You are on page 1of 8

View Article Online / Journal Homepage / Table of Contents for this issue

Catalysis Dynamic Article Links

Science & Technology


Cite this: Catal. Sci. Technol., 2012, 2, 1417–1424

www.rsc.org/catalysis PAPER
Mesoporous Zr-SBA-15 as a green solid acid catalyst for the Prins
reactionw
Dong Minh Do, Stephan Jaenicke and Gaik-Khuan Chuah*
Received 13th February 2012, Accepted 25th March 2012
DOI: 10.1039/c2cy20084h
Published on 27 March 2012. Downloaded on 17/03/2017 12:09:01.

Mesoporous Zr-SBA-15 platelets were prepared with different pore dimensions from 4 to 8 nm
by the simple procedure of varying the amount of water in the synthesis gel. Narrow pore size
distributions were obtained for water–tetraethoxysilane ratios between 208 and 639, but samples
formed using a lower ratio of 100 had a broad pore size distribution. Thermogravimetric
measurements showed that the interaction between the Pluronic template and the inorganic
framework was affected by the amount of water in the synthesis gel. More zirconium was
incorporated into the silica framework when the synthesis was conducted in a more dilute system.
The Zr-SBA-15 obtained from this synthesis forms platelets with relatively short channels.
The catalytic activity was tested for C–C-coupling (Prins reaction). The terpene alcohol Nopol,
the product of an intermolecular Prins reaction between b-pinene and paraformaldehyde, could
be obtained with excellent selectivity. The mesoporous structure of the catalyst together with the
presence of zirconium in the silica framework, which confers strong Lewis acidity as well as
weak Brønsted acidic sites, are essential for the activity and selectivity of the reaction.

Introduction catalysts have been prepared by loading this metal onto high
surface area supports like MCM-41, SBA-15 or kenyaite
Acid-catalyzed reactions are important in petrochemical and (a sodium silicate) through ion exchange, incipient wetness
fine chemical synthesis.1 Using solid acids instead of mineral impregnation and chemical vapor deposition14–17 and by incor-
acids offers green synthetic routes to the desired chemicals, due poration into the silica framework of mesoporous SBA-15.18 In
to ease of product isolation, minimization of waste in work-up particular, Corma and Renz19 reported that Sn-MCM-41 was very
and reuse of catalysts. Hence, efficient and selective solid acids active in the Prins reaction of b-pinene with paraformaldehyde,
are desired. An example of an acid-catalyzed C–C coupling forming Nopol with a 94% selectivity at 93% conversion.
reaction is the Prins reaction where aldehydes are added to The use of butylnitrile instead of toluene as a solvent improved
alkenes.2 Different products such as 1,3-diols, 1,3-dioxanes or the selectivity by moderating the strength of the acidic sites at
unsaturated alcohols can be formed, depending on the reaction the catalyst.
conditions.3,4 Various acids have been reported such as hydro- Non-tin-containing heterogeneous catalysts reported for Nopol
chloric acid, alkyl aluminium chlorides,5,6 stannic chloride,7–9 synthesis include mesoporous iron phosphate,20 mesoporous
indium chloride in ionic liquids,10 and heteropolyacids.11 The Zn-Al-MCM-41,21 ZnCl2 impregnated Indian Montmorillonite,22
Prins reaction of b-pinene with paraformaldehyde is used in the and Fe–Zn double cyanide.23 However, mesoporous iron
synthesis of Nopol, a bicyclic unsaturated primary alcohol phosphate gave only Nopol yields of 2–85% at a catalyst to
(Scheme 1). Its ester with acetic acid, nopyl acetate, is an substrate ratio of 3.3 : 5 mmol.20 An even higher catalyst to
artificial fragrance compound which is found in the formulations substrate molar ratio (6.6 : 5) was required to reach 100% conver-
of many household products such as pesticides, detergents, soaps sion. Similarly, the use of Al-MCM-41 and Zn-Al-MCM-41 gave
and polishes.12 Normally, Nopol is synthesized in homogeneous Nopol yields of 27–84% but at a catalyst : substrate ratio of 0.2 g
systems using either ZnCl2 or acetic acid as the catalyst, or by (B 3.3 mmol) : 0.75 mmol.21
autoclaving the mixture of formaldehyde and b-pinene for
several hours at 150–230 1C.13 As tin was found to be active
for the homogeneously-catalyzed reaction,7–9 various supported

Department of Chemistry, National University of Singapore, 3 Science


Drive, Singapore 117543. E-mail: chmcgk@nus.edu.sg;
Fax: 65 67791691; Tel: 65 6516 2918
w Electronic supplementary information (ESI) available: Fig. S1–S6.
See DOI: 10.1039/c2cy20084h Scheme 1 Prins condensation of b-pinene and paraformaldehyde.

This journal is c The Royal Society of Chemistry 2012 Catal. Sci. Technol., 2012, 2, 1417–1424 1417
View Article Online

We have found that hydrous zirconia and Zr-zeolite beta are for 8 h (or 24 h) before hydrothermal treatment at 100 1C for
good catalysts for the intramolecular Prins cyclisation of 24 h in a Teflon-lined autoclave. The solid was recovered by
citronellal to isopulegols.24,25 These catalysts possess strong filtration, washed, dried at 100 1C, and calcined at 550 1C for
Lewis acid sites and weak Brønsted sites, both of which are 8 h to remove the organic template. The amount of water was
essential for the reaction.24 Further studies using mesoporous varied from 6.67 ml to 56.67 ml to give H2O/Si molar ratios of
TUD-1 with framework incorporation of Al and Zr showed 100, 208, 423 and 639. The molar composition of the synthesis
the presence of synergy between the Brønsted and Lewis gel was 0.017 Pluronic : 1 TEOS : 0.10 ZrOCl28H2O : 7.77
acid sites when applied as catalysts in the cyclisation of HCl : 100–639 H2O. The samples are labeled as 10Zr-w-t where
citronellal.26 However, when used in the more challenging 10 denotes the Si/Zr ratio, w the H2O/Si ratio and t represents
intermolecular Prins reaction between b-pinene and paraformal- the aging time of the synthesis gel. A catalyst with a lower
dehyde, no synergy was observed.27 To investigate if there zirconium content, 50Zr-423-24, was also prepared.
are other suitable catalysts besides the tin-based ones for inter- A procedure for Al-SBA-1544 was adapted for the prepara-
molecular Prins reactions, we focused on zirconium-containing tion of non-platelet mesoporous zirconium silica (10ZrMPS,
materials. The bulky size of b-pinene necessitates the use of Si/Zr = 10). The molar composition of the gel was 0.016
mesoporous catalysts, as previous studies with this substrate have Pluronic : 1 TEOS : 0.10 ZrOCl28H2O : 0.56 HCl : 77 H2O. A
Published on 27 March 2012. Downloaded on 17/03/2017 12:09:01.

shown a higher reaction rate with Sn-MCM-41 than with mixture of 2 g Pluronic P123 in 27.6 ml water was stirred at
microporous Sn-beta.19 room temperature for 3 h before adding 2.4 ml of 5 M HCl,
Among the mesoporous materials, SBA-15 with tunable 4.82 ml TEOS and 0.696 g zirconium oxychloride. The
pore sizes of 4–10 nm arranged in a 2D-hexagonal p6mm solution was stirred for 4 h, hydrothermally treated in a
structure has received much attention in the past decade Teflon-lined autoclave at 100 1C for 2 days before calcination
because of its relatively large pore size and high hydrothermal at 500 1C for 8 h.
stability in comparison with other mesoporous silica materials.
While the synthesis of the pure silica mesoporous material is
Characterization of samples
easily carried out in the presence of an appropriate surfactant,
it has low catalytic activity due to the lack of acidic sites. The The zirconium and silicon content of the samples was determined
incorporation of other metal ions such as zirconium into the by inductively coupled plasma-atomic emission spectroscopy
framework provides Lewis acidity. Mesoporous zirconium- (ICP-AES) after dissolving a weighed amount in HF. X-ray
doped silica is of interest due to the catalytic activity of the diffraction measurements were carried out using a Siemens
metal28–33 and the possibility of forming strong solid acids by D5005 diffractometer equipped with a copper anode and variable
sulfation.34–37 Recently, several novel routes have been developed slits. A step size of 0.0041/step and 0.021/step was used for 2y
for direct synthesis of Zr-SBA-15. For example, Newalkar et al.38 between 0.5–61 and 6–701, respectively. The morphology of the
reported the direct synthesis of Zr-SBA-15 with Si/Zr of 10–80 samples was measured using a scanning electron microscope
using microwave irradiation. The materials formed had thicker (JEOL JSM-5200). The UV-Vis absorption spectra were
walls than those formed by conventional synthesis.39 Materials measured against a barium sulfate reference using a Shimadzu
with high zirconium content (Si/Zr 5–9) were prepared by Du spectrophotometer (UV-2450) equipped with a diffuse-reflectance
et al.40 using urea as a pH adjustor. Cheng and coworkers41 cell. 29Si MAS NMR spectra were measured on a Bruker
showed that the synthesis of Zr-SBA-15 from the zirconium DRX-400 wide-bore solid state spectrometer operating at a
precursor ZrOCl28H2O was possible in the absence of HCl. resonance frequency of 79.46 MHz with a spinning rate of
However, the addition of HCl changed the morphology of the 8 kHz, a pulse length of 4 ms and a recycle time of 20 s. 4 mm
materials from rod-like structures to hexagonal platelets with rotors were used and the 29Si chemical shifts are reported
short mesochannels of 150–350 nm in length, perpendicular to relative to tetramethylsilane. The interaction between Pluronic
the hexagonal planes.42 Due to their structure, these platelet and the inorganic silica–zirconia was studied using thermo-
materials were less susceptible to diffusion limitations and pore gravimetric analysis-mass spectrometry (TGA-MS, MS-Pfeiffer
blockage than conventional rod or fibre-like SBA-15. Hence, Thermo Star) and simultaneous differential thermal analysis-
these materials are potentially useful catalysts in the intermole- thermogravimetric analysis (DTA-TGA, SDT 2960). Prior to
cular Prins reaction, where both the nature and strength of the the measurements, the as-synthesized sample was kept at 100 1C
acidic sites are essential for the formation of Nopol.43 In this for 30 min before heating to 600 1C in air at a rate of 10 1C min1.
study, we varied the water content in the synthesis gel and The evolution of water (m/z 18) and carbon dioxide
examined its effect on the pore size of the materials. (m/z 44) was monitored by the online mass spectrometer.
The adsorption of b-pinene and formaldehyde on 10Zr-423-8
was investigated by DTA-TGA. The catalyst was first im-
Experimental pregnated with b-pinene or formaldehyde in toluene and dried
at room temperature under nitrogen before thermogravimetric
Preparation of zirconium-incorporated SBA-15
analysis. A blank run was carried out using only the catalyst.
First, 0.5 g Pluronic P123 was added to a solution containing The presence of residual products on the catalyst after the
3.33 ml concentrated HCl (12 M) in water. After stirring the reaction was similarly studied. A heating rate of 10 1C min1 in
mixture for 4 h to obtain a clear solution, 1.15 ml tetraethyl nitrogen from room temperature to 500 1C was used.
orthosilicate (TEOS) and 0.162 g zirconium oxychloride The nitrogen adsorption–desorption isotherms were carried
(ZrOCl28H2O) were added. The mixture was stirred at 35 1C out using a Micromeritics Tristar 3000. Samples were degassed

1418 Catal. Sci. Technol., 2012, 2, 1417–1424 This journal is c The Royal Society of Chemistry 2012
View Article Online

under nitrogen flow at 300 1C for 6 h before measurement. The


surface area was calculated according to Brunauer–Emmett–
Teller (BET) theory while the pore size distribution was obtained
by the Barrett–Joyner–Halenda (BJH) method. The total
amount of acidity was determined by temperature-programmed
desorption (TPD) of ammonia. After pretreatment in helium
(50 ml min1) at 500 1C for 2 h, the sample was cooled to 150 1C,
exposed to ammonia gas for 15 min and flushed with helium for
another 2 h to remove any physisorbed ammonia. The tempera-
ture was then increased at a rate of 10 1C min1 from 150 to
500 1C and the evolved gases were analyzed using an online mass
spectrometer (Balzers Prisma 200). The nature of the acid sites
was determined using infrared spectroscopy of adsorbed pyridine.
A self-supporting wafer of the sample (8–10 mg) was mounted in a
Pyrex IR cell with NaCl windows and degassed at 300 1C in a Fig. 1 XRD patterns of calcined Zr-SBA-15 materials (a) 10Zr-100-24
(b) 10Zr-208-24 (c) 10Zr-208-8 (d) 10Zr-423-24 (e) 10Zr-423-8 and
Published on 27 March 2012. Downloaded on 17/03/2017 12:09:01.

vacuum (103 mbar) for 2 h. After cooling to room temperature,


(f) 10Zr-639-24.
pyridine was introduced into the cell at 22 mbar. Following
evacuation for an hour, an IR measurement at 2 cm1 and
64 scans was made using a Bio-Rad FTS 165 FT-IR spectro- Table 1 Textural properties of Zr-SBA-15 samples prepared with
meter. Further measurements were made after the sample had different water amounts and times
been heated at 100 and 200 1C for 1 h at each temperature.
Surf. area/ Pore vol./ Pore diam./ ao / t/
Catalytic studies Sample m2 g1 cm3 g1 nm nm nm
10Zr-100-24 544 0.79 5.6–7.3 — —
For catalytic testing, 0.156 ml (1 mmol) b-pinene, 0.0606 g 10Zr-208-24 597 0.93 6.3 11.2 4.9
(2 mmol) paraformaldehyde and 5 mL toluene were added 10Zr-423-24 627 1.09 7.4 11.3 3.9
into a 2-necked round-bottom-flask equipped with a reflux 10Zr-639-24 711 1.34 8.0 12.2 4.2
10Zr-208-8 621 0.93 6.0 12.0 6.0
condenser. The reaction mixture was heated under stirring in an
10Zr-423-8 629 1.18 7.6 11.9 4.3
oil bath before adding 50 mg of catalyst. The reaction temperature 50Zr-423-24 909 1.31 6.3
was varied between 80–110 1C. Samples were withdrawn at regular 10ZrMPS 616 0.85 5.4 — —
time intervals and analyzed by gas chromatography. The products pffiffiffi
ao ¼ 2d100 = 3, t: wall thickness.
were identified by gas chromatography–mass spectrometry
(GCMS). The reaction order in b-pinene and paraformaldehyde
was studied at 100 1C by keeping one substrate concentration zirconia was detected, suggesting that zirconium is either well-
at 4 mmol and varying the other from 0.25 to 4 mmol in 15 mL incorporated into the silica matrix or the zirconia crystallites
toluene. To test for reuse, the catalyst was recovered by are below the XRD detection limit of B4 nm.
centrifugation after the reaction. Its activity for further batch All samples have a high surface area (544–711 m2 g1) and
reactions was tested following regeneration by either (i) washing large total pore volume (0.79–1.34 cm3 g1) (Table 1). The
with toluene or (ii) treatment with H2O2 solution at 40 1C overnight nitrogen adsorption–desorption curves of the samples are typical
followed by drying at 100 1C. of Type IV isotherms. Hysteresis loops occur at P/P1 B0.6–0.8,
indicative of mesopores between 4–8 nm (Fig. 2). The vertical and
almost parallel adsorption and desorption branches for all samples
Results and discussion except 10Zr-100-24 show that the pores are narrowly distributed
with a spread of 0.5 to 1 nm about the mean pore size.
Textural properties
The pores in 10Zr-100-24, which was prepared with low
Fig. 1 shows the low angle XRD patterns of calcined Zr-SBA-15. water content in the synthesis gel, had a wider distribution,
Except for 10Zr-100-24 prepared with the lowest H2O/Si ratio of from 3.5 to 8.4 nm. The amount of water in the synthesis gel
100, the other samples prepared with higher H2O content showed affects the pore size. As the water content in the synthesis gel
the typical (100), (110) and (200) diffraction peaks of the increased, the resulting materials had bigger pores. The mean
2D-hexagonal p6mm structure. These diffraction peaks were pore diameter in Zr-208-24 was 6.3 nm and increased to
already present after only 8 h aging time, although the 8.0 nm for Zr-639-24.
intensity of the peaks was lower than in samples aged for From the isotherms of the calcined oxides, an aging time of
24 h. The peaks shifted to lower angles with increase of H2O in 8 h is sufficient to form ordered mesopores (Fig. S2w). The
the synthesis gel, indicative of an increase in lattice spacing samples show a Type IV isotherm with parallel hysteresis
(Table 1). A longer aging time counter-acts this lattice expansion. loops at P/P1 of around 0.7–0.8, corresponding to a narrow
It is possible that with longer aging time, syneresis takes place, pore size distribution centered at B7.0 nm. Increasing the
resulting in a contraction of the inorganic framework. The wide- aging time from 8 to 24 h led to a decrease in the surface area
angle XRD spectra of the calcined samples (Fig. S1, ESIw) and total pore volume without affecting the pore size.
showed only a broad hump for 2y between 151 to 351, indicating The effect of water on the interaction of the Pluronic
the amorphous structure of the walls.42 No crystalline phase of template with silicate was probed by thermogravimetry.

This journal is c The Royal Society of Chemistry 2012 Catal. Sci. Technol., 2012, 2, 1417–1424 1419
Published on 27 March 2012. Downloaded on 17/03/2017 12:09:01. View Article Online

Fig. 3 Weight loss and DTG profiles of as-synthesized samples of (a)


Si-SBA-15 (b) 10Zr-100-24 (c) 10Zr-208-24 (d) 10Zr-423-24 and (e)
Fig. 2 (a) N2 adsorption–desorption isotherms and (b) pore size 10Zr-639-24.
distribution curves of Zr-SBA-15 samples prepared with different
water content. Isotherms are offset by 300 cm3 g1.
has been attributed to a quick condensation of silicate around
the micelles which reduces the interconnection between the
The as-synthesized zirconium-containing samples lost weight particles.42 The presence of Zr4+ is important as it accelerates
between B200 1C and 550 1C (Fig. 3). The total weight loss is the self-assembly of Pluronic micelles and TEOS. When the
in the order 30–50%. The evolution of CO2 and H2O during amount of water in the synthesis gel is decreased, the platelets
weight loss was verified by TGA-MS (Fig. S3w), hence this step appeared to agglomerate into bigger secondary particles. For
is associated with the decomposition of Pluronic. In contrast, 10ZrMPS, only big agglomerates of B10 mm were observed
decomposition of the purely siliceous SBA-15 occurred at a instead of platelets.
lower temperature, B150–300 1C. This difference in the The extent of zirconium incorporation into the silica frame-
decomposition temperature for Pluronic suggests that the work was found to depend on the water content of the gel and
interaction of the template with the zirconium-containing the aging time. When the amount of water in the synthesis gel
silica surface is stronger than with pure silica. With an was increased at constant HCl volume, more zirconium could
increased amount of water in the synthesis gel, the decom- be incorporated into the silica matrix (Table 2). This can be
position started at lower temperatures, suggesting that the attributed to a decrease in the solubility of the zirconium
interaction between the template and the zirconium species in species as the pH of the gel rose from 0.17 to 0.37. However,
the mesostructure is weakened. This agrees with the pore size the zirconium content in the solids was still lower than
expansion when the water content in the synthesis gel was expected, with Si/Zr between 13.1 to 18.4, instead of Si/Zr
increased. Chen et al.42 observed that samples with low of 10. When the aging time was increased from 8 h to 24 h, the
zirconium content, Si/Zr 4 20, decomposed in two tempera- Si/Zr ratio rose from 12.9 to 17.9. The lower zirconium in
ture regimes, between 180–200 1C and at a higher temperature the 24 h-aged samples showed that under the acidic conditions
of 270–300 1C. However, samples containing more zirconium the rate of dissolution into the synthesis gel is higher than the
(Si/Zr of 16.7 and 10) showed only the higher temperature rate of deposition.
decomposition, indicating a stronger interaction between the The incorporation of zirconium into the silica framework is
surface and Pluronic. supported by FT-IR measurements. The FT-IR spectra of the
The SEM images showed that the Zr-SBA-15 samples samples showed a number of bands indicative of Si–O bonding
formed hexagonal platelets (Fig. 4). With a longer aging time (Fig. 5). In Si-SBA-15, an intense vibration band at ca.
of 24 h, the platelets were bigger in diameter (2–2.5 mm) than 980–1110 cm1 with a shoulder at ca. 1220 cm1 can be assigned
those prepared with 8 h aging, although the thickness was to Si–O–Si asymmetric stretching modes while the smaller
unaffected (B 340 nm) The formation of plate-shaped SBA-15 band at ca. 935 cm1 is due to the Si–OH group vibration.

1420 Catal. Sci. Technol., 2012, 2, 1417–1424 This journal is c The Royal Society of Chemistry 2012
View Article Online

Fig. 5 FTIR of (a) SBA-15 (b) 10Zr-100-24 (c) 10Zr-208-24


(d) 10Zr-423-24 (e) 10Zr-639-24 (f) 10Zr-208-8 and (g) 10Zr-423-8.
Published on 27 March 2012. Downloaded on 17/03/2017 12:09:01.

Fig. 4 SEM images of (a) 10Zr-100-24 (b) 10Zr-208-24 (c) 10Zr-423-24


(d) 10Zr-639-24 (e) 10Zr-423-8 and (f) 10ZrMPS.

Table 2 Zirconium content and acidity of samples

Lewis/Brønstedc
Sample Si/Zr pHa Acidityb (mmol g1)
Solid 25 1C 100 1C
Fig. 6 Ammonia TPD of 10Zr-w-24 and 10ZrMPS samples.
10Zr-100-24 18.4 0.17 0.17 12.5 3.3
10Zr-208-24 17.9 0.33 0.31 11.2 4.3
10Zr-423-24 14.6 0.65 0.33 14.5 4.1 the samples (Fig. S5w). The adsorption of pyridine at Lewis
10Zr-639-24 13.1 0.73 0.43 15.7 6.1 acid sites is indicated by bands at B1440–1450 cm1 and
10Zr-208-8 12.9 0.33 0.51 N.D. N.D. 1600–1610 cm1, while the pyridinium ion formed by adsorp-
10Zr-423-8 16.4 0.65 0.35 N.D. N.D.
10ZrMPS 15.4 — 0.32 N.D. N.D. tion of pyridine at Brønsted acid sites shows a band at
Zr-beta 100 — — 8.3 2.5 B1540–1550 cm1. The peak at B1491 cm1 is attributed to
a
In the synthesis gel. b From NH3 TPD. c From pyridine IR mea- both Brønsted and Lewis acidities. The incorporation of zirconium
surements. N.D: not determined. into the silica framework can result in the generation of Brønsted
acid sites as suggested by Tanabe et al.45 The relative density
of Lewis/Brønsted acid sites were obtained from the bands
These bands broadened and merged together in the zirconium- B1445 cm1 and B1545 cm1 after normalizing with the
containing samples. The smaller bands centered at 800 and respective molar extinction coefficients.46 Lewis acidity was
458 cm1 are assigned to Si–O–Si symmetric stretching and predominant in all the samples (Table 2). The density of Lewis
rocking modes, respectively, and they also broadened with a acid sites increased with the amount of water in the synthesis
slight decrease in intensity when zirconium was introduced gel while samples formed with less water had more Brønsted
into the silica framework. The 29Si MAS spectra show peaks of acidity. Evacuation of the samples at 100 1C resulted in the
Q4 and Q3 (Si with four and three neighbouring Si) with only a removal of pyridine bound to weak Lewis acid sites, so that
small Q2 peak. The higher intensity of the Q3 peak over Q2 the Lewis/Brønsted ratio decreased.
supports the presence of isolated zirconium ions in the silica
matrix (Fig. S4w).
Catalytic activity
The acidity of the samples was assessed by ammonia
temperature-programmed desorption (Fig. 6). Desorption of The catalytic activity of the 10Zr-SBA catalysts was tested in
ammonia occurred between 170 and 470 1C, showing that the the liquid phase Prins condensation of paraformaldehyde and
samples have a range of acidic sites. The density of acid sites b-pinene to form Nopol (Table 3). The samples exhibited good
increased from 0.17 mmol g1 in 10Zr-100-24 to 0.43 mmol g1 activity. Conversions of 54–74% were obtained after 6 h when
for 10Zr-639-24. This is due to the incorporation of more the reaction was carried out at 80 1C. At 100 1C, the conversion
zirconium into the silica framework as the water content in increased to 69–95%. The highest activity was observed for
the synthesis gel increased. The pyridine IR measurements catalysts formed with H2O/Si of 208 and 423. For samples
indicate the presence of both Brønsted and Lewis acid sites in prepared at H2O/Si of 423, the conversion decreased from 74%

This journal is c The Royal Society of Chemistry 2012 Catal. Sci. Technol., 2012, 2, 1417–1424 1421
View Article Online

in 10Zr-423-24 to 60% in 50Zr-423-24. Despite the different


pore sizes of the platelet samples, Nopol was the only product
formed. In contrast, the selectivity to Nopol over 10ZrMPS
was only 67%, because isomerisation of the b-pinene led to
by-products. Using microporous Zr-zeolite beta as the catalyst
also resulted in a fast reaction rate and, after 2 h, the conversion
was already at 77%. However, the selectivity to Nopol was only
46% as limonene and camphene were also formed. The reaction
rate was even higher over the strongly acidic H-Beta (Si/Al 12.5)
but the Nopol selectivity was only 4%.
To optimize the reaction, the reaction temperature was
varied using 10Zr-208-8 as the catalyst (Fig. 7). After 8 h,
the conversion at 60 1C was 64% as compared to 77% when
the reaction was carried out at 80 1C. Surprisingly, when the Fig. 7 Conversion of b-pinene over 10Zr-208-8 at (m) 60 (’) 80 (J)
reaction temperature was increased to 100 1C, the initial rate 100 and (&) 110 1C.
Published on 27 March 2012. Downloaded on 17/03/2017 12:09:01.

was smaller than that at 80 1C, although conversion continued


to increase with time to reach 95% after 8 h. The selectivity (Lewis/Brønsted B 3.3–6.1). Zeolite H-Beta, which has strong
was unaffected by the higher temperature. The lower final Brønsted acidity, gives a low selectivity to Nopol. Similarly,
conversion at temperatures below 80 1C indicates that the our previous investigations into the catalytic properties of
active sites are blocked, most probably by strong adsorption HITQ-2, a MWW zeolite, had found that as a consequence
of the products. A higher temperature of 100 1C can help to of its higher density of Brønsted acid sites (Lewis/Brønsted 1.3),
remove the products that block the active sites, allowing the no Nopol was formed.43 Instead, the products were all
reaction to proceed. mono-isomers of b-pinene. These results agree well with that
However, due to the higher temperature, fewer reactants for the intramolecular Prins reaction where citronellal is
can adsorb at the surface of the catalyst, leading to a lower cyclised to isopulegols.24 There, we showed that catalysts that
initial reaction rate. At an even higher reaction temperature of combined strong Lewis with weak Brønsted acid sites had
110 1C, the conversion reached 95% in 4 h but mostly isomerisa- good activity and selectivity to isopulegols. Strongly acidic
tion products of b-pinene were formed, so that Nopol selectivity catalysts such as Amberlyst, Nafion and sulfated zirconia also
was only 8%. Under these conditions, the surface concentration of gave high conversions but the products were mainly due to
paraformaldehyde is low so that formation of camphene, limonene dehydration, cracking and etherification whereas silica, a weak
and other mono-isomers of b-pinene became predominant. This is acid, gave only very low conversion.24 The coordinatively
supported by thermogravimetric measurements on 10Zr-423-8, unsaturated zirconium ion in Zr-SBA-15 is a much stronger
where the differential thermogravimetric profile shows a high Lewis acid site compared to silicon. Coordination of b-pinene
temperature peak between 160 and 270 1C for the desorption of and formaldehyde to the zirconium brings the molecules close
b-pinene, in addition to desorption of weakly bound b-pinene together (Scheme 2). The platelet morphology exposes high
below 150 1C (Fig. S6w). In contrast, the desorption of numbers of coordinatively unsaturated zirconium ions at edge
formaldehyde occurred below 150 1C. Indeed, the reaction sites, whereas in a granular sample most of the active sites are
order in paraformaldehyde was found to be 1.72 compared to located deep within the channels. The edge sites allow more
0.79 for b-pinene (Fig. S7w). The higher rate dependence on flexibility for the substrate molecules to orient themselves for
paraformaldehyde indicates that its adsorption at the surface intermolecular reaction. Hence, it is envisaged that steric
of the catalyst is weaker than b-pinene. hindrance should not be a major factor (as evidenced from
The high selectivity to Nopol over the Zr-SBA-15 platelets the rate studies) despite the bulky size of b-pinene (Fig. S8w).
may in part be attributed to a unique combination of acid After the molecules are adsorbed and coordinated to the Zr
sites with a predominance of Lewis over Brønsted acid sites center, the reaction is initiated when hydrogen is removed by a
neighbouring oxygen. This is followed by the formation of the
C–C bond. The binding of the two reactants has to be
Table 3 Activity of catalysts for Nopol formation
balanced because too strong a binding of either one will
Entry Catalyst Conv.a (%) Nopol sel. (%) Initial TOF (h1) adversely affect the rate and selectivity of the reaction. The
1 10Zr-100-24 63 100 23
lower reaction order in b-pinene indicates that it adsorbs
2 10Zr-208-24 72 100 19 stronger than formaldehyde at the surface of the catalyst. In
3 10Zr-423-24 74 100 16 the presence of Brønsted acid sites, isomerisation leads to side
4 10Zr-639-24 54 100 13 products such as camphene and limonene. However, the
5 10Zr-208-8 74 100 18
6 10Zr-423-8 68 100 16 strength and density of acidic sites for 10ZrMPS is very similar
7 50Zr-423-24 64 100 42 to that of the 10Zr-SBA-15 platelets and therefore unlikely to
8 10ZrMPS 95 67 37 be solely responsible for the low Nopol selectivity. Moreover,
9 Zr-betab 77 46 163
the pores of 10ZrMPS are in the order of B3–8 nm, and,
10 H-betab 100 4 29
a
although not narrowly distributed, they are within the pore
After 6 h. b After 2 h. Reaction conditions: 1 mmol b-pinene, 2 mmol
size range of the Zr-SBA-15 platelets. Hence, a plausible
paraformaldehyde, 5 ml toluene, 50 mg catalyst, 80 1C.
explanation for the lower Nopol selectivity over 10ZrMPS

1422 Catal. Sci. Technol., 2012, 2, 1417–1424 This journal is c The Royal Society of Chemistry 2012
View Article Online

that are narrowly distributed about a mean pore size. In


contrast, the sample prepared using a H2O/Si ratio of 100
had a broader pore size distribution. The mean pore diameter
varied from 4 to 8 nm, depending on the amount of water in
the synthesis gel. The expansion in pore size with increased
water in the synthesis gel was attributed to a reduced inter-
action of the Pluronic template with the inorganic surface. The
Si/Zr in the calcined oxides decreased from 18.4 to 13.1 when
the H2O/Si ratio was increased from 100 to 639. The increased
zirconium content in the samples resulted in a higher density
of Brønsted and Lewis acid sites, with a predominance of the
latter. The Zr-SBA-15 platelets showed good catalytic activity
in the Prins condensation of b-pinene and paraformaldehyde.
Scheme 2 Proposed mechanism for the intermolecular Prins reaction The selectivity to Nopol was essentially 100%. In comparison,
and isomerisation of b-pinene. the selectivity over particulate 10ZrMPS, microporous Zr-beta
Published on 27 March 2012. Downloaded on 17/03/2017 12:09:01.

and H-beta was lower. The catalysts could be regenerated for


further batch reactions following treatment with H2O2.

Acknowledgements
Financial support from National University of Singapore under
grant number R-143-000-418-112 is gratefully acknowledged.

Notes and references


1 K. Wilson and J. H. Clark, Pure Appl. Chem., 2000, 72, 1313–1319.
2 B. B. Snider, in Comprehensive Organic Synthesis, ed. B. M. Trost,
Fig. 8 Reuse of catalyst 10Zr-208-8 (’) conversion after 8 h (&)
I. Fleming and C. H. Heathcock, Pergamon, Oxford, 1991, vol. 2,
selectivity. pp. 527–561.
3 M. L. Clarke and M. B. France, Tetrahedron, 2008, 64, 9003–9031.
may lie in its big particle sizes, where in diffusing through the 4 I. M. Pastor and M. Yus, Curr. Org. Chem., 2007, 11, 925–957.
long tortuous channels, b-pinene encounters several acidic 5 J. T. Williams, P. S. Bahia and J. S. Snaith, Org. Lett., 2002, 4,
sites for isomerisation. Similarly, the higher density of 3727–3730.
6 B. B. Snider and G. B. Phillips, J. Org. Chem., 1983, 48, 464–469.
Brønsted acid sites in Zr-beta and H-beta compared to the 7 L. M. Stephenson and M. Orfanopoulos, J. Org. Chem., 1981, 46,
10Zr-SBA-15 samples and the diffusion limitation posed by 2200–2201.
the microporous nature of zeolites could contribute to 8 H. Kwart and M. Brechbiel, J. Org. Chem., 1982, 47, 5409–5411.
9 N. H. Andersen, S. W. Hadley, J. D. Kelly and E. R. Bacon,
the lower Nopol selectivity. The short mesoporous channels
J. Org. Chem., 1985, 50, 4144–4151.
in Zr-SBA-15 platelets offer easy access to the active sites 10 J. S. Yadav, B. V. S. Reddy and G. Bhaishya, Green Chem., 2003,
leading to improved activity and utilization of the catalyst. 5, 264–266.
11 G. X. Li, Y. L. Gu, Y. Ding, H. P. Zhang, J. M. Wang, Q. Gao,
Reuse of the catalyst L. Yan and J. S. Suo, J. Mol. Catal. A: Chem., 2004, 218, 147–152.
12 H. Surburg and J. Panten, Common Fragrance and Flavor Materials,
The recovery and reusability of the catalysts were investigated. Preparation, Properties and Uses, Wiley-VCH, Weinheim, Germany,
After the reaction, the catalyst was washed with toluene, 2006, pp. 67–68.
13 J. P. Bain, J. Am. Chem. Soc., 1946, 68, 638–641.
centrifuged and dried in the oven for the next run. The results 14 A. L. Villa de P, E. Alarcón and C. M. de Correa, Chem. Commun.,
showed that the activity was lower. It would appear that some 2002, 2654–2655.
active sites remained blocked by the residual products. Treating 15 A. L. Villa de P, E. Alarcón and C. M. de C, Catal. Today, 2005,
the used catalyst with H2O2 solution at 40 1C overnight resulted in 107–108, 942–948.
16 M. Selvaraj and P. K. Sinha, New J. Chem., 2010, 34, 1921–1929.
a rejuvenated catalyst with activity 4 95% conversion and 100% 17 E. A. Alarcón, L. Correa, C. Montes and A. L. Villa, Microporous
selectivity to Nopol (Fig. 8). With H2O2, fouling compounds that Mesoporous Mater., 2010, 136, 59–67.
cannot be removed by washing with toluene can be oxidized at low 18 M. Selvaraj and Y. Choe, Appl. Catal., A, 2010, 373, 186–191.
19 A. Corma and M. Renz, Arkivoc, 2007, viii, 40–48.
temperatures. The differential thermogravimetric profile of the 20 U. R. Pillai and E. Sahle-Demessie, Chem. Commun., 2004, 826–827.
used catalyst confirmed that, despite washing with toluene, small 21 A. M. Selvaraj and S. Kawi, J. Mol. Catal. A: Chem., 2006, 246,
amounts of strongly adsorbed residues were still present and could 218–222.
only be removed at temperatures of up to 500 1C (Fig. S9w). 22 M. K. Yadav and R. V. Jasra, Catal. Commun., 2006, 7, 889–895.
23 M. V. Patil, M. K. Yadav and R. V. Jasra, J. Mol. Catal. A: Chem.,
2007, 273, 39–47.
Conclusions 24 G. K. Chuah, S. H. Liu, S. Jaenicke and L. J. Harrison, J. Catal.,
2001, 200, 352–359.
The amount of water present during the synthesis of Zr-SBA-15 25 Y. Z. Zhu, Y. T. Nie, S. Jaenicke and G. K. Chuah, J. Catal., 2005,
229, 404–413.
affects the dimension and distribution of pores. Samples prepared 26 S. Telalović, J. F. Ng, R. Maheswari, A. Ramanathan,
with a H2O/Si ratio of 208–639 formed platelets with pores G. K. Chuah and U. Hanefeld, Chem. Commun., 2008, 4631–4633.

This journal is c The Royal Society of Chemistry 2012 Catal. Sci. Technol., 2012, 2, 1417–1424 1423
View Article Online

27 S. Telalović, A. Ramanathan, J. F. Ng, R. Maheswari, C. Kwakernaak, 37 C. L. Chen, S. F. Cheng, H. P. Lin, S. T. Wong and C. Y. Mou,
F. Soulimani, H. C. Brouwer, G. K. Chuah, B. M. Weckhuysen and Appl. Catal., A, 2001, 215, 21–30.
U. Hanefeld, Chem.–Eur. J., 2011, 17, 2077–2088. 38 B. L. Newalkar, J. Olanrewaju and S. Komarneni, J. Phys. Chem.
28 M. Morandin, R. Gavagnin, F. Pinna and G. Strukul, J. Catal., B, 2001, 105, 8356–8360.
2002, 212, 193–200. 39 K. Szczodrowski, B. Prelot, S. Lantenois, J. Zajac, M. Lindheimer,
29 M. S. Wong, H. C. Huang and J. Y. Ying, Chem. Mater., 2002, 14, D. Jones, A. Julbe and A. van der Lee, Microporous Mesoporous
1961–1973. Mater., 2008, 110, 111–118.
30 Y. Z. Zhu, S. Jaenicke and G. K. Chuah, J. Catal., 2003, 218, 40 Y. C. Du, S. Liu, Y. L. Zhang, F. Nawaz, Y. Y. Ji and F. S. Xiao,
396–404. Microporous Mesoporous Mater., 2009, 121, 185–193.
31 S. Y. Chen, J. F. Lee and S. Cheng, J. Catal., 2010, 270, 41 S. Y. Chen, L. Y. Jang and S. Cheng, Chem. Mater., 2004, 16,
196–205. 4174–4180.
32 A. Ramanathan, M. C. C. Villalobos, C. Kwakernaak, S. Telalović 42 S. Y. Chen, C. Y. Tang, W. T. Chuang, J. J. Lee, Y. L. Tsai, J. C. C.
and U. Hanefeld, Chem.–Eur. J., 2008, 14, 961–972. Chan, C. Y. Lin, Y. C. Liu and S. Cheng, Chem. Mater., 2008, 20,
33 D. Barreca, M. P. Copley, A. E. Graham, J. D. Holmes, 3906–3916.
M. A. Morris, R. Seraglia, T. R. Spalding and E. Tondello, Appl. 43 J. Wang, S. Jaenicke, G. K. Chuah, W. M. Hua, Y. H. Yue and
Catal., A, 2006, 304, 14–20. Z. Gao, Catal. Commun., 2011, 12, 1131–1135.
34 O. Y. Gutierrez, G. A. Fuentes, C. Salcedo and T. Klimova, Catal. 44 A. Vinu, V. Murugesan, W. Bolhlmann and M. Hartmann,
Today, 2006, 116, 485–497. J. Phys. Chem. B, 2004, 108, 11496–11505.
35 W. M. Hua, Y. H. Yue and Z. Gao, J. Mol. Catal. A: Chem., 2001, 45 K. Tanabe, M. Misono, Y. Ono and H. Hattori, New Solid Acids and
Published on 27 March 2012. Downloaded on 17/03/2017 12:09:01.

170, 195–202. Bases, Their Catalytic Properties, in Studies in Surface Science and
36 H. Matsuhashi, M. Tanaka, H. Nakamura and K. Arata, Appl. Catalysis 51, ed. B. Delmon and J. T. Yates, Kodansha, Tokyo, 1989.
Catal., A, 2001, 208, 1–5. 46 C. A. Emeis, J. Catal., 1993, 141, 347–354.

1424 Catal. Sci. Technol., 2012, 2, 1417–1424 This journal is c The Royal Society of Chemistry 2012

You might also like