygen. Catal. Lett. Separation. Chem. Eng. Sci. 2007, 62 (24), 7282–7289.
1999, 59 (2-4), 161–163. (59) Yaws, C. L.; Yang, H.-C. Thermodynamic and Physical Property
(40) Hoelderich, W. F. ’One-Pot’ Reactions: A Contribution to Envi- Data; Gulf Publishing Company: Houston, TX, 1992.
ronmental Protection. Appl. Catal. A: Gen. 2000, 194, 487–496. (60) Miyano, Y.; Fukuchi, K. Henry’s Constants of Propane, Propene,
(41) Hoelderich, W. F.; Kollmer, F. Oxidation Reactions in the Synthesis Trans-2-butene and 1,3-Butadiene in Methanol at 255-320 K. Fluid Phase
of Fine and Intermediate Chemicals Using Environmentally Benign Oxidants Equilib. 2004, 226, 183–187.
and the Right Reactor System. Pure Appl. Chem. 2000, 72 (7), 1273–1288. (61) Perry, R. H.; Green, D. W. Perry’s Chemical Engineers’ Handbook;
(42) Grey, R. A.; Rangasamy, P. Aqueous epoxidation process using McGraw-Hill: New York, 2007.
modified titanium zeolite. U.S. Patent 6,194,591, 2001. (62) Reid, R. C.; Prausnitz, J. M.; Poling; B. E. The Properties of Gases
(43) Jenzer, G.; Mallat, T.; Maciejewski, M.; Eigenmann, F.; Baiker, & Liquids; McGraw-Hill: New York, 1998.
A. Continuous Epoxidation of Propylene with Oxygen and Hydrogen on a ReceiVed for reView January 13, 2010
Pd-Pt/TS-1 Catalyst. Appl. Catal. A: Gen. 2001, 208 (1-2), 125–133. ReVised manuscript receiVed May 28, 2010
(44) Laufer, W.; Hoelderich, W. F. Direct Oxidation of Propylene and Accepted July 1, 2010
Other Olefins on Precious Metal Containing Ti-catalysts. Appl. Catal. A:
Gen. 2001, 213 (2), 163–171. IE100083U

You might also like