You are on page 1of 11

Reac Kinet Mech Cat

DOI 10.1007/s11144-010-0177-z

Effects of boehmite synthesis conditions


on the epoxidation of styrene

_
Ismail Boz • Özge Kerkez

Received: 9 November 2009 / Accepted: 16 March 2010


Ó Akadémiai Kiadó, Budapest, Hungary 2010

Abstract The influence of synthesis conditions on the properties of nanocrystal-


line boehmite catalysts was studied for styrene epoxidation reaction in liquid phase.
At hydrothermal synthesis of boehmite, effects of pH, aging time and precursor
aluminum salts were studied by X-ray diffraction (XRD), fourier transform infrared
spectroscopy (FTIR), thermogravimetric analysis (TGA) and total surface area.
Characterization results showed that boehmite samples were in the nanosize range
and single-phase with high crystallinity. Styrene epoxidation reaction in liquid
phase was performed over boehmite catalysts at 80 °C in ethanol, using 30%
aqueous hydrogen peroxide solution and urea ? 30% aqueous hydrogen peroxide
solution as oxidants. The progress of the reaction was followed by gas chroma-
tography (GC). Higher styrene conversion and higher epoxide selectivities were
achieved when urea was mixed with H2O2 solution. Boehmite catalyst samples aged
for 8 h resulted in higher epoxide selectivities.

Keywords Epoxidation  Styrene  Boehmite  Urea–hydrogen peroxide

Introduction

Catalytic epoxidation of alkenes with hydrogen peroxide from a view point of green
organic synthesis is a challenging problem in comparison to the established
epoxidation reactions. The strained three-membered ring of epoxides easily reacts
with a nucleophile. Unlike ordinary ethers, epoxides are quite reactive. By the
reaction of a wide variety of nucleophiles, epoxides can be transformed into useful
chemicals [1, 2]. Production of epoxide by the reaction of alkenes with hydrogen

_ Boz (&)  Ö. Kerkez


I.
Chemical Engineering Department, Faculty of Engineering, Istanbul University,
Avcilar, 34320 Istanbul, Turkey
e-mail: ismailb@istanbul.edu.tr

123
_ Boz, Ö. Kerkez
I.

peroxide (H2O2) is an ideal synthetic route since H2O2 is inexpensive, easy to


handle, and possesses a high content of active oxygen. Alkene epoxidation with
H2O2 preferably requires a heterogeneous catalyst. The important step is to
synthesize a heterogeneous catalyst using hydrogen peroxide as oxidant.
Epoxidation of styrene is a commercially important reaction for the production of
styrene oxide, an important organic intermediate. Earlier studies reported for the
epoxidation of alkenes were based on the use of TS-1 [3, 4], TiO2–SiO2 [5],
hyrotalcites [6], alumina supported Mn(II) complexes [7] using different oxidizing
agents, such as TBHP [8], aqueous H2O2 [5, 9]. In the present work, we report the
effect of synthesis conditions on the properties of boehmite and the changes in
conversion and selectivity of styrene oxide. We also report that boehmite, a cheaper,
active/selective catalyst and less sensitive to water, was developed and utilized in
the epoxidation reaction.

Experimental

Synthesis of the catalysts and characterization

Aluminum nitrate, Al(NO3)39H2O ([95% Merck), and aluminum acetate,


Al(CH3COO)3 ([98.9% Baker & Adamson) were used as starting chemicals.
Aluminum nitrate (0.2 M) and aluminum acetate (0.2 M) solutions were made by
dissolving the aluminum salts in 100 mL of deionized water. NH3 solution (5 M)
was added to the solution drop by drop under vigorous stirring to achieve pH = 7
and 10. White precipitates were filtered to prevent anions attaching to the surface
and washed with deionized water. Then, precipitates were taken in water again and
redispersed. The aging of the precipitated particles was carried out in a Teflon lined
autoclave at 150 °C. After cooling the content, the precipitates were subsequently
filtered and washed with distilled water several times [10].
X-ray diffraction (XRD) patterns were taken at room temperature using a Rigaku
Ultima ? diffractometer (Cu Ka radiation with k = 0.15418 nm). The diffraction
intensity of boehmite samples was measured in the 2h range between 10° and 80°
with 2°/min scan speed. The crystallite sizes were calculated using the Scherer
equation:
kk
Dhkl ¼
ðb cos hÞ
where k is a constant between 0.8 and 1.39, k is the wavelength of the X-rays, b is
the pure full width of the diffraction line at half of the maximum intensity and h is
the Bragg angle.
Fourier transform infrared (FT-IR) spectra were recorded using a Thermo Nicolet
380 spectrometer in the wave number range from 400 to 4,000 cm-1, with samples
prepared by the KBr method, where the sample to KBr ratio was 1:100.
Thermogravimetric analysis of the samples was conducted in Shimadzu TG-50A
under flowing air at a heating rate of 10 °C/min up to 1,000 °C. The total surface area

123
Effects of boehmite synthesis conditions on the epoxidation of styrene

was determined by a single point BET method using a Costech Instruments


Sorptometer 1042. Catalysts were labeled as follows with a three digit code. The first
digit is either 1 or 2, 1 represents Al(NO3)39H2O and 2 represents Al(CH3COO)3
being used as precursor. The second digit represents pH at which preparation was
carried out. Finally, the last digit represents the aging time in hours. For instance,
1-7-2 is an acronym for the catalyst prepared from Al(NO3)39H2O precursor,
precipitated in pH = 7 and aged for 2 h.

Catalytic reactions

The catalytic epoxidation of styrene over boehmite catalysts was carried out in a
stirred glass reactor (volume: 50 cm3) under reflux at 80 °C using aqueous H2O2
and urea added aqueous H2O2 solution (30% H2O2 in water). Initially, 9 mL ethanol
as solvent was loaded into the reactor. The reactant (20 mmol styrene), 0.28 g
decane as internal standard and 0.2 g catalyst were added consecutively. After the
reaction mixture reached 80 °C, the reaction was timed right beginning at the
addition of oxidant (40 mmol H2O2).
Aliquots of the reaction mixture were taken after equal reaction times for a total
period of 5 h. The unconverted styrene and reaction products were analyzed using
Agilent Technologies 6890N gas chromatography equipped with a HP Innowax
(Crosslinked Polyethyleneglycol) column and a flame ionization detector. Products
were quantified using calibration curves obtained with standard solutions. The
conversion and product selectivity were calculated as follows: conversion
(%) = (moles of reactant converted 9 100) 7 (moles of reactant in feed); product
selectivity (%) = (moles of product formed) 9 100) 7 (moles of reactant con-
verted). Repeated experiments were carried out to determine reproducibility and it
was better than 15% in conversion calculations and 10% in selectivity.

Results and discussion

Catalyst characterization

X-ray diffraction analysis results are presented in Figs. 1 and 2. In XRD patterns, all
catalysts showed all the reflections matching with boehmite [JCPDS 21-1307]. No
other phases, such as, bayerite, diaspore, were observed in the X-ray diffraction
pattern. Music et al. also found that only boehmite phase was detected in the
precipitation after 5 h at 160 °C [11]. However, it was found in this study that 2 h of
aging time at 150 °C were enough to synthesize a single phase of boehmite.
Conversion of aluminum hydroxide to boehmite was dependent on synthesis
parameters, such as pH, temperature, type of starting aluminum salt and aging time
[11–13].
From the XRD analysis of intense peaks, it was determined that the crystallite
sizes of boehmite were in the nanosize range as tabulated in Table 1. The dominant
parameters, which affected the crystallite size of boehmite samples, were the aging
time and the precursor aluminum salt. Samples prepared from Al(CH3COO)3 had

123
_ Boz, Ö. Kerkez
I.

Fig. 1 XRD patterns of boehmite samples prepared from Al(NO3)39H2O

Fig. 2 XRD patterns of boehmite samples prepared from Al(CH3COO)3

smaller crystallite sizes in contrast to samples prepared from Al(NO3)39H2O, which


had relatively larger crystallite sizes. The influence of the anion on crystallite size
could be explained with the size of anions, structure and hydration degree of anion
[14]. At both conditions of pH = 7 and 10, the crystallite size of the samples were
increased with increasing aging time from 2 to 8 h.
FTIR spectra of all samples are presented in Figs. 3 and 4. The shoulder at
3,482 cm-1 and the broad band at 1,636 cm-1 were assigned to the stretching and
bending modes of adsorbed water [15–17]. In Fig. 3, the strong band at 1,385 cm-1
was attributed to the presence of nitrate anions and the absorption band at
approximately 1,426 cm-1 in Fig. 4 was due to the stretching modes of C=O

123
Effects of boehmite synthesis conditions on the epoxidation of styrene

Table 1 Properties and preparation conditions of catalysts prepared by hydrothermal method


Catalysts Styrene Styrene oxide Crystallite size Surface area Weight Weight
codes conversion selectivity in nm of d(020) in m2/gcat loss (%) loss (%)
(%) (%) 25–110 °C 110–600 °C

1-7-2 81.2 8.6 4.38 225 3.65 18.22


1-7-8 55.0 23.3 5.68 200 3.83 18.80
1-10-2 77.8 14.0 3.65 256 5.80 20.63
1-10-8 43.0 23.2 5.90 169 3.36 18.24
2-7-2 72.3 10.2 3.70 221 3.30 19.76
2-7-8 53.0 18.2 5.37 164 2.22 18.40
2-10-2 83.0 13.3 2.70 268 5.00 17.00
2-10-8 48.0 21.0 4.30 171 3.33 15.63

Fig. 3 FTIR spectra of boehmite samples prepared from Al(NO3)39H2O

functional group of acetate anion. The FTIR spectrum of the catalyst samples gave
an intense broad band with peaks at 3,094 and 3,292 cm-1, which were assigned to
the symmetric and asymmetric stretching vibrations of (Al)O–H, respectively.
The absorption band at 1,068 cm-1 and the shoulder at 1,160 cm-1 were
assigned to the symmetric and asymmetric bending vibrations of (Al)O–H,
respectively. Generally, stretching modes and bending modes of O–H group were
expected at 3,600–3,400 and 1,100–900 cm-1. The shifting of the stretching mode
towards lower frequency (about 300 cm-1) and the bending mode towards higher
frequency suggested that the boehmite samples were tightly hydrogen-bonded [16].
The two weak bands observed at 2,099 and 1,971 cm-1 were assigned to a
combination band. Although strict attention was paid to the removal of anions,
peaks belonging to the corresponding anions show that there are still some remnants
of anions. The presence of large amount of water attached to the surface was also of
significance.

123
_ Boz, Ö. Kerkez
I.

Fig. 4 FTIR spectra of boehmite samples prepared from Al(CH3COO)3

Fig. 5 Thermo gravimetric weight loss pattern of boehmite sample (1-7-2)

Major weight losses of two steps were observed in TGA curve given in Fig. 5, at
about 100 and 400 °C. The first weight loss at 100 °C was related to the removal of
physically adsorbed water, while the second weight loss at 400 °C was attributed to
dehydroxylation and nitrate decomposition.
If single-phase boehmite is decomposed by the release of one mole of water, a
total weight loss of ca. 15% is expected [17]. As deduced from decomposition
reactions occurring between 100 and 400 °C, weight loss from TGA curves was
around 15%. In all cases, approximately 15% weight losses were calculated, thereby
proving that all the samples were in the form of boehmite with different water
uptakes.

123
Effects of boehmite synthesis conditions on the epoxidation of styrene

Surface areas of the boehmite samples were also reported in Table 1. Boehmite
catalyst samples prepared by the hydrothermal method had high specific surface
areas. Potdar et al. [17] determined the specific surface area of commercial Capatal-
B boehmite samples prepared at similar synthesis conditions. Among the prepared
samples, 1-10-2 and 2-10-2 had the highest surface areas. If we associate this with
weight losses of 1-10-2 and 2-10-2 at 25–110 °C from Table 1, the samples which
had higher surface areas showed higher weight losses at TG analysis, too. Another
inference from Table 1 was that surface area of samples decreased with increasing
aging time from 2 to 8 h in contrast with crystallite size. Crystallite sizes of
boehmite, as determined by XRD, were in the nanosize range, it was concluded that
the particles consisted of nanosized boehmite crystallites.

Epoxidation of styrene

The epoxidation of styrene using boehmite as heterogeneous catalyst and hydrogen


peroxide as the oxidant at atmospheric pressure can be influenced by such factors as
the surface area in the samples, the hydrothermal synthesis conditions, the nature of
the solvent and the reaction temperature. In this respect, several preliminary
experiments were performed in order to elucidate the effect of selected variables on
the epoxidation reaction of styrene. First, a blank test (with no catalyst) was
performed leading to null conversion, which showed that a catalyst was essential for
this reaction. The epoxidation of styrene was carried out with various reaction times
(1–24 h) but for most reaction tests, the conversion of styrene and yield as well as
selectivity of styrene oxide slightly changed with increasing reaction time after 5 h.
The reaction profiles for the epoxidation of styrene over boehmite catalysts prepared
at different pH and aging times, using urea and hydrogen peroxide solution mixture
as oxidant were depicted in Figs. 6, 7, 8, and 9. The highest styrene conversion of
81.20% after 5 h was observed with catalyst 1-7-2, whereas a conversion of 55%
was obtained with catalyst 1-7-8 after 5 h. Although a lower styrene conversion was
obtained for catalyst 1-7-8, higher styrene oxide selectivity (23.27%) was achieved
with catalyst 1-7-2. The same trend exists for catalysts 1-10-2 and 1-10-8 pairs.
Namely, all boehmite samples which were aged for 8 h showed higher styrene oxide
selectivities in comparison to those obtained by the samples aged for 2 h (Table 1).
These results clearly showed that the aging time was an important parameter which
affected the catalytically active species on the surface of boehmite for epoxidation.
For the epoxidation reactions of styrene over boehmite catalysts, H2O2 solution
with urea resulted in better results than using the aqueous H2O2 solution only. The
reaction results were obtained by 1-7-8 catalyst for aqueous H2O2 solution only and
urea added aqueous H2O2 solution. When urea and H2O2 solution (1:1 mol) was
used, the styrene conversion increased from 28 to 55%, as well as the selectivity to
styrene oxide increased from 0 to 24%. A similar trend was obtained for catalyst
2-10-2, the styrene conversion increased from 42 to 83% by changing H2O2 to
urea ? H2O2 solution (1:1 mol). When urea was added to the H2O2 solution, the
benzaldehyde selectivity was decreased as expected. This finding was explained by
the proposal that urea and hydrogen peroxide make up an intermediate complex, so
hydrogen peroxide decomposes in the reaction mixture in a controlled manner

123
_ Boz, Ö. Kerkez
I.

Fig. 6 Catalytic performance of boehmite samples prepared from Al(NO3)39H2O in the epoxidation of
styrene

Fig. 7 Catalytic performance of boehmite samples prepared from Al(CH3COO)3 in the epoxidation of
styrene

thereby not leading to further reactions giving by-product benzaldehyde [18]. In a


similar study, Laha and Kumar [3] explained that urea added to aqueous hydrogen
peroxide acted not only as a dehydrating agent but also as a buffer for the system,
which minimized further acid-catalyzed isomerization and hydrolysis of the desired
epoxide. Choudhary et al. performed styrene epoxidation with commercial
boehmites dried at 200 °C without the removal of reaction water and using

123
Effects of boehmite synthesis conditions on the epoxidation of styrene

Fig. 8 The changes in the selectivity of styrene epoxide catalyzed by boehmite samples prepared from
Al(NO3)39H2O

Fig. 9 The changes in the selectivity of styrene epoxide catalyzed by boehmite samples prepared from
Al(CH3COO)3

aqueous H2O2 solution 50%. The styrene conversion was very low (\10%) and the
selectivity for the styrene oxide (*1%) was also very poor [9]. In this study, with
catalyst 2-10-2 showed higher styrene conversion and styrene oxide selectivity,
although we have used aqueous 30% H2O2 solution. If epoxidation reactions were
performed with the removal of water, the selectivities would be higher. Without
removing the water from the reaction system, urea acted as dehydrating agent by

123
_ Boz, Ö. Kerkez
I.

doing so, exposing the catalytically active surfaces to epoxidation reactions.


Addition of reactive compounds strongly affects the product yield. In a recent study,
it was found that addition of trifluoroacetic acid (TFA) lead to the noticeable
acceleration of all oxidation reactions and enhancement of the product yield [19].
The composition of products of the olefin oxygenation (epoxide/diol ratios, etc.)
strongly depends on the nature of the substrate [20]. In our study, we have found
that there was no other isomer other than epoxides, styrene epoxide.
The styrene conversion correlates well with the surface area but not with average
pore size, suggesting that under present study, the reaction was not limited by
diffusion of reactants and products. All catalysts essentially exhibited the same
ratios of direct epoxidation as well as selectivities to styrene epoxide relative to their
surface areas.
Thus, the optimum conditions for obtaining maximum yield of styrene oxide can
be summarized as follows: catalyst, 2-10-2; reaction time, 5 h; urea to H2O2 ratio
1:1.

Conclusions

Boehmite catalysts, for the catalytic epoxidation of styrene in liquid phase have
been synthesized by the hydrothermal method from aqueous solutions of abundant
aluminum salts. The hydrothermal method produced single-phase boehmite
crystallites with high surface area and high crystallinity. Boehmite samples were
in the nanosize range. The crystallite size of boehmite powders were found to
depend on the aging time. The boehmite sample which was prepared from
aluminum acetate at pH = 10 and for 2 h of aging time showed higher conversion
and selectivity values with only aqueous hydrogen peroxide solution in respect to
the styrene conversion and epoxide selectivity. Higher styrene conversion and
higher epoxide selectivities were determined in the reactions when urea was
admixed to H2O2 oxidant.

Acknowledgement This work was supported by Research Fund of the Istanbul University. Project
number T-1398.

References

1. Jones CW (1999) Applications of hydrogen peroxide and derivatives. RSC, Cambridge


2. Sheldon RA, Bekkum HV (2001) Fine chemicals through heterogeneous catalysis. Wiley-VCH,
Weinheim
3. Laha SC, Kumar R (2002) Highly selective epoxidation of olefinic compounds over TS-1 and TS-2
redox molecular sieves using anhydrous urea-hydrogen peroxide as oxidizing agent. J Catal 208:339–
344
4. Laha SC, Kumar R (2001) Selective epoxidation of styrene to styrene oxide over TS-1 using urea-
hydrogen peroxide as oxidizing agent. J Catal 204:64–70
5. Oki AR, XU Q, Shpeizer B, Clearfield A, Qui X, Kirumakki S, Tichy S (2007) Synthesis, charac-
terization and activity in cyclohexene epoxidation of mesoporous TiO2–SiO2 mixed oxides. Catal
Commun 8:950–956

123
Effects of boehmite synthesis conditions on the epoxidation of styrene

6. Kirm I, Medina F, Rodriguez X, Cesteros Y, Salagre P, Sueiras J (2004) Epoxidation of styrene with
hydrogen peroxide using hydrotalcites as heterogeneous catalysis. Appl Catal A 272:175–185
7. Salavati-Niasari M, Farzaneh F, Ghandi M (2002) Oxidation of cyclohexene with tert-butylhydro-
peroxide and hydrogen peroxide catalyzed by alumina-supported manganese(II) complexes. J Mol
Catal A 186:101–107
8. Choudhary VR, Jha R, Chaudhari NK, Jana P (2007) Supported copper oxide as a highly active/
selective catalyst for the epoxidation of styrene by TBHP to styrene oxide. Catal Commun 8:1556–
1560
9. Choudhary VR, Patil NS, Chaudhari NK, Bhargava SK (2005) Epoxidation of styrene by anhydrous
hydrogen peroxide over boehmite and alumina catalysts with continuous removal of the reaction
water. J Mol Catal A 227:217–222
10. Butler EG, Gu X, He JY, Kaya C (2002) Nanostructured ceramic powders by hydrothermal synthesis
and their applications. Microporous Mesoporous Mater 54:37–49
11. Music S, Dragcevic D, Popovic S, Vdovic N (1998) Microstructural properties of boehmite formed
under hydrothermal condition. Mater Sci Eng B52:145–153
12. Mishra D, Anand S, Panda RK, Das RP (2000) Hydrothermal preparation and characterization of
boehmites. Mater Lett 42:38–45
13. Okada K, Nagashima T, Kameshima Y, Yasumari A, Tsukada T (2002) Relationship between for-
mation conditions, properties and crystalline size of boehmite. J Colloid Interf Sci 253:308–314
14. Mishra D, Anand S, Panda RK, Das RP (2002) Effect of anions during hydrothermal preparation of
boehmites. Mater Lett 53:133–137
15. Priya GK, Padmaja P, Warrier KGK, Damodaran AD, Aruldhas G (1997) Dehydroxylation and high
temperature phase formation in sol–gel boehmite characterized by fourier transform infrared spec-
troscopy. J Mater Sci Lett 16:1584–1587
16. Hochepied JF, Nortier P (2002) Influence of precipitation conditions (pH and temperature) on the
morphology and porosity of boehmite particles. Powder Technol 128:268–275
17. Potdar HS, Jun K, Bae JW, Kim S, Lee Y (2007) Synthesis of nano-sized porous c-alumina powder
via a precipitation/digestion route. Appl Catal A 321:109–116
18. Rinaldi R, Schuchardt U (2005) On the paradox of transition metal-free alumina-catalysed epoxi-
dation with aqueous hydrogen peroxide. J Catal 236:335–345
19. Mandelli D, do Amaral ACN, Kozlov YN, Shul’pina LS, Bonon AJ, Carvalho WA, Shul’pin GB
(2009) Hydrogen peroxide oxygenation of saturated and unsaturated hydrocarbons catalyzed by
montmorillonite or aluminum oxide. Catal Lett 132:235–243
20. Bonon AJ, Mandelli D, Kholdeeva OA, Barmatova MV, Kozlov YN, Shul’pin GB (2009) Oxidation
of alkanes and olefins with hydrogen peroxide in acetonitrile solution catalyzed by a mesoporous
titanium-silicate Ti-MMM-2. Appl Catal A 365:96–104

123

You might also like