Vous êtes sur la page 1sur 9

View Online

PAPER

www.rsc.org/dalton | Dalton Transactions

Studies on bis(halogeno) dioxomolybdenum(VI)-bipyridine complexes: Synthesis and catalytic activity


Alev G unyar,a Ming-Dong Zhou,a Markus Drees,a Paul N. W. Baxter,b Ghada Bassioni,c Eberhardt Herdtwecka and Fritz E. K uhn*a
Received 7th May 2009, Accepted 11th August 2009 First published as an Advance Article on the web 28th August 2009 DOI: 10.1039/b909075d Dioxomolybdenum(VI) complexes with the general formula [MoO2 Cl2 L2 ] (L2 =3,3-dimethyl-2,2-bipyridine, 5,5-dimethyl-2,2-bipyridine, 6,6-dimethyl-2,2-bipyridine, 4,4-dibromo-2,2-bipyridine, 5,5-dibromo-2,2-bipyridine, 5,5-diamino-2,2-bipyridine; 5,5-dinitro-2,2-bipyridine; 5,5-di-ethoxycarbonyl-2,2-bipyridine; 6-phenyl-2,2-bipyridine; 2,2:6,2-terpyridine) have been prepared and characterised. [MoO2 Cl2 (5,5-diethoxycarbonyl-2,2-bipyridine)] has been examined by single crystal X-ray analysis. The complexes were applied as homogenous catalysts for the epoxidation of cyclooctene with tert-butyl hydroperoxide (TBHP) as oxidising agent. The new compounds show an overall high activity and are highly selective catalysts in the epoxidation of cyclooctene. The stability of the complexes and differences in the catalytic activity can be clearly attributed to electronic contributions of the functional groups on bipyridine ligands and to steric restrictions. DFT calculations have assisted in a better understanding of the stability of the complexes and are in agreement with experiment. The inuence of the terminal oxo ligands and the Lewis base ligands on the Mo center keep the compounds on quite a stable level of electron density.

Downloaded by Universiteit Gent on 05 November 2010 Published on 28 August 2009 on http://pubs.rsc.org | doi:10.1039/B909075D

Introduction
In the last decade, signicant research efforts both in industry and academia have been dedicated to nding efcient routes to valuable epoxides. As described earlier, epoxidation can be performed in a variety of ways.1 The route we are following in this work applies a higher concentration of a hydroperoxide (usually produced in a preliminary reaction generally by autoxidation with air or O2 ), which transfers an oxygen atom in a second step. This method has established a dominant position in industry. Among the variety of efcient catalysts known today for olen epoxidation are several organometallic and inorganic oxides containing a metal in its highest oxidation state.2 It has been reported that molybdenum(VI) complexes are versatile catalysts for the oxidation of organic compounds.3 Treatment of [MoO2 X2 ] species (X = halide, OR, OSiR3 ) with mono- or bidentate Lewis bases (L or L2 ), such as pyridine, 2,2-bipyridine and 1,10phenanthroline, and with donor solvents, such as acetonitrile and THF, give rise to adducts of the composition [MoO2 X2 L2 ].2,4-6 In the presence of tert-butyl hydroperoxide (TBHP) as the oxygen source,7-9 some of these complexes have shown excellent catalytic
a Molecular Catalysis, Catalysis Research Centre, Technische Universit at M unchen, Lichtenbergstrasse 4, D-85747, Garching, Germany. E-mail: fritz.kuehn@ch.tum.de; Fax: +49 89 289 13473; Tel: +49 89 289 13081 b Charles Sadron, Universit e Louis Pasteur, 23 rue du Loess BP 20, 67083, Strassbourg, France. E-mail: baxter@ics.u-strasbg.fr c On leave from: Chemistry Department, Faculty of Engineering, Ain Shams University, Cairo, Egypt. E-mail: Ghada.Bassioni@lrz.tu-muenchen.de Electronic supplementary information (ESI) available: 1 H NMR spectra, TGA curves of the synthesised [MoO2 Cl2 L2 ] complexes 1, 3, 4, 6, 7, 913. CCDC reference number 730946 (11). For ESI and crystallographic data in CIF or other electronic format see DOI: 10.1039/b909075d

activities. The advantages of TBHP as an oxidising agent are its good thermal stability and ease of handling.10 Furthermore, the byproduct of the reaction, tert-butanol, is easily separated by distillation and can be reused.11 Important properties, such as the solubility of the complex and the Lewis acidity of the metal centre, can be ne tuned by varying either X or L. These adjustments allow the preparation of highly active and selective catalysts.12 The [MoO2 X2 L2 ] complexes present a distorted octahedral geometry, with the oxo ligands adopting a cis position in order to maximise backdonation into the empty t2g set orbitals.13 Kinetic and mechanistic investigations of catalytically applicable molybdenum(VI) compounds have focused largely on inorganic Mo oxide catalysts.14 In all cases, the experimental data agree with the formation of a coordinated alkyl peroxo species as the active intermediate, which transfers an oxygen atom to the substrate. Involvement of a h2 -peroxo Mo complex as the active species in the catalytic reaction is considered to be unlikely.14 From the dioxomolybdenum complexes studied for the epoxidation of cyclooctene, [MoO2 Cl2 L2 ] proved to be among the most efcient catalysts.14a Different mechanisms have been suggested and it is still a subject of discussion whether they can explain the reactivity of these complexes, despite a large amount of information that has been derived from NMR studies and catalytic reactivity patterns.14e It has however been reported that the compound [MoO2 Cl2 (4,4-dimethyl-2,2-bipyridine)] is not an active catalyst of the epoxidation of cyclooctene with molecular oxygen at atmospheric pressure, but is active with TBHP as oxidant at 55 C, yielding cyclooctene oxide as the only product.14 The catalytic activity of the complexes is assumed to depend strongly on the substitution pattern of the donor ligand, in agreement with previous studies on related oxomolybdenum complexes.14 This journal is The Royal Society of Chemistry 2009

8746 | Dalton Trans., 2009, 87468754

View Online

In this work, the synthesis and spectroscopic properties of [MoO2 Cl2 L2 ] (1: L2 = 3,3-dimethyl-2,2-bipyridine, 3: L2 = 5,5dimethyl-2,2-bipyridine, 4: L2 = 6,6-dimethyl-2,2-bipyridine, 6: L2 = 4,4-dibromo-2,2-bipyridine, 7: L2 = 5,5-dibromo2,2-bipyridine, 9: L2 = 5,5-diamino-2,2-bipyridine, 10: L2 = 5,5-dinitro-2,2-bipyridine, 11: L2 = 5,5-diethoxy-carbonyl-2,2bipyridine, 12: L2 = 6-phenyl-2,2-bipyridine, 13: L2 = 2,2:6,2terpyridine) are reported. The electronic contributions of the different functional groups in different positions on the bipyridine rings and the implications of these differences on the catalytic activity are the subjects of the present study.

Downloaded by Universiteit Gent on 05 November 2010 Published on 28 August 2009 on http://pubs.rsc.org | doi:10.1039/B909075D

Results and discussion


Synthesis and characterisation The synthesis of a series of bipyridine derivatives of dioxomolybdenum(VI) complexes has been performed following a literature known route depicted in Scheme 1.15 Compounds 1, 3, 4, 6, 7, 9, 10, 11, 12 and 13 are obtained as either pale-yellow or white microcrystalline powders.

Scheme 2 [MoO2 Cl2 L2 ] complexes, as well as the hypothetical complexes 5 and 8. Scheme 1 Synthesis of [MoO2 Cl2 L2 ] complexes.

Most of the products are signicantly less soluble in almost all organic solvents than the starting materials and precipitate from the reaction mixture within a few minutes after their formation. They can be puried (e.g. from excess ligand) by washing with diethyl ether. All compounds are air and moisture stable, which is in remarkable contrast to the starting material [MoO2 Cl2 (THF)2 ], which can only be handled and stored under an inert gas atmosphere. In order to obtain a better insight into the catalytic activities of [MoO2 Cl2 L2 ]-type compounds, a series of bipyridine ligands bearing different functional groups at different positions of the bipyridine rings and the inuence of these changes on the activity of cyclooctene epoxidation were monitored. Complex 2, where the dimethyl groups are in the 4,4-position of the bipyridine ligand, has been previously reported.14 Complexes 5 and 8, shown in Scheme 2, are hypothetical compounds. To the best of our knowledge, the ligand for complex 5 has so far not been synthetically accessible, most likely due to the pronounced steric hindrance of the two bromides in close proximity to each other (see Scheme 2 and discussion below). The attempt to synthesise complex 8 failed, affording only free ligand and MoO2 Cl2 separately in the reaction solution. This journal is The Royal Society of Chemistry 2009

NMR spectroscopy. An important feature of [MoO2 Cl2 L2 ]type complexes is that they are insoluble or fairly soluble in most of the common organic solvents. Nevertheless, the majority of the prepared compounds are soluble enough in polar, noncoordinating organic solvents to give 1 H NMR spectra of sufcient resolution. In general, the spectroscopic data for these complexes are in good agreement with those obtained from the other Lewisbase adducts (see the ESI).14 The proton signals of all the examined compounds are shifted to lower eld after complexation. The chemical shift differences between the free ligands and the corresponding complexes largely depend on the electron donating ability of each ligand to the metal centre. The proton signals of complex 10, bearing strong electron withdrawing nitro groups on the aromatic rings, appear at lower eld than the other complexes, followed by ethoxycarbonyl and bromo substituted complexes 11, 7 and 6. For compounds having electron donating groups on the ligands (for instance, methyl, phenyl and amino groups), the proton signals appear at comparatively higher eld. The strong electron donating amino substituted complex 7 shows the proton signals at the highest eld among all the examined complexes. Calculating the coupling constants J of adjacent protons in the ring of the ligand before and after complexation provides some information about the structure. For example, the J coupling of H6,6 in 3,3-dimethyl-2,2-bipyridine is 8.0 Hz, while after Dalton Trans., 2009, 87468754 | 8747

View Online

Downloaded by Universiteit Gent on 05 November 2010 Published on 28 August 2009 on http://pubs.rsc.org | doi:10.1039/B909075D

complexation this coupling becomes 4.8 Hz; an indication of a twisted molecular geometry. The assumption of a somewhat twisted bipy ligand is in good agreement with the predicted structure according to the DFT calculations (see below and Fig. 2). 95 Mo NMR spectroscopy is in most cases a sensitive tool to determine the electron density on the metal centre.14a,c Nevertheless, 95 Mo NMR spectra of complexes 3, 6, 7, 9 and 13 could not be obtained due to their poor solubility. Concerning the obtained spectra for complexes 1, 4, 8 and 1012 (see Table 1), the differences between the 95 Mo chemical shifts, however, are comparatively small. Complexes of the type [MoO2 Cl2 L2 ] generally show their signals in the range 160220 ppm.16 As expected, the electronic contribution of a methyl group in the para position (4,4position with respect to the bipyridine nitrogen atoms), as in the case of compound 2, reects the most signicant difference in its 95 Mo spectrum as compared to the solvent substituted reference compound [MoO2 Cl2 (THF)2 ]. It seems that with the methyl groups in different positions, the electronic contribution on the 95 Mo chemical shifts follows the order para, 2 > ortho, 4 > meta, 1. IR spectroscopy. The FT-IR spectroscopic data for the examined complexes are generally consistent with other Lewisbase adducts of bis(halogeno)dioxomolybdenum(VI) described before.14a The symmetric and asymmetric IR stretching vibrations for the cis-dioxo unit of the complexes are found in the expected range between 904 and 949 cm-1 . The observed stretching vibrations (see the Experimental section) as well as the calculated force constants (see Table 1) show that the Mo=O bond strengths increase in the order 4 > 1 > 3 among the complexes with CH3 substituents on the bipy ligands. For the two Br substituted derivatives, complex 6 displays stronger Mo=O bonds than complex 7. The inuence of different substituents on the ligand donor ability is clearly reected by the Mo=O force constants calculated from the IR data. When comparing

Fig. 1 ORTEP plot25f of the molecular geometry of complex 11 in the solid state. Thermal ellipsoids are drawn at the 50% probability level. The hydrogen atoms are omitted for clarity. Selected bond lengths ] and angles [ ]: Mo1Cl1 2.355(1), Mo1Cl2 2.354(1), Mo1O1 [A 1.696(2), Mo1O2 1.692(2), Mo1N11 2.332(2), Mo1N21 2.335(2); Cl1Mo1Cl2 157.29(2), Cl1Mo1O1 96.21(6), Cl1Mo1O2 96.99(6), Cl1Mo1N11 80.96(5), Cl1Mo1N21 80.98(4), Cl2Mo1O1 97.28(6), Cl2Mo1O2 96.70(6), Cl2Mo1N11 80.44(5), Cl2Mo1N21 80.33(5), O1Mo1O2 106.12(8), O1Mo1N11 91.82(7), O1Mo1N21 161.06(7), O2Mo1N11 162.06(7), O2Mo1N21 92.82(7), N11Mo1N21 69.24(6).

Fig. 2 DFT optimised structures of complexes 113.

8748 | Dalton Trans., 2009, 87468754

This journal is The Royal Society of Chemistry 2009

View Online

Table 1 Formation constants, selected catalytic and spectroscopic data for the complexes 113 Complex Ref. 1 2 3 4 6 7 9 10 11 12 13 K eq in THF

l (nm) in THF

Decomp. Temp. ( C)

TOF (h-1 )a

-1 f (Mo=O) mdyn A

d (95 Mo) ppm 174 191 205

(1,6 0.2) 104 (9 3) 106 (2.4 0.5) 106 (6.9 0.7) 103 (5.7 0.6) 104 (4.7 0.2) 103 (3.8 0.9) 106 (1.5 0.5) 101 (3.7 0.4) 104 (2.4 0.2) 103 (1.9 0.1) 103

330 320 318 330 325 350 450 360 340 350 330

275 not determined 375 225 375 300 400 250 275 not determined 250

1862

(6.85)

1015 1916 1584 1124 212 1836 1792 1867 1871

(6.79) (6.96) (6.89) (6.85) (6.74) (6.94) (6.91) (6.91) (6.89)

194

186 194 197


_

Downloaded by Universiteit Gent on 05 November 2010 Published on 28 August 2009 on http://pubs.rsc.org | doi:10.1039/B909075D

The TOF values are calculated after 3 min. Reaction time for all examined complexes and dened as mol epoxide/(mol catalyst time(h)) Ref: [MoO2 Cl2 (DMSO)2 ]. The TOF of the complexes 6 and 7 were calculated with a conversion reached within 35 min time; for complex 9 after 510 min of the reaction.

the 5,5-substituted complexes 3 (CH3 substituents), 7 (Br substituents), 9 (NH2 substituents), 10 (NO2 substituents) and 11 (C(O)OEt substituents), the force constants increase in the order 10 > 11 > 7 > 3 > 9. This is exactly in accordance with the donor/acceptor capabilities of the various substituted ligands, with NH2 being the strongest donor and NO2 the strongest acceptor. The stronger the donor ability of the ligands, the higher the electron density of the metal centre and therefore the weaker are the Mo=O bonds. As compared to the free ligands, the CN stretching vibrations and the CC stretching vibrations shift to higher wavenumbers after coordination with MoO2 Cl2 . For example, the pyridine CN and CC vibrations of 5,5-dimethyl2,2-bipyridine (ligand 3) lie at 1523 and 1553 cm-1 and they shift to 1586 and 1600 cm-1 after coordination. UV/Vis spectroscopy. The formation constants of the title complexes have been calculated by tting the UV/vis absorbance data at equilibrium conditions (see Table 1).17 The higher the value, the more stable are the compounds formed. The calculated formation constants increase in the order 4 > 1 > 3 among the complexes with methyl substituents on the bipy ligands. With respect to the bromo derivatives on the bipy ligands, compound 6 is more stable than compound 7. In general, the stronger the donor ability of the ligand, the higher the stability of the complexes formed. In the case of all complexes with different substituents at the 5,5-positions, the stabilities are in the order 9 > 3 > 11 > 7 > 10. This is also in agreement with that obtained from the calculated Mo=O force constants. Thermogravimetry. Thermogravimetric analysis (TGA) was performed for all the compounds and the curves for each compound are presented (see the ESI). The onset of the rst decomposition temperature for the CH3 -substituted complexes increases in the following order: complex 3 (375 C) > complex 1 (275 C) > complex 4 (225 C) (Table 1). The result is in agreement with the equilibrium constant values obtained from UV-Vis spectroscopy (see Table 1).17 Compound 4 has an equilibrium constant almost three orders of magnitude below that of compound 3. The reason may be due to the unfavorable steric situation of compound 4 originating from the methyl groups of the ligand pointing towards the MoO2 Cl2 moiety and thus hindering ligand coordination to the metal. In the case of the Br-substituted This journal is The Royal Society of Chemistry 2009

complexes, 6 is more stable than 7. Presumably, the electron withdrawing effect of the bromides is more effective in the 5,5 than the 4,4-positions, and therefore, the donor ability of ligand 7 is weaker than that of ligand 6. Accordingly, ligand 7 forms a comparatively weaker coordination bond with the metal than ligand 6. Concerning the different functional groups residing at the 5,5-positions, the rst decomposition onset temperature increases in the order 9 (400 C) > 3 (375 C) > 7 (300 C) > 11 (275 C) > 10 (250 C). Again, the order of the rst decomposition onset is in good agreement with the calculated Mo=O force constants and the complex formation constants. Complexes that have a high complex formation constant also have a high onset of the rst decomposition point but a weak Mo=O force constant. The weakest bonding interaction of the molecules are, however, the NMo donor interactions. The stronger the donor capability of the ligand, the more stable the obtained complex. X-ray single crystal structure analysis. Some non-substituted bipy adducts of MoO2 X2 had been characterised by X-ray crystallography before.14a The MoCl, Mo=O and MoN bond distances for compound 11 are very similar to those of the complexes determined previously. Although the MoCl bond distances in compound 11 appear to be very slightly shorter and the MoN bond distances seem to be slightly longer than in [MoO2 Cl2 (2,2-bipyridine)], the differences are too small to allow a detailed discussion. The donor capability differences are seemingly too small to have a large impact on the bond distances of the molecules. It has also been observed that tiny differences in the donor/acceptor capability of the ligands have only a small inuence on the catalytic activities. A detailed discussion can be found in the catalysis section.2 Fig. 1 shows the crystal structure of compound 11. DFT calculations. An explanation for the results obtained by NMR spectroscopy can be provided by DFT calculations (B3LYP/6-31G* with Stuttgart-ECP for Mo). 95 Mo NMR spectra of complexes 1 and 4 show comparable chemical shifts, which is explained by the calculated atomic charge on the metal centre in both cases, i.e. +1.116 and +1.102, respectively. Therefore a similar electron density should occur at the molybdenum atom in both complexes 1 and 4. Because of the lack of suitable crystals, the ground states for complexes 113 have been calculated to gain a Dalton Trans., 2009, 87468754 | 8749

View Online

Table 2 Calculated 13 C NMR shifts, Mulliken charges for Mo and N atoms and distortion of the bipyridine backbone of complexes 113 Complex 1 2 3 4 5 6 7 8 9 10 11 12 13
13

C GIAO 2,2-C (ppm)

13

C GIAO 6,6-C (ppm)

Mulliken Charge Mo 1.12 1.14 1.14 1.10 1.11 1.14 1.14 1.16 1.14 1.14 1.14 1.13 1.12

Mulliken Charge N,N -0.48 -0.51 -0.51 -0.51/-0.53 -0.46 -0.50 -0.50 -0.39/-0.43 -0.52 -0.51 -0.51 -0.50/-0.51 -0.51

Distortion angle ( ) 32.7 0.1 0.0 15.5 37.6 0.0 0.0 24.5 0.1 0.0 0.0 17.7 22.7

Downloaded by Universiteit Gent on 05 November 2010 Published on 28 August 2009 on http://pubs.rsc.org | doi:10.1039/B909075D

159.9 157.6 155.4 160.6/161.1 160.6 157.0 154.9 162.9/164.9 147.8 160.2 163.2/161.4 160.2/160.4 160.3/161.0

158.4 159.6 160.3 171.3/172.1 157.8 159.8 161.4 172.8/176.0 145.5 157.3 159.3/159.1 161.1/174.6 161.2/172.5

deeper insight into the geometry of these complexes. The optimised structures of dimethyl bipyridine ligated compounds, the dibromo bipyridine ligated complexes and the other complexes are shown in Fig. 2. It is obvious from those representations that only the 4,4 (2) and 5,5 (3) substituted bipyridines remain planar after complex formation. For the 6,6 (4) and especially for the 3,3 (1) complexes it is energetically more favorable to display a distorted bipyridine geometry. As can be seen from the structure plots, all the complexes carrying functional groups like esters, amines or nitro moieties on the bipyridine ligand show that the bipyridine remains planar after addition to the molybdenum centre. Only the bipyridine derivative bearing one additional aromatic ring loses its planarity after complexation. The distortion of the bipyridine in the complexes can also be derived from both the computed and experimental NMR chemical shifts. An indicator for the electron density at the MoN contact is the chemical shift of the neighboring carbons (2,2 and 6,6). Table 2 summarises these chemical shifts calculated with GIAO-DFT and compares them with the dihedral angle spanned by the two pyridine halves of the bipyridine ligands and with the Mulliken charge of the Mo centre and the two bipyridine nitrogen atoms. As can be seen from Table 2, distorted bipyridines in the case of 6,6(4,8) and 3,3(1, 5) substituted compounds lead to a downeld shift of the carbons in the 6,6 and 2,2 positions. This indicates that these two carbons, adjacent to the nitrogen atoms, lose electron density. Additionally, the metal centre has some changes in the electronic distribution. Planar bipyridine ligands lead in all cases to an atomic charge of 1.14 at the molybdenum and around -0.5 for the nitrogen atoms. In the case of methyl, a distorted ligand leads to lower charges with slightly more electron density being available at the metal. In the case of the bromines being at the 6,6 positions (8), electron density is removed from the Mo atom and the coordinating nitrogen leading to a higher positive charge both at the metal and the coordinating nitrogen atoms. For the hypothetical compound 5, the metal would be expected to receive more electron density while the nitrogen atoms lose some. The thermodynamic stability of adducts is controlled by enthalpic and entropic factors. For example, the enthalpy calculated from the experimental UV/Vis data obtained for the 6,6-dimethyl complex 4 is DH = -42.1 kJ mol-1 .17 This energy gain (exothermic reaction) is due to the replacement of the weak donor ligand THF by the stronger donor bipy. The entropy change results from the release of two solvent molecules and the coordination
8750 | Dalton Trans., 2009, 87468754

Table 3 Calculated enthalpies and free energies of formation for complexes 113 Complex 1 2 3 4 5 6 7 8 9 10 11 12 13 Substituent Me Me Me Me Br Br Br Br NH2 NO2 C(O)OEt Ph Py Position(s) 3,3 4,4 5,5 6,6 3,3 4,4 5,5 6,6 5,5 5,5 5,5 6 6 DH (kJ mol-1 ) -80.4 -104.3 -104.3 -93.4 -62.0 -81.6 -76.6 +2.1 -114.1 -56.8 -86.2 -16.0 -14.5 DG (kJ mol-1 ) -21.8 -47.3 -51.5 -40.2 -2.9 -26.8 -21.4 +56.5 -58.9 -2.5 -31.4 -2.7 -0.8

of the bipyridine ligand. The overall reaction entropy is slightly negative, indicating the complex stability being mainly due to an enthalpy change rather than to entropic effects. DFT calculations demonstrate that all bipy-methyl derivatives show negative DH and DG values for the formation of the corresponding bipyridine complex from MoO2 Cl2 . In this case the calculations do not start from solvent ligated [MoO2 Cl2 (THF)2 ] giving rise to higher formation energies than experimentally observed (see above). Therefore, these values reect more the stability trends amongst the different complexes, rather than the real formation energy in a THF solution. The calculated DH and DG values are summarised in Table 3. According to the calculations, the order of the stability of methyl substituted complexes is: 5,5 (3) = 4,4 (2) > 6,6 (4) > 3,3 (1). In the case of the bromide substituted bipyridines, the order changes and complex 8 shows an endothermic and endergonic formation. This might help to explain why complex 8 is not synthetically feasible. The order of stability is 4,4 (6) > 5,5 (7) > 3,3 (5) 6,6 (8). Considering substituent effects, the complex formation energies of compounds that bear different substituents in the 5,5 positions are compared. The general tendency is that substituents with electron donating groups (amine, methyl) show better stabilities than complexes with electron withdrawing groups (bromide, ester, nitro). The order according to the DFT calculations is NH2 (9) > CH3 (3) C(O)OEt (11) > Br (7) NO2 (10), tting well with the conclusions drawn from the experimental results described above. This journal is The Royal Society of Chemistry 2009

View Online

Downloaded by Universiteit Gent on 05 November 2010 Published on 28 August 2009 on http://pubs.rsc.org | doi:10.1039/B909075D

Fig. 3 Time-dependent yield of cycloocteneepoxide in the presence of MoO2 Cl2 and compounds 113 as catalysts at 55 C with 1 mol% catalyst charged. The yield obtained within the initial 30 min is shown in the inset of the gures.

In the case of compounds 12 and 13, the bipy complexes bear only one substituenta third aryl ring in the ortho position to the nitrogen. Because of the close proximity of the aryl substituent to the MoO2 Cl2 centre, these ligands have a reduced planarity. The stability of both 12 (R=Ph) and 13 (R=Py) is reduced due to the steric demand of the third ring, while 12 is somewhat more stable than 13. Application in epoxidation catalysis The synthesised Mo compounds were tested as catalysts for the epoxidation of cyclooctene with TBHP (Fig. 3). The details concerning the catalytic reaction are given in the Experimental section. In the absence of a catalyst, no signicant amount of epoxide is formed. Two series of experiments with different catalyst : substrate : oxidant ratios (1 : 100 : 150 & 1 : 1000 : 1500) were undertaken in order to compare the catalytic potential of the systems. The catalytic reactions show time-dependent curves in which the yield increases quickly in the initial part of the reaction and thereafter slows down when the reaction approaches completeness, due to an increasing lack of substrate (Fig. 3). All examined compounds show quite high catalytic activity. Particularly complexes 1, 4, 1013 display TOFs higher than 1000 h-1 and reach yields well exceeding 90% within the rst hour of reaction time. Yet, complexes 6, 7 and 9 show an induction period within the rst minutes of reaction. In these cases, the precatalysts dissolved very slowly after adding TBHP into the reaction media which leads to a low catalyst concentration and hence to a slow reaction in the rst minutes, whereas the other complexes immediately form the catalytically active, well soluble peroxospecies. After the catalysts have totally dissolved, the reaction speeds up quickly. The situation with respect to the catalytic activity of complexes having different substituents on the same position of the ring is rather subtle. The strongest donor substituted complex 9 shows the lowest turnover frequency among all the complexes (see Table 1); however, the TOFs of complexes 3, 7, 10, 11 are quite similar. This is not surprising when looking at the Mulliken charge at the Mo atom (Table 2), being identical for these complexes. The degree of electron density at the Mo centre should be decisive for the reaction, according to the previously published reaction mechanism based largely on theoretical calculations.14 Accordingly, a charge balance must be established at the Mo This journal is The Royal Society of Chemistry 2009

centre, depending on the donor/acceptor capabilities of the ligands, rendering the electron density at the Mo centre rather independent from the substituents attached to the ligand. Only ligands with greatly different donor capabilities should inuence the catalytic activity in a more pronounced way. Surprisingly, changing the catalyst : substrate : oxidant ratio from 1 : 100 : 150 to 1 : 1000 : 1500 does not change the activity signicantly. It appears that optimal activities for the system have been reached. Lowering the reaction temperature to 0 C, however, lowers the activity. In the case of compound 4, the TOF decreases from ca. 1150 h-1 (at 55 C) to ca. 750 h-1 (at 0 C). Adding a second batch of substrate after the rst run, the reaction reaches more than 95% product yield. However, the reaction is not so fast as in the rst run. This phenomenon is already known from related complexes and has been ascribed to a reaction of the catalyst with the byproduct t-BuOH, rather than a catalyst decomposition.10 It is important to note that the reaction is selective to the desired epoxide and that no diol formation is observed.

Conclusion
Molybdenum(VI) complexes of the general formula [MoO2 Cl2 L2 ] are easily synthesised from [MoO2 Cl2 (THF)2 ] and 2,2-bipyridine derivatives. The stability of the products is strongly inuenced by the position of substituents on the 2,2-bipyridine ligand. Steric hindrance in the 6,6 and 3,3-positions may prevent complex formation if larger substituents (e.g. Br) are in these positions. DFT calculation is a useful tool to explain the experimental results and to predict tendencies such as substituent effects. All the complexes prepared in this work are active catalysts for cyclooctene epoxidation. Particularly in comparison to the previously reported complexes with unsubstituted bipy ligands,14 the new compounds exhibit signicantly higher epoxide conversions. TOFs are around 1000 h-1 and the selectivity towards cyclooctene epoxide is close to 100%. The independence of TOFs from the ligand substituents and accordingly from the complex stability is exemplarily shown for the sterically feasible 5,5-substitution. This also shows that most presumably, the bipy ligand is not cleaved from the Mo centre during the reaction and the number of ligands is not reduced during the catalytic cycle. Hence, Mo complexes with more weakly coordinating bipyridine ligands could exhibit better catalytic activities. The mechanism previously published by Calhorda et al. 14a is in excellent agreement with the results Dalton Trans., 2009, 87468754 | 8751

View Online

presented here. In order to easily separate the catalyst from the reaction solution and reduce the negative effects of excess byproduct t-BuOH, the immobilisation of this type of catalyst is recommended and the related experiments are currently being undertaken in our laboratories. The linker to the carrier material, however, should be placed in the 4,4 or 5,5 positions of the 2,2-bipyridine ligand to avoid steric problems.

1051(s), 937(vs) (Mo=Osym ), 904(vs) (Mo=Oasym ). 1 H NMR (400 MHz, [d6 ]DMSO, 20 C): d = 2.41 (d, 6H, CH3 ), 8.64 (s, 2H, py-H6,6 ), 8.46 (m, J H,H = 7.2, 2H, py-H4,4 ), 8.1 (m, 2H, py-H3,3 ). Anal Calc. for C12 H12 Cl2 MoN2 O2 (383.19): C 37.62, H 3.13, N 7.31. Found: C 36.80, H 3.29, N 7.12. (4). [MoO2 Cl2 [MoO2 Cl2 (6,6-dimethyl-2,2-bipyridine)] (THF)2 ] (0.9 mmol). Yield: 0.15 g (93%). Color: white. Selected IR (KBr): n (cm-1 ) = 1599(vs), 1568(s), 1455(vs), 1031(s), 949(vs) (Mo=Osym ), 914(vs) (Mo=Oasym ). 1 H NMR (400 MHz, [d6 ]DMSO, 20 C): d = 1.98 (s, 6H, CH3 ), 7.59 (d, J H,H = 7.43, 2H, py-H3,3 ), 7.28 (t, J H,H = 8.49, 2H, py-H4,4 ), 6.76 (d, J H,H = 7.43, 2H, py-H5,5 ). 13 C NMR (400 MHz, [d6 ]DMSO, 20 C): d = 157.27 (py-C6,6 ), 154.04 (py-C2,2 ), 137.64 (py-C4,4 ), 123.53 (py-C5,5 ), 117.75 (py-C3,3 ), 23.94 (CH3 C). 95 Mo NMR (400 MHz, [d6 ]DMSO, 20 C): d = 191.16. Anal. Calc. for C12 H12 Cl2 MoN2 O2 (383.19): C 37.61, H 3.13, N 7.31. Found: C 36.84, H 3.21, N 7.16. (6). [MoO2 Cl2 [MoO2 Cl2 (4,4-dibromo-2,2-bipyridine)] (THF)2 ] (0.6 mmol). Yield: 0.144 g (46%). Color: off-white. Selected IR (KBr): n (cm-1 ) = 1589(vs), 1541(w), 1468(s), 1394(s), 1037(s), 942(vs) (Mo=Osym ), 912(vs) (Mo=Oasym ). 1 H NMR (400 MHz, CD3 NO2 , 20 C): d = 9.20 (d, J H,H = 6.7, 2H, py-H5,5 ), 8.60 (s, 2H, py-H3,3 ), 8.00 (d, J H,H = 6.7, 2H, py-H6,6 ). Anal. Calc. for C10 H6 Cl2 Br2 MoN2 O2 (512.93): C 23.4, H 1.2, N 5.45. Found: C 23.59, H 1.44, N 5.36. (7). [MoO2 Cl2 [MoO2 Cl2 (5,5-dibromo-2,2-bipyridine)] (THF)2 ](0.6 mmol). Yield: 0.119 g (38%). Color: off-white. Selected IR (KBr): n (cm-1 ) = 1585(s), 1562(w), 1460(s), 1376(s), 1037(s), 941(vs) (Mo=Osym ), 908(vs) (Mo=Oasym ). 1 H NMR (400 MHz, CD3 NO2 , 20 C): d = 9.44 (s, 2H, py-H6,6 ), 8.4 (d, J H,H = 8.4, 2H, py-H4,4 ), 8.13 (d, J H,H = 8.4, 2H, py-H3,3 ). Anal. Calc. for C10 H6 Cl2 Br2 MoN2 O2 (THF)0.5 (548.93): C 26.2, H 1.82, N 5.1. Found: C 25.30, H 1.96, N 5.15. [MoO2 Cl2 (5,5-diamino-2,2-bipyridine)] (9). The complex [MoO2 Cl2 (THF)2 ] (0.8 mmol) was dissolved in THF (5 mL) and treated with one equiv. of ligand that was also dissolved in THF (10 mL). The resulting turbid solution was stirred for 1 h: Yield: 0.20 g (84%). Color: dark red. Selected IR (KBr): n (cm-1 ) = 1617(s), 1491(s), 1384(vs), 1039(w), 935(s) (Mo=Osym ), 899(s) (Mo=Oasym ). 1 H NMR (400 MHz, [d6 ]DMSO, 20 C): d = 8.71 (s, 2H, py-H3,3 ), 8.06 (d, J H,H = 8.88, 2H, py-H6,6 ), 7.31 (d, J H,H = 8.63, 2H, py-H4,4 ), 6.26 (b, 4H, NH2 ). Anal. Calc. for C10 H10 Cl2 MoN4 O2 (383.19): C 31.18, H 2.59, N 14.54. Found: C 31.34, H 2.74, N 13.13. (10). [MoO2 Cl2 [MoO2 Cl2 (5,5-dinitro-2,2-bipyridine)] (THF)2 ] (0.2 mmol). Yield: 0.066 g (74%). Color: pale orange. Selected IR (KBr): n (cm-1 ) = 1609(s), 1531(vs), 1464(w), 1346(vs), 1051(s), 947(vs) (Mo=Osym ), 913(vs) (Mo=Oasym ). 1 H NMR (400 MHz, CD3 NO2 , 20 C): d = 10.23 (s, 2H, py-H6,6 ), 9.14 (dd, J H,H = 8.61, 2H, py-H3,3 ), 8.94 (d, J H,H = 8.75, 2H, py-H4,4 ). 13 C NMR(400 MHz, CD3 NO2 , 20 C): d = 153.60 (py-C2,2 ), 148.81 (py-C5,5 ), 148.11 (py-C6,6 ), 138.51 (py-C3,3 ), 127.72 (py-C4,4 ). 95 Mo NMR (400 MHz, CD3 NO2 , 20 C): d = 185.60 Anal. Calc. for C10 H6 Cl2 MoN4 O6 (THF)0.5 (481.13): C 29.92, H 2.07, N 11.6. Found: C 28.69, H 2.03, N 10.78. (11). [MoO2 Cl2 (5,5-di-ethoxycarbonyl-2,2-bipyridine)] [MoO2 Cl2 (THF)2 ] (1 mmol). Yield: 0.5 g (94%). Color: This journal is The Royal Society of Chemistry 2009

Experimental
General remarks All preparations and manipulations were carried out under an oxygen- and water-free argon atmosphere with standard Schlenk techniques. MoO2 Cl2 was purchased from Aldrich and used as received. Solvents were dried by standard procedures, distilled, and kept under argon over molecular sieves. (THF, and diethyl ether over Na/benzophenone; CH2 Cl2 over CaH2 ) The ligands 3,3dimethyl-2,2-bipyridine, 5,5-dimethyl-2,2-bipyridine, 6,6-dimethyl-2,2-bipyridine, 4,4-dibromo-2,2-bipyridine, 5,5-dibromo-2,2-bipyridine, 5,5-diamino-2,2-bipyridine; 5,5-dinitro2,2-bipyridine; 5,5-di-ethoxycarbonyl-2,2-bipyridine; 6-phenyl2,2-bipyridine; 2,2:6,2-terpyridine were prepared according to published procedures.18 Elemental analyses were performed with a Flash EA 1112 series elemental analyser. 1 H, 13 C NMR and 95 Mo NMR were measured in DMSO or CD3 NO2 with a 400 MHz Bruker Avance DPX-400 and Varian 400 spectrometer. IR spectra were recorded with a Perkin Elmer FT-IR spectrometer using KBr pellets as the IR matrix. Catalytic runs were monitored by GC methods on a Varian CP-3800 instrument equipped with a FID and a VF-5ms column. Thermogravimetric analyses were performed using a Netzsch TG209 system at a heating rate of 10 K min-1 under argon. Warning: TBHP (in decane solution) is toxic, possibly mutagenic, corrosive and a strong oxidising agent. It is a combustible liquid and is readily absorbed through the skin and must be stored below 38 C. Syntheses of the complexes The complex [MoO2 Cl2 (THF)2 ] was dissolved in CH2 Cl2 (5 mL) and treated with one equiv. of ligand dissolved in CH2 Cl2 (10 mL). The resulting turbid solutions were stirred for 1 h, with the solvent then being removed under vacuum, and the product being washed with diethyl ether (2 5 mL) and dried under vacuum. (1). [MoO2 Cl2 [MoO2 Cl2 (3,3-dimethyl-2,2-bipyridine)] (THF)2 ] (1 mmol). Yield: 0.48 g (89%). Color: off-white. Selected IR (KBr): n (cm-1 ) = 1653(s), 1583(s), 1559(w), 1457(s), 1409(w), 1039(w), 938(vs) (Mo=Osym ), 910(vs) (Mo=Oasym ). 1 H NMR (400 MHz, [d6 ]DMSO, 20 C): d = 1.89 (s, 6H, CH3 ), 8.34 (d, J H,H = 4.8 Hz, 2H, py-H6,6 ), 7.74 (d, J H,H = 8.0 Hz, 2H, py-H4,4 ), 7.32 (dd, J H,H = 4.8, 2H, py-H5,5 ). 13 C NMR (400 MHz, [d6 ]DMSO, 20 C): d = 153.99 (py-C2,2 ), 144.96 (py-C6,6 ), 140.31 (py-C4,4 ), 132.57 (py-C3,3 ), 124.10 (py-C5,5 ), 17.76 (CH3 C). 95 Mo NMR (400 MHz, [d6 ]DMSO, 20 C): d =194. Anal. Calc. for C12 H12 Cl2 MoN2 O2 (383.19): C 37.61, H 3.13, N 7.31. Found: C 35.34, H 3.21, N 6.77. (3). [MoO2 Cl2 [MoO2 Cl2 (5,5-dimethyl-2,2-bipyridine)] (THF)2 ] (1 mmol). Yield: 0.24 g (64%). Color: pale yellow. Selected IR (KBr): n (cm-1 ) = 1600(s), 1586(w), 1483(s), 1386(s),
8752 | Dalton Trans., 2009, 87468754

Downloaded by Universiteit Gent on 05 November 2010 Published on 28 August 2009 on http://pubs.rsc.org | doi:10.1039/B909075D

View Online

white. Selected IR (KBr): n (cm-1 ) = 1730(vs), 1607(s), 1471(w), 1059(w), 945(vs) (Mo=Osym ), 911(vs) (Mo=Oasym ). 1 H NMR (400 MHz, CD3 NO2 , 20 C): d = 9.96 (s, 2H, py-H3,3 ), 8.88 (dd, J H,H = 8.59, 2H, py-H6,6 ), 8.73 (d, J H,H = 8.04, 2H, py-H4,4 ), 4.51 (q, 4H, CH2 ), 1.44 (t, J H,H = 7.92, 6H, CH3 ). 13 C NMR(400 MHz, CD3 NO2 , 20 C): d = 164.88 (C=O), 154.71 (py-C2,2 ), 153.34 (py-C6,6 ), 143.79 (py-C4,4 ), 132.20 (py-C5,5 ), 126.05 (py-C3,3 ), 69.16 (CH2 ), 14.62 (CH3 ). 95 Mo NMR (400 MHz, CD3 NO2 , 20 C): d = 190.58. Anal. Calc. for C16 H16 Cl2 MoN2 O6 (499.94): C 38.5, H 3.20, N 5.61. Found C 38.65, H 3.27, N 5.18. [MoO2 Cl2 (6-phenyl-2,2-bipyridine)] (12). [MoO2 Cl2 (THF)2 ] (0.4 mmol). Yield: 0.17 g (94%). Color: yellow. Selected IR (KBr): n (cm-1 ) = 1602(s), 1596(s), 1488(s), 1446(vs), 1023(vs), 946(vs) (Mo=Osym ), 910(vs) (Mo=Oasym ). 1 H NMR (400 MHz, [d6 ]DMSO, 20 C). d = 8.74 (s, 1H, py-H6 ), 8.62 (d, J HH = 8.01 Hz, 1H, py-H6 ), 8.08 (s, 1H, py-H4 ), 8.37 (s, 1H, py-H5 ), 8.25 (d, J HH = 8.01 Hz, 2H, py-H3,3 ), 8.04 (s, 1H, py-H4,4 ), 7.54 (m, 3H, py-H3 , py-H4 , py-H5 ), 7.48 (d, J H,H = 6.46 Hz, 2H, py-H5 ). 13 C NMR(400 MHz, [d6 ]DMSO, 20 C): d = 155.62 (py-C6 ), 154.37 (py-C2,2 ), 153.9 (pyC6 ), 149.4 (py-C4 ), 148.5 (benzyl-C1 ), 138.5 (py-C4 ), 129.3 (benzylC3 ,C5 ), 127.8 (benzyl-C1 ,C6 ), 126.7 (benzyl-C4 ), 126.0 (py-C5 ), 124.6 (py-C3,3 ), 121.2 (py-C5 ). 95 Mo NMR (400 MHz, [d6 ]DMSO, 20 C): d = 197. Anal. Calc. for C12 H6 Cl2 MoN2 O4 (409.88): C 44.56, H 2.78, N 6.49. Found: C 44.96, H 2.94, N 6.48. (13). [MoO2 Cl2 (THF)2 ] [MoO2 Cl2 (2,2:6,2-terpyridine)] (0.7 mmol). Yield: 0.23 g (72%). Color: yellow. Selected IR (KBr): n (cm-1 ) = 1586(s), 1445(vs), 1427(vs), 1024(s), 944(vs) (Mo=Osym ), 910(vs) (Mo=Oasym ). 1 H NMR (400 MHz, CD3 NO2 , 20 C): d = 9.04 (d, J H,H = 5.29, 2H, py-H6,6 ), 8.84 (d, J H,H = 7.86, 2H, py-H3,3 ), 8.71 (s, 2H, py-H3 , py-H5 ), 8.65 (d, J H,H = 8.82, 2H, py-H5,5 ), 8.44 (t, J H,H = 9.39, 1H, py-H4 ), 8.14 (s, 2H, py-H4,4 ). Anal. Calc. for C15 H11 Cl2 MoN3 O2 (432.22): C 41.68, H 2.54, N 9.72. Found: C 41.40, H 2.75, N 9.39. Catalytic reactions The epoxidation of cyclooctene was carried out at 55 C in air inside a batch reactor equipped with a magnetic stirrer. The reactor was loaded with 73 mmol of complex, 7.3 mmol of cyclooctene (1% molar ratio of catalyst/substrate) and 11 mmol tert-butyl hydroperoxide (5.5 M in decane) as oxidant. Samples were withdrawn periodically and analysed with a gas chromatograph (Varian 3800) equipped with a capillary column (SPB-5, 20 m 0.25 mm 0.25 mm) and a ame ionisation detector. The products were identied by gas chromatography-mass spectrometry (HP, 5890 Series II GC; HP 5970 Series Mass Selective Detector) using He as carrier gas. X-ray structure analysis of compound 11 Crystal data and details of the structure determination: formula: C16 H16 Cl2 MoN2 O6 ; M r = 499.15; crystal color and shape: colourless; crystal dimensions = 0.10 0.25 0.25 mm; crystal system: triclinic; space group P1;(no. 2); a = 8.4974(3), b = ; a = 100.241(2) , b = 95.708(2) , 10.5235(4), c = 11.0314(4) A 3 ; Z = 2; m(MoKa ) = 1.017 mm-1 ; g = 101.873(2) , V =940.45(6) A -3 rcalcd = 1.763 g cm ; q range = 3.0425.35 ; data collected: 28 892; independent data [I o > 2s (I o )/all data/Rint ]: 3207/3340/0.063; data/restraints/parameters: 3340/0/246; R1 [I o > 2s (I o )/all This journal is The Royal Society of Chemistry 2009

data]: 0.0250/0.0260; wR2 [I o >2s (I o )/all data]: 0.0639/0.0652; -3 . Suitable single crystals GOF = 1.081; Drmax/min : 0.97/-0.36 e A for the X-ray diffraction study were grown at room temperature by slow diffusion of hexane of complex 11 in acetone. Preliminary examination and data collection were carried out on an area detecting system (APEX II, k -CCD) at the window of a rotating anode (Bruker AXS, FR591) and graphite monochromated MoKa ). Data collection was performed radiation (l = 0.71073 A at 173 K. Raw data were corrected for Lorentz, polarisation, and, arising from the scaling procedure, for latent decay and absorption effects. All non-hydrogen atoms were rened with anisotropic displacement parameters. All hydrogen atom positions were calculated in ideal positions (riding model).25 Computational details All calculations were performed with GAUSSIAN-0319 using the density functional/HartreeFock hybrid model Becke3LYP20 and the split valence double-z (DZ) basis set 6-31G*,21 for Mo the Stuttgart-RSC ECP22 was applied. No symmetry or internal coordinate constraints were applied during optimisations. All reported structures were veried as being true minima by the absence of negative eigenvalues in the vibrational frequency analysis. Approximate free energies (DG) and enthalpies (DH ) were obtained through thermochemical analysis of frequency calculations, using the thermal correction to Gibbs free energy as reported by GAUSSIAN-03. This takes into account zeropoint effects, thermal enthalpy corrections, and entropy. All energies reported in this paper, unless otherwise noted, are free energies or enthalpies at 298 K, using unscaled frequencies. NMR chemical shifts were calculated via the GIAO formalism23 (gauge independent atomic orbital) as implemented in GAUSSIAN03. The GIAO calculations were carried out as single points with the triple-zeta basis set 6-311++G**24 including the already mentioned Stuttgart-ECP for Mo. The 13 C chemical shifts are scaled to a tetramethylsilane (TMS) calculation with the same level of theory.

Downloaded by Universiteit Gent on 05 November 2010 Published on 28 August 2009 on http://pubs.rsc.org | doi:10.1039/B909075D

Acknowledgements
A. G. and F. E. K. thank the Elitenetzwerk Bayern (Graduate School NanoCat) for nancial support. M. D. Z. is grateful to the international Graduate School for Science and Engineering (IGSSE) for a Ph.D. Grant. The authors thank Dr. Hugh ChaffeyMillar and Dr. Mirza Cokoja for valuable discussions and for the proofreading of the manuscript.

Notes and references


1 K. Weissermel and H.-J. Arpe, Industrial Organic Chemistry, Fourth Edition, Wiley-VCH, Germany, 2003, 146. 2 (a) F. E. Kuhn, A. M. Santos and M. Abrantes, Chem. Rev., 2006, 106, and W. A. Herrmann, Chem. Rev., 2455; (b) C. C. Rom ao, F. E. Kuhn 1997, 97, 3197. 3 (a) K. A. Jrgensen, Chem. Rev., 1989, 89, 431; (b) M. H. Dickman and M. T. Pope, Chem. Rev., 1994, 94, 569. 4 G.-S. Kim, D. Huffman and C. W. De Kock, Inorg. Chem., 1989, 28, 1279. 5 M. H. Chisholm, K. Folting, J. C. Huffman and C. C. Kirkpatrick, Inorg. Chem., 1984, 23, 1021. Petrovski, M. Pillinger, A. A. Valente, I. S. Gonc 6 Z. alves, A. Hazell and C. C. Rom ao, J. Mol. Catal. A: Chem., 2005, 227, 67.

Dalton Trans., 2009, 87468754 | 8753

View Online

7 (a) F. E. Kuhn, E. Herdtweck, J. J. Haider, W. A. Herrmann, I. S. Gonc alves, A. D. Lopes and C. C. Rom ao, J. Organomet. Chem., 1999, 583, 3; (b) F. E. Kuhn, A. M. Santos, A. D. Lopes, I. S. Gonc alves, E. Herdtweck and C. C. Rom ao, J. Mol. Catal. A: Chem., 2000, 164, 25; (c) F. E. Kuhn, A. M. Santos, I. S. Gonc alves, C. C. Rom ao and A. D. Lopes, Appl. Organomet. Chem., 2001, 15, 43. 8 (a) W. A. Herrmann, J. J. Haider, J. Fridgen, G. M. Lobmaier and M. Spiegler, J. Organomet. Chem., 2000, 603, 69; (b) S. BelleminLaponnaz, K. S. Coleman, P. Dierkes, J.-P. Massonand and J. A. Osborn, Eur. J. Inorg. Chem., 2000, 1645; (c) F. E. Kuhn, A. M. Santos, A. D. Lopes, I. S. Gonc alves, J. E. Rodr guez-Borges, M. Pillinger and C. C. Rom ao, J. Organomet. Chem., 2001, 621, 207; (d) I. S. Gonc alves, F. E. Kuhn, A. M. Santos, A. D. Lopes, J. E. Rodr guez-Borges, M. Pillinger, P. Ferreira, J. Rocha and C. C. Rom ao, J. Organomet. Chem., 2001, 626, 1; (e) A. A. Valente, I. S. Gonc alves, A. D. Lopes, J. E. Rodr guez-Borges, M. Pillinger, C. C. Rom ao, J. Rocha and X. Garc aMera, New J. Chem., 2001, 25, 959. 9 A. M. Santos, F. E. Kuhn, K. Bruus-Jensen, I. Lucas, C. C. Rom ao and E. Herdtweck, J. Chem. Soc., Dalton Trans., 2001, 1332. 10 J.-M. Br` egeault, Dalton Trans., 2003, 3289. 11 S. Gago, J. E. Rodr guez-Borges, C. Teixeira, A. M. Santos, J. Zhao, M. Petrovski, T. M. Santos, F. E. Kuhn, Pillinger, C. D. Nunes, Z. C. C. Rom ao and I. S. Gonc alves, J. Mol. Catal. A: Chem., 2005, 236, 1. 12 S. Gago, M. Pillinger, A. A. Valente, T. M. Santos, J. Rocha and I. S. Gonc alves, Inorg. Chem., 2004, 43, 5422. 13 G. Barea, A. Lledos, F. Maseras and Y. Jean, Inorg. Chem., 1998, 37(13), 3321. Bencze, E. Herdtweck, A. Prazeres, 14 (a) F. E. Kuhn, M. Groarke, E. A. M. Santos, M. J. Calhorda, C. C. Rom ao, I. S. Gonc alves, A. D. Lopes and M. Pillinger, Chem.Eur. J., 2002, 8, 2370; (b) M. Groarke, I. S. Gonc alves, W. A. Herrmann and F. E. Kuhn, J. Organomet. Chem., 2002, 649, 108; (c) A. M. Al-Ajlouni, A. A. Valente, C. D. Nunes, M. Pillinger, T. M. Santos, J. Zhao, C. C. Rom ao, I. S. Gonc laves and F. E. Kuhn, Eur. J. Inorg. Chem., 2005, 1716; (d) L. F. Veiros, A. Prazeres, P. J. Costa, C. C. Rom ao, F. E. Kuhn and M. J. Calhorda, Dalton Trans., 2006, 1383; (e) J. M. Mitchell and N. S. Finney, J. Am. Chem. Soc., 2001, 123, 862. 15 F. E. Kuhn, A. D. Lopes, A. M. Santos, E. Hertdweck, J. J. Haider, C. C. Rom ao and A. G. Santos, J. Mol. Catal. A: Chem., 2000, 151, 147. 16 J. Zhao, K. R. Jain, E. Herdtweck and F. E. Kuhn, Dalton Trans., 2007, 5567. 17 A. M. Al-Ajlouni, A. Gunyar, M.-D. Zhou, P. N. W. Baxter and F. E. Kuhn, Eur. J. Inorg. Chem., 2009, 1019. 18 (a) F. H. Case, J. Am. Chem. Soc., 1946, 68, 2574; C. P. Whittle, J. Heterocycl. Chem., 1977, 14, 191; (b) E. Rajalakshmanan and V. Alexander, Synth. Commun., 2005, 35, 891; (c) M. Tiecco, L. Testaferri, M. Tingoli, D. Chianelli and M. Montanucci, Synthesis, 1984, 736; (d) P. F. H. Schwab, F. Fleischer and J. Michl, J. Org. Chem., 2002, 67,

19

20

21 22 23

24 25

443; (e) G. Maerker and F. H. Case, J. Am. Chem. Soc., 1958, 80, 2745; (f) J. E. Parks, B. E. Wagner and R. H. Holm, J. Organomet. Chem., 1973, 56, 53. M. J. Frisch, G. W. Trucks, H. B. Schlegel, G. E. Scuseria, M. A. Robb, J. R. Cheeseman, J. Montgomery, J. A. T. Vreven, K. N. Kudin, J. C. Burant, J. M. Millam, S. S. Iyengar, J. Tomasi, V. Barone, B. Mennucci, M. Cossi, G. Scalmani, N. Rega, G. A. Petersson, H. Nakatsuji, M. Hada, M. Ehara, K. Toyota, R. Fukuda, J. Hasegawa, M. Ishida, T. Nakajima, Y. Honda, O. Kitao, H. Nakai, M. Klene, X. Li, J. E. Knox, H. P. Hratchian, J. B. Cross, V. Bakken, C. Adamo, J. Jaramillo, R. Gomperts, R. E. Stratmann, O. Yazyev, A. J. Austin, R. Cammi, C. Pomelli, J. W. Ochterski, P. Y. Ayala, K. Morokuma, G. A. Voth, P. Salvador, J. J. Dannenberg, V. G. Zakrzewski, S. Dapprich, A. D. Daniels, M. C. Strain, O. Farkas, D. K. Malick, A. D. Rabuck, K. Raghavachari, J. B. Foresman, J. V. Ortiz, Q. Cui, A. G. Baboul, S. Clifford, J. Cioslowski, B. B. Stefanov, G. Liu, A. Liashenko, P. Piskorz, I. Komaromi, R. L. Martin, D. J. Fox, T. Keith, M. A. Al-Laham, C. Y. Peng, A. Nanayakkara, M. Challacombe, P. M. W. Gill, B. Johnson, W. Chen, M. W. Wong, C. Gonzalez, J. A. Pople, Gaussian03, Rev. C.02 Ed., Gaussian Inc., Wallingford, CT, 2004. (a) S. H. Vosko, L. Wilk and M. Nusair, Can. J. Phys., 1980, 58, 1200; (b) C. Lee, W. Yang and R. G. Parr, Phys. Rev. B: Condens. Matter Mater. Phys., 1988, 37, 785; (c) A. D. Becke, J. Chem. Phys., 1993, 98, 5648. (a) W. J. Hehre, R. Ditcheld and J. A. Pople, J. Chem. Phys., 1972, 56, 2257; (b) M. M. Francl, W. J. Petro, W. J. Hehre, J. S. Binkley, M. S. Gordon, D. J. DeFrees and J. A. Pople, J. Chem. Phys., 1982, 77, 3654. A. Bergner, M. Dolg, W. Kuechle, H. Stoll and H. Preuss, Mol. Phys., 1993, 80, 1431. (a) R. McWeeny, Phys. Rev., 1962, 126, 1028; (b) R. Ditcheld, Mol. Phys., 1974, 27, 789; (c) J. L. Dodds, R. McWeeny and A. J. Sadlej, Mol. Phys., 1977, 34, 1779; (d) K. Wolinski, J. F. Hinton and P. Pulay, J. Am. Chem. Soc., 1990, 112, 8251. R. Krishnan, J. S. Binkley, R. Seeger and J. A. Pople, J. Chem. Phys., 1980, 72, 650. (a) APEX suite of crystallographic software, APEX 2 Version 2008, 4.Bruker AXS Inc., Madison, Wisconsin, USA, (2008); (b) SAINT, Version 7.56a and SADABS Version 2008/1, Bruker AXS Inc., Madison Wisconsin, USA, (2008); (c) A. Altomare, G. Cascarano, C. Giacovazzo, A. Guagliardi, M. C. Burla, G. Polidori and M. Camalli, SIR 92, J. Appl. Crystallogr., 1994, 27, 435; (d) International Tables for Crystallography, Vol. C, Tables 6.1.1.4, 500502; 4.2.6.8, 219222; 4.2.4.2, 193199, A. J. C. Wilson, Ed., Kluwer Academic Publishers, Dordrecht, The Netherlands, 1992; (e) G. M. Sheldrick, SHELXL-97, University of G ottingen, Gottingen, Germany, (1998); (f) A. L. Spek, PLATON, A Multipurpose Crystallographic Tool, Utrecht University, Utrecht, The Netherlands, (2008); (g) L. J. Farrugia, WinGX (Version 1.70.01 January 2005), J. Appl. Crystallogr., 1999, 32, 837.

Downloaded by Universiteit Gent on 05 November 2010 Published on 28 August 2009 on http://pubs.rsc.org | doi:10.1039/B909075D

8754 | Dalton Trans., 2009, 87468754

This journal is The Royal Society of Chemistry 2009

Vous aimerez peut-être aussi