Download as pdf or txt
Download as pdf or txt
You are on page 1of 8

Journal of Molecular Catalysis A: Chemical 423 (2016) 248–255

Contents lists available at ScienceDirect

Journal of Molecular Catalysis A: Chemical


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

Editor’s choice paper

Photo-epoxidation of cyclohexene, cyclooctene and 1-octene with


molecular oxygen catalyzed by dichloro
dioxo-(4,4 -dicarboxylato-2,2 -bipyridine) molybdenum(VI) grafted on
mesoporous TiO2
Henry Martínez a,∗ , María F. Cáceres b , Fernando Martínez a , Edgar A. Páez-Mozo a ,
Sabine Valange c,∗ , Nelson J. Castellanos d , Daniel Molina e , Joël Barrault c ,
Henri Arzoumanian f
a
Centro de Investigaciones en Catálisis-CICAT, Universidad Industrial de Santander, Escuela de Química, Km 2 vía El Refugio, Piedecuesta, Santander,
Colombia
b
Laboratório de Química Atmosférica (LQA), Universidade Católica do Río de Janeiro, Rua Marquês de Sao Vicente, 225, Gávea-Rio de Janeiro, RJ, Brazil
c
Institut de Chimie des Milieux et Matériaux de Poitiers (IC2MP), Université de Poitiers, CNRS, ENSIP, B1, 1 rue M. Doré, F-86073 Poitiers, France
d
Estado Sólido y Catálisis Ambiental (ESCA), Universidad Nacional de Colombia, Carrera 30 No. 45-03, Bogotá, Colombia
e
Laboratorio de Resonancia Magnética Nuclear, Universidad Industrial de Santander, Km 2 vía El Refugio, Piedecuesta, Colombia
f
Chirosciences UMR CNRS 7313 Institut des Sciences Moléculaires de Marseille (iSm2), Aix-Marseille Université, Campus St Jérôme, 13397 Marseille, France

a r t i c l e i n f o a b s t r a c t

Article history: The photo-assisted oxygen transfer to cyclohexene, cyclooctene and 1-octene was investigated
Received 12 November 2015 using molecular O2 as a primary oxidant with a dichloro-dioxo-(4,4 -dicarboxylato-2,2 -bipyridine)
Received in revised form 31 May 2016 molybdenum(VI) complex grafted on the surface of mesoporous TiO2 supports exhibiting different tex-
Accepted 4 July 2016
tural properties. Such mesoporous titania samples were prepared following a hydrothermal treatment
Available online 6 July 2016
(HT) and through an original procedure under supercritical CO2 (SC) conditions. Non porous TiO2 Degussa
P-25 was also used for comparison purposes. The porosity of the titania supports, as well as the amount
Keywords:
and dispersion of the covalently anchored dioxo Mo(VI) complex were shown to strongly influence the
Dioxomolybdenum complex
Mesoporous TiO2
resulting photo-catalytic properties. A particular attention was also devoted to the stability of the grafted
Photocatalysis catalysts by performing long-term epoxidation experiments. The new hybrid Mo(VI) -derivatized meso-
Selective oxidation porous SC TiO2 catalyst proved to be the most efficient in the photo-assisted oxygen transfer to various
Epoxidation alkenes. Additionally, the high stability of such catalyst was ascertained by recycling studies.
© 2016 Elsevier B.V. All rights reserved.

1. Introduction chemicals, pharmaceuticals, agrochemical products, epoxy paints


and dyes [6,7]. However, the production of epoxide compounds is
Traditional oxidation reactions based on stoichiometric oxi- still a big issue because several side reactions can take place, such
dants tend to produce waste side products generating a negative as oxidation in the allylic positions, ring opening of the epoxides
environmental impact [1]. In this respect, the use of environmen- or complete oxidation to CO2 [8]. Furthermore, the epoxidation of
tally friendly oxidants such as molecular oxygen is a more attractive terminal olefins such as 1-octene is difficult and therefore requires
and challenging option in order to generate oxygenates without long time reaction [9]. Traditionally, epoxides are prepared by
formation of by-products [2–4]. partial oxidation of alkenes using “exotic” oxidants like iodosyl-
Epoxidation of alkenes is a suitable route prime of importance benzene, hypochlorite or organic hydroperoxides [10]. Hydrogen
for the preparation of relevant industrial intermediates [5]. Epox- peroxide or t-butyl hydroperoxide (TBHP) are known as efficient
ides are versatile and reactive compounds for the synthesis of fine oxidants [11] with the assistance of homogeneous transition metal-
based catalysts [12–14]. The epoxidation of ethylene and propylene
was recently reported with molecular oxygen or air as a primary
∗ Corresponding authors. oxidant with a heterogeneous silver catalyst [15,16]. However, the
E-mail addresses: henry.martinez1@correo.uis.edu.co (H. Martínez), reaction conditions are rather severe and risky. Thus, the develop-
sabine.valange@univ-poitiers.fr (S. Valange). ment of catalytic systems that use molecular oxygen or air in mild

http://dx.doi.org/10.1016/j.molcata.2016.07.006
1381-1169/© 2016 Elsevier B.V. All rights reserved.
H. Martínez et al. / Journal of Molecular Catalysis A: Chemical 423 (2016) 248–255 249

conditions remains a challenging target [17]. In previous studies excess of precursor were separated from the reaction medium dur-
[18–21], we have reported on the selective production of organic ing the CO2 expansion. The TiO2 sample synthesized under SC CO2
oxygenated compounds by Oxygen Atom Transfer (OAT) reactions, was referred to as TiO2 (SC-150).
using O2 as the primary oxidant, with a dioxoMo(VI) complex
grafted on a commercial TiO2 (MoO2 L/TiO2 ) and UV–vis irradiation.
Here, the photo-assisted oxygen transfer to cyclohexene, 2.2.3. Mo(VI) Cl2 O2 Bipy grafting procedure
cyclooctene and 1-octene was investigated on new grounds with The covalent grafting of the dichloro-dioxo-(4,4 -dicarboxylato-
a dichloro-dioxo-(4,4 -dicarboxylato-2,2 -bipyridine)-Mo(VI) com- 2,2 -bipyridine)-Mo(VI) complex on the surface of both dehydrated
plex (labelled as Mo(VI) Cl2 O2 Bipy) grafted on the surface of various titania supports was carried out according to the following
mesoporous titania powders. Two different TiO2 synthesis routes procedure. Hexamethyldisilazane was added to a benzene sus-
were explored with the aim of highlighting the influence of the pension of TiO2 , stirred 24 h at room temperature, filtered,
porosity on the dispersion of the Mo(VI) Cl2 O2 Bipy complex, and washed thoroughly with benzene and dried under vacuum. A
hence, on the resulting photo-catalytic properties. High surface benzene suspension of freshly prepared trimethylsilylated TiO2
area mesoporous TiO2 supports with anatase walls were prepared was then added to a benzene suspension of 2,2 -bipyridyl-4,4 -
following a CTAB-assisted hydrothermal process, as well as by a dicarboxylic acid, stirred 72 h at room temperature, filtered
sol-gel route under supercritical carbon dioxide (SC CO2 ). As a non- and thoroughly washed with benzene. Finally, a THF suspen-
toxic and environmentally benign green solvent, SC CO2 exhibits sion of MoO2 Cl2 was added to a benzene suspension of the
many unique features and great versatility in materials process- previously obtained solid, stirred 4 h at room temperature and
ing and synthesis. Such characteristics allow for a better control of the excess of solvent evaporated. The solid product was then
the nanoarchitecture and the textural properties of the resulting washed with acetonitrile and dried at room temperature [20]. The
materials [22]. dioxo Mo(VI) grafted titania samples were designated as follows:
Several complementary techniques, including XRD, nitrogen Mo(VI) Cl2 O2 Bipy/TiO2 (HT-400), Mo(VI) Cl2 O2 Bipy/TiO2 (SC-150) and
sorption isotherm, Raman, ATR-FTIR, UV–vis and 13 C CP-MAS NMR Mo(VI) Cl2 O2 Bipy/TiO2 (P-25), where the label in brackets stands for
spectroscopies were used to characterize the parent mesoporous the titania support prepared under hydrothermal treatment, super-
TiO2 supports and the Mo(VI) Cl2 O2 Bipy grafted counterparts. critical conditions and the reference Degussa TiO2 , respectively.
Finally, the stability of these hybrid catalysts was evaluated after
long-term photo-epoxidation experiments (56 h).
2.3. Catalysts characterization

2. Experimental
The TiO2 supports and the grafted catalysts were character-
ized by powder X-ray diffraction (XRD) using a Bruker AXS D8
2.1. Materials
Advance DaVinci geometry with monochromatized Cu K␣ radia-
tion (␭ = 1.5418 Å) at 40 kV and 30 mA. The diffraction patterns were
All chemical reagents were purchased and used without fur-
recorded in the 2 value range of 20–70◦ (with a step size of 0.01◦
ther purification unless otherwise specified. All solvents were
and a step time of 0.4 s) for wide-angle analysis and at 2 = 2–10◦
thoroughly degassed prior to use. Acetonitrile was distilled and
for low-angle analysis. Nitrogen adsorption/desorption isotherms
kept under argon. Cetyltrimethylammonium bromide (CTAB) and
at 77 K were measured on a Micromeritics Tristar 3000 apparatus
HCl (37%) were purchased from Acros, while ethanol (absolute
at −196 ◦ C. Before analysis, the samples were degassed under vac-
for analysis) and NaOH were obtained from Aldrich. Titanium
uum at 150 ◦ C for 12 h. The specific surface area was determined
oxide (Degussa-P25) with a specific area of 50 m2 g−1 and a crys-
from the linear part (0–0.23 P/Po ) of the BET plot. The total pore
talline structure composed of 80% anatase and 20% rutile was
volume was measured from the isotherms at P/Po = 0.95 and the
used for comparison. All supports were dehydrated at 60 ◦ C under
mean pore diameter was determined by the BJH method applied
5 × 10−4 mbar for 48 h prior to the anchoring procedure.
to the adsorption branch. Raman analysis of both supports and
catalysts was carried out on a LabRAM HR800UV HORIBA Jobin
2.2. Catalysts synthesis Yvon spectrometer with a BXFM microscope, equipped with a CCD
detector and an internal HeNe laser (632.81 nm) and an external
2.2.1. CTAB-assisted hydrothermal route Ar laser (514.532 nm). All grafted Mo(VI) Cl2 O2 Bipy/TiO2 catalysts
The hydrothermal preparation route of the mesoporous TiO2 were characterized by attenuated total reflectance (ATR FT-IR)
was based on the following procedure. An aqueous solution of tita- spectroscopy on a Perkin Elmer spectrometer and 13 C CP-MAS
nium sulphate (1.1402 g) was first prepared and added to a CTAB NMR spectroscopy on a Bruker Avance-400 MHz NMR spectrome-
ethanolic solution under stirring. The molar ratio of Ti(SO4 )2 :CTAB ter equipped with a 4 mm magic angle spinning (MAS) probe head
is 1:0.12. The resulting mixture was then aged at room tempera- for solids. The CP technique was applied during Magic Angle Spin-
ture for 12 h before being transferred to an autoclave at 100 ◦ C for ning (MAS) of the rotor at 10 kHz. A ramped 1H-pulse starting at
72 h. After the hydrothermal treatment, the solid was recovered by 100% power and decreasing until 50% was used during contact
centrifugation, washed with water and ethanol, dried at 120 ◦ C and time (2 ms) in order to circumvent Hartmann–Hahn mismatches.
then calcined at 400 ◦ C in air for 6 h with a heating rate of 2 ◦ C min−1 In order to obtain a good signal-to-noise ratio in 13 C CPMAS exper-
[23]. The TiO2 sample was labelled as TiO2 (HT-400). iment, 2048 scans were accumulated using a delay of 2 s. The
13 C chemical shifts were referenced to tetramethylsilane and cal-

2.2.2. SC CO2 -assisted sol-gel route ibrated with the adamantane signal set at 38.48 ppm [24]. Diffuse
This procedure consists in the elaboration of mesoporous TiO2 reflectance spectra (200–700 nm) were measured on a Cary 5000
by a sol-gel route under supercritical CO2 through hydrolysis- Varian spectrophotometer. The amount of grafted complex was
condensation of titanium tetraisopropoxide at 150 ◦ C under determined by thermo gravimetric analyses (TGA) under N2 atmo-
30 MPa. Briefly, titanium tetraisopropoxide was introduced into the sphere between 30 and 900 ◦ C with a 10 ◦ C min−1 heating ramp
reactor prior to the injection of CO2 . As soon as the desired reac- using a SDT-Q600 apparatus. Molybdenum elemental analysis was
tion was reached, an excess of water was added under supercritical carried out with a Thermo S4 atomic absorption spectrophotometer
conditions. The alcohol formed through hydrolysis as well as the (AAS), after acid digestion of the samples.
250 H. Martínez et al. / Journal of Molecular Catalysis A: Chemical 423 (2016) 248–255

Fig. 1. Wide and small-angle (inset) XRD patterns obtained for (a) TiO2 (HT-400), (b) Fig. 2. N2 adsorption-desorption isotherms for the mesoporous (a) TiO2 (SC-150)
Mo(VI) Cl2 O2 Bipy/TiO2 (HT-400), (c) TiO2 (SC-150) and (d) Mo(VI) Cl2 O2 Bipy/TiO2 (SC- and (b) TiO2 (HT-400). Insets showed the corresponding pore size distribution
150). obtained from the adsorption loop.

2.4. Photocatalytic experiments


XRD patterns (inset). The diffractograms were similar and exhib-
The photocatalytic-assisted alkene epoxidation was carried out ited a sharp diffraction peak located at 2 = 3◦ . The TiO2 sample
by using a 15 mL batch microreactor (ACEGLASS) equipped with synthesized through the SC CO2 -assisted sol-gel route displayed
a mercury lamp (UV PenRay, ␭ ≥ 360 nm) at a constant temper- however a broadening coupled to a reduction in the intensity of
ature of 19 ◦ C. All substrates dissolved in CH3 CN (10−2 mol L−1 ) this low-angle peak. After the Mo(VI) Cl2 O2 Bipy grafting procedure,
were thoroughly deoxygenated by bubbling N2 for several hours the samples still exhibited a single strong XRD peak indicating that
before the addition of the catalyst (15 mg). The reaction system was the TiO2 mesoporous structure was maintained. Moreover no shift
allowed to equilibrate during 1 h under O2 atmosphere in the dark. of this diffraction peak was observed after the functionalization
The photo-reaction was performed for 5 h under O2 and UV–vis step, thereby confirming that no change occurred in the lattice
light. Then the reaction system was maintained in the dark with parameter and in the wall thickness of the titania supports.
O2 during 12 h. The illumination-darkness cycle was repeated four The N2 adsorption-desorption isotherms of the titania supports
more times. In order to evaluate the stability of the catalyst, the and their corresponding pore size distributions are shown in Fig. 2.
powder was separated from the reaction medium at the end of the According to the IUPAC classification, these isotherms are of type
photo-oxidation process and washed with acetonitrile. Then the IV and are attributed to mesoporous materials. The shape of the
recovered catalyst was introduced into a new CH3 CN-based olefin isotherms also indicated that the TiO2 (HT-400) sample contains
solution prior to the photo-epoxidation reaction. The illumination- larger mesopores than the support prepared under supercritical
darkness cycle was repeated five times. Long term photo-oxidation conditions. A wider pore size distribution was indeed observed
reactions were also carried out under O2 atmosphere and UV–vis for the TiO2 (HT-400) titania (Fig. 2b) compared with that of the
illumination during 35 h and 56 h. The catalyst reuse was evaluated TiO2 (SC-150) sample (Fig. 2a). The specific surface areas, mean
by performing six consecutive oxidation runs with the same cata- pore diameters and pore volumes of the TiO2 samples before and
lyst, each time using a fresh alkene solution. All photo-catalytic after the Mo(VI) Cl2 O2 Bipy grafting procedure are summarized in
experiments were performed with the MoCl2 O2 Bipy/TiO2 (P-25) Table 1. The corresponding data for the reference P-25 TiO2 and
catalyst used for comparison purposes. the Mo(VI) Cl2 O2 Bipy/TiO2 (P-25) counterpart are also provided for
The reaction products were analysed by using a gas chromatog- comparison. It appeared that the TiO2 support prepared by the sol-
raphy (GC-HP-6890) equipped with a flame ionization detector and gel route under supercritical CO2 exhibited a higher BET surface
a HP-INNOWAX column (30 m × 0.32 mm × 0.25 ␮m). The quantifi- area coupled to a higher pore volume with respect to the meso-
cation of the products was performed using benzene as internal porous titania obtained after hydrothermal treatment. As expected
standard. after the dioxo-Mo(VI) complex immobilization step, a decrease of
adsorbed N2 , surface area, pore volume and of the pore diameter
3. Results and discussion was observed for all grafted samples when compared to the parent
TiO2 supports.
3.1. Catalysts characterization Fig. 3 shows the Raman spectra of bare TiO2 supports and
their Mo(VI) Cl2 O2 Bipy/TiO2 analogous systems. Both titania sup-
Wide-angle X-ray diffraction analyses indicated that anatase ports exhibited the characteristic peaks of the anatase phase at
is the main phase for both HT-400 and SC-150 TiO2 supports 145 cm−1 (symmetric stretching vibration of O Ti O), 390 cm−1
(JCPDS No. 21-1272), along with traces of rutile in the case of the (symmetric bending vibration of O Ti O), 509 cm−1 (antisymmet-
mesoporous titania prepared under hydrothermal heating. Fig. 1 ric bending vibration of O-Ti-O) and 632 cm−1 , corresponding to
confirmed the presence of the (101), (004), (200), (105) and (211) the Eg(1) , B1g(1) , (A1g + B1g(2) ) and Eg(2) modes, respectively [26,27].
diffraction peaks characteristics of anatase [25] for the parent For the Mo(VI) Cl2 O2 Bipy/TiO2 (HT and SC) systems, all the vibra-
supports, as well as for both TiO2 grafted with the dichloro-dioxo- tion bands for anatase can still be found (inset). Additionally, the
(4,4 -dicarboxylato-2,2 -bipyridine)-Mo(VI) complex. The thermal Raman spectra of the grafted samples showed the vibrational bands
stability of the TiO2 mesostructures prepared under hydrothermal related to the symmetric stretching of the bipyridyl ligand at 1631,
or supercritical conditions were clearly evidenced by the low-angle 1561, 1505, 1429 and 1301 cm−1 , as well as the ring breathing
H. Martínez et al. / Journal of Molecular Catalysis A: Chemical 423 (2016) 248–255 251

Table 1
Textural properties, density of TiO2 surface hydroxyl groups and content of grafted MoCl2 O2 Bipy complex for the titania supports before and after immobilization.

Solid S BET (m2 g−1 ) Vp (cm3 g−1 ) Mean Øp (nm) OH (mmol g−1 )a MoCl2 O2 Bipy (mmol g−1 )b

TiO2 (P-25) 50 0.08 – 0.37 –


Mo(VI) Cl2 O2 Bipy/TiO2 (P-25) 47 0.07 – 0.08 0.14
TiO2 (HT-400) 92 0.26 11.3 1.31 –
Mo(VI) Cl2 O2 Bipy/TiO2 (HT-400) 84 0.20 10.2 0.92 0.19
TiO2 (SC-150) 150 0.32 7.5 3.51 –
Mo(VI) Cl2 O2 Bipy/TiO2 (SC-150) 126 0.18 6.1 2.53 0.48
a
Determined by TGA.
b
Mo(VI) -dioxo complex concentration determined by Mo elemental analysis.

Fig. 3. Raman spectra of the mesoporous anatase TiO2 supports and the correspond- Fig. 4. Total reflectance (ATR-FTIR) spectra of a) reference TiO2 (P-25),
ing MoCl2 O2 Bipy/TiO2 systems. (b) Mo(VI) Cl2 O2 Bipy/TiO2 (P-25), c) Mo(VI) Cl2 O2 Bipy/TiO2 (HT-400) and d)
Mo(VI) Cl2 O2 Bipy/TiO2 (SC-150).

at 1025 cm−1 . The characteristic Mo = O symmetric vibration was


observed at 949 cm−1 [12]. All these bands confirmed the presence
of the Mo(VI) Cl2 O2 Bipy moiety in both TiO2 supports. Moreover we assignments are provided in Table S1 (ESI). The asymmetric and
can also observed that the Raman band intensity of the immobi- symmetric stretching vibrations of the cis MoO2 group were
lized dioxo-Mo(VI) complex was higher for TiO2 (SC-150) than for observed at 940 cm−1 and 918 cm−1 , respectively (Fig. 4) [18].
TiO2 (HT-400), thereby suggesting that the titania prepared through The stretching vibration bands related to the carboxylate lig-
the supercritical route contained an increased amount of grafted and were observed at 1717 cm−1 (C O), 1271 cm−1 (asymmetric
Mo(VI) Cl2 O2 Bipy than that of the hydrothermally heated one. COO− ) and 1125 cm−1 (symmetric COO− ) [19]. The Ti COO− link-
The amount of the dioxomolybdenum(VI) complex covalently age was identified by the 1396 cm−1 asymmetric and 1368 cm−1
anchored on all TiO2 supports was determined by Mo elemental symmetric scissor vibration modes [20,28]. The IR spectra of
analysis (Table 1). The use of atomic absorption spectrophotom- the Mo(VI) Cl2 O2 Bipy/TiO2 samples clearly showed that the dioxo
etry was indeed preferred in order to get an accurate value of molybdenum complex was covalently bonded to the mesoporous
the grafted Mo-dioxo active centers. The determination of the TiO2 supports.
loading of the Mo(VI) complex through C, H and N elemental anal- Diffuse reflectance UV–vis analysis provided additional evi-
ysis before and after the grafting procedure was also performed dence for the successful anchoring of the MoCl2 O2 Bipy complex
but proved to be not well suited for a precise evaluation of the on the surface of the two mesoporous titania supports and of the
Mo content. The percentage of carbon, hydrogen and nitrogen reference P-25 TiO2 . The comparison of the DR UV–vis spectra of
determined for all Mo(VI) Cl2 O2 Bipy/TiO2 samples was systemati- the titania before and after the dioxo-MoVI immobilization indi-
cally higher than that of the stoichiometric amount corresponding cated a new absorption band at ca. 560 nm (Fig. 5), related to the
to the dichloro-dioxo-(4,4 -dicarboxylato-2,2 -bipyridine)-Mo(VI) presence of the Mo = O group (n → ␲* transition) [21]. The inten-
complex, suggesting the presence of solvent molecules trapped sity of this band was higher for the TiO2 (SC-150) sample than for
within the pores of the TiO2 supports. TiO2 (HT-400), in accordance with the higher amount of complex
Important differences were observed regarding the concentra- immobilized on the surface of the TiO2 synthesized under super-
tion of the grafted Mo(VI) complex for all catalysts. While this critical CO2 (Table 1).
amount was in the same range for the TiO2 (HT-400) and P-25 sup- Fig. 6 shows the 13 C NMR spectra of the mesoporous
ports, the mesoporous TiO2 synthesized under supercritical CO2 and P-25 Mo(VI) Cl2 O2 Bipy/TiO2 samples. The occurrence of the
was shown to exhibit a far increased content of covalently anchored C( O) O Ti4+ group was observed at ␦ ≈ 167 ppm, while the peaks
dioxo-Mo(VI) complex as high as 0.43 mmol g−1 (Table 1). Such dif- corresponding to the pyridinic ligand [29] were detected between
ference could be due to the higher surface density of hydroxyl 160 and 120 ppm.
groups exhibited by the TiO2 (SC-150) support with respect to the All the spectroscopic characterizations of the functional-
titania prepared under hydrothermal treatment (TiO2 (HT-400)). ized samples clearly showed that the dichloro-dioxo-(4,4 -
The anchoring of the Mo(VI) Cl2 O2 Bipy complex on both meso- dicarboxylato-2,2 -bipyridine) molybdenum(VI) complex was duly
porous TiO2 was further confirmed by FTIR and the band immobilized on the surface of the mesoporous titania supports.
252 H. Martínez et al. / Journal of Molecular Catalysis A: Chemical 423 (2016) 248–255

Fig. 7. Cyclohexene photo-conversion over the various Mo(VI) O2 Cl2 Bipy/TiO2 sys-
Fig. 5. UV–vis diffuse reflectance spectra of (a) Mo(VI) Cl2 O2 Bipy/TiO2 (P-25) (b) tems.
Mo(VI) Cl2 O2 Bipy/TiO2 (HT-400), (c) Mo(VI) Cl2 O2 Bipy/TiO2 (SC-150) and of the corre-
sponding TiO2 supports (grey lines).
Table 2
Cyclohexene photo-oxidation with various fresh and used Mo(VI) Cl2 O2 Bipy/TiO2
catalysts. Yield to cyclohexene oxide and selectivities to cyclohexene oxide, cyclo-
hexenone and 2-cyclohexen-1-ol.

Fresha and usedb catalysts Yield (%)c Selectivity (%)

Mo(VI) Cl2 O2 Bipy/TiO2 P-25 fresh 50.6 70 17 13


Mo(VI) Cl2 O2 Bipy/TiO2 P-25 used 46.2 68 19 13
Mo(VI) Cl2 O2 Bipy/TiO2 HT-400 fresh 51.8 74 15 11
Mo(VI) Cl2 O2 Bipy/TiO2 HT-400 used 41.9 70 18 12
Mo(VI) Cl2 O2 Bipy/TiO2 SC-150 fresh 54.7 83 10 7
Mo(VI) Cl2 O2 Bipy/TiO2 SC-150 used 50.6 79 13 8
a
Measured after 56 h of reaction.
b
Fresh: new catalyst after 4 reaction cycles (56 h).
c
Used: catalyst filtered after 4 reaction cycles (56 h), then employed with a new
reaction solution during 4 reaction cycles (56 h).

Fig. 6. 13 C CP-MAS NMR spectra recorded at 10 kHz of (a) Mo(VI) Cl2 O2 Bipy/TiO2 (P- epoxide product was always higher than 70%, thereby indicat-
25) (b) Mo(VI) Cl2 O2 Bipy/TiO2 (HT-400) and (c) Mo(VI) Cl2 O2 Bipy/TiO2 (SC-150). ing that the OAT is the predominant process. An effect of the
TiO2 porosity and hence on the dispersion of the dioxo-Mo(VI)
3.2. Photo-catalytic oxidation of cyclohexene species was also noticed since the selectivity to the epoxide
values were 70, 74 and 83% for Mo(VI) Cl2 O2 Bipy/TiO2 (P-25),
Blank photo-oxidation experiments for bare TiO2 under O2 Mo(VI) Cl2 O2 Bipy/TiO2 (HT-400) and Mo(VI) Cl2 O2 Bipy/TiO2 (SC-150),
atmosphere and UV irradiation were first performed so as to con- respectively. Recycling studies were also performed over 56 h of
firm that there was no significant alkene conversion (less than 1%). reaction and they provided evidence of the stability of the catalysts.
Moreover no activity was observed for the anchored catalytic sys- Interestingly the yield to cyclohexene oxide and the selectivity val-
tems in the dark. ues were quite similar for the fresh and used catalysts based on
Results presented in Fig. 7 showed the cyclohexene the TiO2 (SC-150) and P-25 supports, while a significant drop of
photo-conversion in the presence of the various dioxo- yield was observed for the Mo(VI) O2 Cl2 Bipy/TiO2 (HT-400) system
molybdenum(VI) anchored-TiO2 catalysts. It appeared that after 56 h of reaction. Again such data confirmed that the TiO2
there was a strong influence of the hybrid Mo(VI) -TiO2 based synthesized in supercritical CO2 medium exhibited appropriate
catalysts on the cyclohexene conversion at the beginning of characteristics for acting as support for the immobilization of the
the reaction (300 min), whereas a quasi-linear increase of dioxo-Mo(VI) complex.
conversion with reaction time was observed. All the cata- Fig. 8a shows the evolution of the moles of cyclohexene oxide
lysts were active in cyclohexene oxidation in the following per mole of dioxo-Mo(VI) complex as a function of irradiation time
order: Mo(VI) Cl2 O2 Bipy/TiO2 (SC-150) > Mo(VI) Cl2 O2 Bipy/TiO2 (HT- for the various Mo(VI) Cl2 O2 Bipy/TiO2 systems. When compared
400) > Mo(VI) Cl2 O2 Bipy/TiO2 (P-25). The dioxo-Mo(VI) immobilized with the commercial TiO2 P-25 matrix, a significant increase was
on the TiO2 SC-150 sample proved to be the most active with obtained with the mesoporous titania supports, in accordance with
respect to the other catalysts. Such behavior could be related to their higher surface area that promotes the absorption-desorption
an increased amount of exposed dioxo-Mo(VI) species over the reaction between the substrate and the product. One can also note
surface of the mesoporous TiO2 prepared under supercritical CO2 the higher ability of the complex immobilized onto the mesoporous
(Table 1). TiO2 -SC support to transfer oxygen atoms to cyclohexene.
The results presented in Table 2 indicated that no allylic The stability of the catalyst based on the mesoporous TiO2
products were observed. In each case, the selectivity to the support synthesized under SC medium during the epoxidation of
H. Martínez et al. / Journal of Molecular Catalysis A: Chemical 423 (2016) 248–255 253

Fig. 9. Cyclohexene oxidation with the different Mo(VI) Cl2 O2 Bipy/TiO2 catalysts.
Time evolution of the mol of epoxide per mol of dioxo-Mo(VI) species in successive
cycles of light under N2 and O2 in darkness.

for such differences of behavior, silylation of both mesoporous TiO2


supports was performed. The aim was to determine the amount
of reactive OH groups [31] able to serve as active sites for the
anchoring of the Mo(VI) complex. The number of reactive OH of the
hydrothermally treated titania support (TiO2 (HT-400)) was found
to be around 1/3 of that of the mesoporous TiO2 (SC-150). Leach-
ing of Mo(VI) species may therefore be due to a relative amount
of complex physisorbed instead of anchored on the mesoporous
TiO2 (HT-400) support.
The photo-oxidation of cyclohexene was also performed under
inert nitrogen atmosphere and UV–vis light irradiation over the
three Mo(VI) Cl2 O2 Bipy/TiO2 systems, to show the real effect of the
anchored dioxo-Mo entity in stoichiometric conditions. As shown
Fig. 8. (a) Evolution of cyclohexene oxide under molecular oxygen atmosphere and in Fig. 9, Mo(VI) Cl2 O2 Bipy/TiO2 (P-25), Mo(VI) Cl2 O2 Bipy/TiO2 (HT-
UV–vis light during 35 h with continuous irradiation and (b) evolution of cyclohex- 400) and Mo(VI) Cl2 O2 Bipy/TiO2 (SC-150) catalysts undergo an
ene oxide during 4 reaction cycles (56 h) for the fresh and used Mo(VI) Cl2 O2 Bipy/TiO2 oxygen atom transfer to cyclohexene during the first illumina-
SC-150 catalyst.
tion cycle under N2 atmosphere. The amount of cyclohexene oxide
produced corresponded to the transfer of one oxygen atom from
cyclohexene is presented in Fig. 8b. The Mo(VI) Cl2 O2 Bipy/TiO2 (SC- the MoO2 unit (stoichiometric amount of the OAT reaction), which
150) catalyst was used in a second reaction cycle during long-term resulted in the formation of the reduced anchored MoO(IV) species,
photo-oxidation. Interestingly, the catalytic activity is perfectly as depicted in Scheme 1.
maintained after 56 h of reaction with almost no reduction in the Then light was tuned off and the dinitrogen evacuated and
yield to cyclohexene oxide, which is a strong indication of the sta- replaced by O2 in order to reoxidize the reduced MoO(IV) species.
bility of the covalently anchored dioxo molybdenum complex on During this period no cyclohexene epoxidation was observed since
the TiO2 elaborated under supercritical CO2 . no OAT occurred. The dioxygen was further evacuated, replaced by
However, when the dioxo-Mo(VI) complex was immobilized on N2 and the light turned back on. After this stage, all Mo(VI) Cl2 O2 Bipy
the surface of the titania prepared under hydrothermal conditions anchored systems exhibited immediately an oxidation reaction.
(Mo(VI) Cl2 O2 Bipy/TiO2 (HT-400) system), a significant drop of activ- The OAT process was twice more important than in the first irra-
ity was observed (not shown). Such behavior was related to a diation cycle, thereby indicating that two oxygen atoms per Mo
partial leaching of the Mo(VI) Cl2 O2 Bipy complex as ascertained by center had been transferred. After reoxidation with O2 in the dark,
Mo elemental analysis in solution. The amount of Mo leached was 2 mol of cyclohexene oxide were indeed produced during the sec-
found to be 6.5 wt%, which prompted us to carry out the photo- ond and third cycles of illumination, which corresponds to 2 oxygen
oxidation with the supernatant solution under irradiation and O2 atoms transferred from the MoO(O2 ) unit [32–34], in agreement
atmosphere. The presence of the epoxide was observed in the with Scheme 1.
supernatant solution of the recovered Mo(VI) Cl2 O2 Bipy/TiO2 (HT-
400) sample, while no photo-oxidation reaction occurred under 3.3. Photo-oxidation of cyclooctene and 1-octene
these conditions for the analogous Mo(VI) Cl2 O2 Bipy/TiO2 (P-25) and
Mo(VI) Cl2 O2 Bipy/TiO2 (SC-150) catalysts (ESI, Fig. S1). Evidence of Remarkably, we also found that cyclooctene epoxidation over
Mo(VI) species in the supernatant solution of the catalyst based on Mo(VI) Cl2 O2 Bipy/TiO2 (SC-150) under O2 atmosphere and UV–vis
the TiO2 -HT sample was confirmed by UV–vis analysis. The UV–vis light during 35 h of reaction led to cyclooctene oxide with 100%
spectrum of this solution (after solid filtration) showed a strong yield (Fig. 10). Photo-oxidation of 1-octene with such catalyst
absorption band at 560 nm (ESI, Fig. S2), corresponding to the pres- resulted in 98% of selectivity to the corresponding 1-octene oxide,
ence of the Mo = O group [30]. In order to provide some explanations indicating 100% OAT process. In general, cyclic olefins exhibit a
254 H. Martínez et al. / Journal of Molecular Catalysis A: Chemical 423 (2016) 248–255

Scheme 1. Stoichiometric oxygen transfer from the MoO2 (VI) entity to cyclohexene under inert N2 atmosphere and UV–vis irradiation, as well as O2 activation by the
Mo(VI) Cl2 O2 Bipy/TiO2 system under mild conditions.

Fig. 10. Selectivity and conversion values for the photo-oxidation of various alkenes
over Mo(VI) Cl2 O2 Bipy/TiO2 (SC-150) after 35 h of reaction.
Fig. 11. Epoxides production after 35 h of reaction under O2 and UV–vis light with
the Mo(VI) Cl2 O2 Bipy/TiO2 (SC-150) catalyst.

higher reactivity than that of linear olefins. As can be observed in


Fig. 10, higher conversion values were indeed obtained in the case medium by filtration after a catalytic run for 6 h, then washed with
of cyclohexene and cyclooctene with respect to 1-octene. acetone and finally dried before being subjected to another cycle of
The evolution of the mol of epoxides (cyclooctene oxide, cyclo- reaction. Remarkably, the activity of the Mo(VI) Cl2 O2 Bipy/TiO2 (SC-
hexene oxide and 1-octene oxide) per mol of dioxo-Mo(VI) species 150) catalyst was preserved after the six cyclooctene epoxidation
as a function of irradiation time is presented in Fig. 11 for the cycles during 6 h under O2 atmosphere and light. The yield to
Mo(VI) Cl2 O2 Bipy/TiO2 (SC-150) catalyst. It can be seen that the cyclooctane epoxide was perfectly maintained as well as the con-
activity of the mesoporous TiO2 -SC support modified with the version values after six reaction cycles (ESI, Table S2).
dioxomolybdenum(VI) complex was maintained during 35 h in the
photo-epoxidation of the various olefins in the presence of O2 as 4. Conclusions
O-donor and light.
In order to assess the stability and reusability of the dioxo- In this paper, it was demonstrated that dichloro-dioxo-[4,4 -
Mo(VI) modified mesoporous TiO2 prepared under SC CO2 during dicarboxylato-2,2 -bipyridine]-Mo(VI) complexes anchored on the
the photo-oxidation of cyclooctene, a series of six successive runs surface of various mesoporous TiO2 were highly active, selec-
was performed. The solid catalyst was separated from the reaction tive and stable for the photo-assisted epoxidation of alkenes with
H. Martínez et al. / Journal of Molecular Catalysis A: Chemical 423 (2016) 248–255 255

molecular oxygen as a primary oxidant and UV–vis light. The activ- [5] J.W. Brown, Q.T. Nguyen, T. Otto, N.N. Jarenwattananon, S. Glöggler, L.-S.
ity may be related to the density of the titania surface hydroxyl Bouchard, Catal. Commun. 59 (2015) 50–54.
[6] K. Shimizu, T. Kaneko, T. Fujishima, T. Kodama, H. Yoshida, Y. Kitayama, App.
groups that favored the dispersion of the Mo(VI) complex on the Catal. A: Gen. 225 (2002) 185–191.
surface of the supports and hence the catalytic properties. In this [7] S. Ouidri, C. Guillard, V. Caps, H. Khalaf, App. Clay Sci. 48 (2010) 431–437.
respect, the mesoporous TiO2 prepared under supercritical CO2 [8] M.S. Hamdy, G. Mul, Catal. Sci. Technol. 2 (2012) 1894–1900.
[9] N.S. Patil, B.S. Uphade, P. Jana, S.K. Bharagava, V.R. Choudhary, J. Catal. 223
conditions exhibited the highest content and dispersion of the (2004) 236–239.
grafted Mo(VI) Cl2 O2 Bipy complex. Such mesoporous TiO2 support [10] T. Ohno, K. Nakabeya, M. Matsumara, J. Catal. 176 (1998) 76–81.
was shown to possess the most appropriate textural characteristics [11] S. Huber, M. Cokoja, F.E. Kühn, J. Organomet. Chem. 751 (2014) 25–32.
[12] M. Bagherzadeh, M. Zare, T. Salemnoush, S. Ozkar, S. Akbayrak, Appl. Catal. A
and proved after the dioxo-Mo(VI) immobilization step to be the
475 (2014) 55–62.
most efficient in terms of activity and selectivity to the formation [13] C. Muller, N. Grover, M. Cokoja, F.E. Kühn, Adv. Inorg. Chem. 65 (2013) 33–83.
of various epoxides. Moreover such catalyst was stable under the [14] T.R. Amarante, A.C. Gomez, P. Neves, F.A. Almeida-Paz, A.A. Valente, M.
Pillinger, I.S. Gonçalves, Inorg. Chem. Commun. 32 (2013) 59–63.
reaction conditions as demonstrated by several reuses even after
[15] S.A. Hauser, M. Cokoja, F.E. Kühn, Catal. Sci. Technol. 3 (2013) 552–561.
long-term photo-oxidation experiments. [16] Y. Li, X. Zhou, S. Chen, R. Luo, J. Jiang, Z. Liang, H. Ji, RSC Adv. 5 (2015)
30014–30020.
[17] S.M. Islam, K. Ghosh, R.A. Molla, A.S. Roy, N. Salam, M.A. Iqubal, J. Organomet.
Acknowledgements
Chem. 774 (2014) 61–69.
[18] C.A. Páez, N.J. Castellanos, F. Martinez, F. Ziarelli, G. Agrifoglio, E. Páez-Mozo,
This work was financially supported by the Universidad Indus- H. Arzoumanian, Catal. Today 133 (2008) 619–624.
trial de Santander, Colombia (project DIEF 9313 and sabbatical year [19] C.A. Páez, O. Lozada, N.J. Castellanos, F. Martinez, F. Ziarelli, G. Agrifoglio, E.
Páez-Mozo, H. Arzoumanian, J. Mol. Catal. A 299 (2009) 53–59.
of FM), HM and NJC were supported by COLCIENCIAS, Colombia [20] N.J. Castellanos, F. Martinez, F. Lynen, S. Biswas, P. Van Der Voort, H.
(Fondo Apoyo a Doctorados Nacionales). FM was financially sup- Arzoumanian, Trans. Met. Chem. 38 (2012) 119–127.
ported during sabbatical year by the Region Poitou-Charentes, the [21] H. Arzoumanian, N.J. Castellanos, F. Martinez, E.A. Paez-Mozo, F. Ziarelli, Eur. J.
Inorg. Chem. (2010) 1633–1641.
CNRS and the University of Poitiers, France (contract 09/RPC-R- [22] X. Zhang, S. Heinonen, E. Levänen, RSC Adv. 4 (2014) 61137–61152.
055). The authors are also very much indebted to Jean-Christophe [23] T. Peng, D. Zhao, K. Dai, W. Shi, K. Hirao, J. Phys. Chem. B 109 (2005)
Ruiz from CEA Bagnols sur Cèze (France) for fruitful discussions 4947–4952.
[24] N. Preda, L. Mihut, M. Baibarac, I. Baltog, M. Husanu, C. Bucur, T. Velula, Rom. J.
regarding the TiO2 synthesized under supercritical CO2 . Phys. 54 (7–8) (2000) 667–675.
[25] K.H. Leong, P. Monash, S. Ibrahim, P. Saravanan, Sol. Energy 101 (2014)
Appendix A. Supplementary data 321–332.
[26] G.R. Hearne, J. Zhao, A.M. Dawe, V. Pischedda, M. Maaza, M.K. Nieuwoudt, P.
Kibasomba, Phys. Rev. 70 (2004) 134102–134110.
Supplementary data associated with this article can be found, in [27] F. Tian, Y. Zhang, J. Zhang, C. Pan, J. Phys. Chem. C 116 (2012) 7515–7519.
the online version, at http://dx.doi.org/10.1016/j.molcata.2016.07. [28] N.L. Alpert, W.E. Keiser, H.A. Szymanski, IR; Theory and Practice of Infrarred
Spectroscopy, Plenum Press, New York, 1970, pp 209, 261 and 268.
006.
[29] V. Amani, A. Abedi, S. Ghabeshi, H. Reza, S.M. Hosseini, N. Safari, Polyhedron
79 (2014) 104–115.
References [30] H. Arzoumanian, G. Agrifoglio, M.V. Caparelli, R. Atencio, A. Briceño, A.
Alvarez-Larena, Inorg. Chim. Acta 359 (2006) 81–89.
[31] M.C. Capel-Sanchez, J.M. Campos-Martin, J.L.G. Fierro, Catal. Today 158 (2010)
[1] B. Puértolas, A.K. Hill, T. García, B. Solsona, L. Torrente-Murcino, Catal. Today
103–108.
248 (2015) 115–127.
[32] J.M. Berg, R.H. Holm, J. Am. Chem. Soc. 107 (1985) 925–932.
[2] X.L. Wei, X.-H. Lu, X.-T. Ma, C. Peng, H.-Z. Jiang, D. Zhou, Q.-H. Xia, Catal.
[33] C. Lorber, M.R. Plutino, L.I. Elding, E. Nordlander, J. Chem. Soc. Dalton Trans. 2
Commun. 61 (2015) 48–52.
(1997) 3997–4003.
[3] Y. Yan, X. Tong, K. Wang, X. Bai, Catal. Commun. 43 (2014) 112–115.
[34] J. Cipot-Wechsler, D. Covelli, J.M. Pratorius, N. Hearns, O.V. Zenkina, E.C. Keske,
[4] A. Feliczak-Guzik, A. Wawrzynczak, I. Nowak, Microporous Mesoporous
R. Wang, P. Kennepohl, C.M. Crudden, Organometallics 31 (2012) 7306–7315.
Mater. 202 (2015) 80–89.

You might also like