Dioxotungsten(VI) complexes with isoniazid-related hydrazones as (pre)catalysts for olefin epoxidation: solvent and ligand substituent effects

Višnja Vrdoljak*a, Jana Piskabc, Biserka Prugovečkia, Dominique Agustin*bcd, Predrag Novaka and Dubravka Matković-Čalogovića
aUniversity of Zagreb, Faculty of Science, Department of Chemistry, Horvatovac, 102a, 10000 Zagreb, Croatia. E-mail: visnja.vrdoljak@chem.pmf.hr; Fax: +385 1 4606341; Tel: +385 1 4606353
bUniversité de Toulouse, Institut Universitaire de Technologie Paul Sabatier, Département de Chimie, Av. Georges Pompidou, CS 20258, F-81104 Castres Cedex, France. E-mail: dominique.agustin@iut-tlse3.fr; Fax: +33 5 63 351 910; Tel: +33 5 63 621 172
cCNRS, LCC (Laboratoire de Chimie de Coordination), 205 route de Narbonne, BP 44099, F-31077 Toulouse Cedex 4, France
dUniversité de Toulouse, UPS, INPT, F-31077 Toulouse Cedex 4, France

Received 25th February 2016 , Accepted 5th April 2016

First published on 6th April 2016


Abstract

The mononuclear dioxotungsten(VI) complexes [WO2(L3OMe)(D)] (1a and 1b), [WO2(L4OMe)(D)] (2a and 2b) and [WO2(LH)(D)] (3a and 3b) (D = EtOH (1a–3a) or MeOH (1b–3b); L3OMe = 3-methoxy-2-oxybenzaldehyde isonicotinoyl hydrazonato, L4OMe = 4-methoxy-2-oxybenzaldehyde isonicotinoyl hydrazonato, LH = 2-oxybenzaldehyde isonicotinoyl hydrazonato) were synthesized by the reaction of [WO2(acac)2]·0.5C6H5Me with the respective isoniazid-related hydrazone. The compounds were characterized by microanalysis, FT-IR and NMR spectroscopy, thermogravimetric analysis, and powder X-ray diffraction method. The crystal and molecular structures of 1a, 1b, 3a and [WO2(acac)2]·0.5C6H5Me were determined by single crystal X-ray diffraction. The structures of 1a, 1b, 3a are mononuclear and form hydrogen bonded centrosymmetric dimers. In all three complexes, the dimers are also held together by π⋯π interactions between aromatic rings. The catalytic performances (activity and selectivity) of 1a–3a and 1b–3b towards alkene epoxidation by tert-butyl hydroperoxide (TBHP) were investigated under different conditions.


Introduction

The chemistry of hydrazones receives ongoing attention in various fields. Thus, isoniazid-related aroylhydrazones and their coordination compounds are of particular interest because of their biological activities1 and their structural diversity.2 Although various metallosupramolecular assemblies2 as well as mononuclear3 dioxomolybdenum(VI) complexes with isoniazid-based aroylhydrazone derivatives are known, related dioxotungsten(VI) complexes are rare. To the best of our knowledge, only the dioxotungsten(VI) compounds [WO2(NIH)]n and [WO2(VIH)]4 (where NIH2− = 2-oxy-1-naphthaldehyde isonicotinoyl hydrazonato and VIH2− = 3-methoxy-2-oxybenzaldehyde isonicotinoyl hydrazonato) were structurally characterized by the powder X-ray diffraction method (PXRD).2a

Compounds with active metal centres play an important role as functional materials in diverse fields.4 In this context, metal complexes with coordinatively unsaturated metal ions as well as those ligated by weakly coordinating ligands are of particular interest. In addition to having other functions, they also act as catalysts for the epoxidation of alkenes, an important process of wide chemical and biochemical significance. A variety of complexes has been used in this way and those containing the {MO2}2+ unit are known as highly selective catalysts.5–7

The catalytic conversion of olefins to epoxides may involve oxidants such as tert-butylhydroperoxide (TBHP),3b,5a–5c,7 O2,8 or H2O2,5d,5e usually in organic solvents. In accordance with the requirements of “green” chemistry, these reactions have also been performed under organic solvent-free conditions. However, contrary to dioxomolybdenum(VI),3b,5a,5c,6,7b–7d investigations of solvent-free epoxidations involving dioxotungsten(VI) complexes are quite scarce. Examples are [WO2(L′)(solv)] (solv = alcohol, L′ = Schiff base)3b,9 [WO2L] (L = aminobisphenolato ligand)10 and [WO2(salen)] (salen = bis(salicylidenediamine)) complexes.11 We have previously compared the catalytic activity of dioxomolybdenum(VI) and dioxotungsten(VI) complexes with ligands derived from 4-hydroxy-benzhydrazone.12 It was shown that the MoVI complexes are more efficient catalysts for cyclooctene epoxidation by tert-butyl hydroperoxide (TBHP) under organic solvent-free conditions.

Considering these features, we have decided to investigate the influence of the coordinated ligands (solvent (D) and isoniazid-related hydrazones (LR)) on the solid-state structure and catalytic activity of the [WO2(LR)(D)] complexes. To achieve this aim, the investigation included a systematic variation of: (i) the tridentate ligand (2-hydroxybenzaldehyde isonicotinoyl hydrazone, H2LH, and two methoxy-substituted derivatives, i.e. 3-methoxy-2-hydroxybenzaldehyde isonicotinoyl hydrazone, H2L3OMe, and 4-methoxy-2-hydroxybenzaldehyde isonicotinoyl hydrazone, H2L4OMe, see Scheme 1), and (ii) the solvent D (MeOH and EtOH). The mononuclear WVI complexes [WO2(LR)(EtOH)] (1a–3a) and [WO2(LR)(MeOH)] (1b–3b), where R = 3OMe, 4OMe or H, have been investigated as (pre)catalysts for the epoxidation of cyclooctene in the presence of organic solvent as well as in its absence, using TBHP (in decane or in water) as an oxidant.


image file: c6ra05067k-s1.tif
Scheme 1 The isoniazid-based hydrazones H2LR (H2L3OMe, H2L4OMe and H2LH). The substituent is shown in red colour.

Results and discussion

Synthesis of bis(acetylacetonato)dioxotungsten(VI) toluene solvate (2/1), [WO2(acac)2]·0.5C6H5Me

The synthesis of the starting tungsten precursor [WO2(acac)2]·0.5C6H5Me was performed from freshly prepared WO2Cl2 and an excess of dry acetylacetone in toluene. Crystals suitable for X-ray structure analysis were obtained directly from the reaction mixture. The product can be recrystallized from a toluene/acetylacetone mixture. The previously reported pseudopolymorph was identified as [WO2(acac)2].13 Surprisingly, the crystal and molecular structure of this compound has not been reported previously.

Synthesis of dioxotungsten(VI) complexes chelated by ONO tridentate ligands

Compound [WO2(acac)2]·0.5C6H5Me was used as a precursor for the synthesis of dioxotungsten(VI) complexes of the type [WO2(LR)(D)], Scheme 2, with isoniazid-related hydrazone ligands. The products were obtained by the ligand exchange strategy using the corresponding aroylhydrazone ligand in ethanol (1a–3a) or methanol (1b–3b). The reactions were performed under a dry argon atmosphere.
image file: c6ra05067k-s2.tif
Scheme 2 Mononuclear [WO2(LR)(D)] complexes (D = MeOH or EtOH). The R′ substituent is shown in red colour in Scheme 1.

Synthesis in the presence of a small amount of DMF proved to be effective for producing products in crystalline form. Good quality single crystals of 1a, 1b and 3a suitable for X-ray diffraction analysis were obtained in this way. Otherwise, in the absence of DMF, yellow powdered solids were obtained.

Thermal analysis

The thermal stability of the complexes was investigated under an oxygen atmosphere. The mass loss corresponding to the first step in the TG curves of 1a–3a, 1b–3b was related to the loss of the coordinated ethanol or methanol molecule. This process occurred in the range 135–186 °C (1a), 159–210 °C (1b), 141–197 °C (2a), 154–193 °C (2b), 190–234 °C (3a) and 175–237 °C (2b). These temperatures are generally much higher than those required for alcohol removal from similar dioxomolybdenum(VI) complexes, for instance 110–120 °C for EtOH loss from a salicylidene-2-aminophenolato derivative,14 indicating that the alcohol donor forms a stronger bond with W relative to Mo. The same trend was observed in a previous report.12 All remaining unsolvated products started to decompose at 298 °C for 1a and 1b, 308 °C for 2a and 2b, and 295 °C for 3a and 3b. The final residue was identified as WO3.

The first step in the thermogravimetric curve of [WO2(acac)2]·0.5C6H5Me was related to the loss of toluene (65–104 °C) and was followed, upon further heating, by significant weight loss in the range 128–449 °C due to ligand decomposition.

Crystal and molecular structure of [WO2(acac)2]·0.5C6H5Me

In [WO2(acac)2]·0.5C6H5Me the tungsten atom exhibits the distorted octahedral coordination sphere. It is coordinated by two cis oxo oxygen atoms and by four oxygen atoms from the two acetylacetonate ligands (Fig. 1). The asymmetric unit consists of the complex molecule and one half of the toluene molecule of crystallization (Fig. S1, see ESI).
image file: c6ra05067k-f1.tif
Fig. 1 ORTEP drawing of [WO2(acac)2]·0.5C6H5Me with the atom-labelling scheme (displacement ellipsoids of non-hydrogen atoms are drawn at the 50% probability level). Toluene molecule is omitted for clarity.

The Cambridge Structural Database15 was searched for structures with the XO2 unit (X was any element with the coordination number of 6: two oxygen atoms from two dioxo groups and four oxygen atoms from two acac ligands, including substituted acac ligands). The search gave only 9 hits (6 Mo complexes, 2 U compounds and a W complex). Since the data for the W complex (having 1,3-phenyl substituted acacetylacetonate) include only the unit cell and lack the atomic coordinates, a comparison of bonds and angles with the present structure is not possible. The bond distances and angles around the W atom and within the chelate rings are given in Table S1, see ESI. There are no classical hydrogen bonds in the crystal structure, only a weak C–H⋯O interaction between the toluene molecule and an oxo oxygen atom (d(C15⋯O5[−1/2−x, 1/2 + y, 1/2−z]) = 3.337(12) Å).

Crystal and molecular structure of [WO2(LR)(D)]

In all compounds, the tungsten atoms exhibit the distorted octahedral coordination sphere. The ligand is coordinated in the dianionic form to the cis-{WO2}2+ core via the ONO donor atoms. The remaining sixth coordination site is occupied by the oxygen atom from the solvent molecule (ethanol in 1a and 3a, methanol in 1b). The ORTEP plots of the molecular structures of 1a, 1b and 3a are given in the ESI (Fig. S2, see ESI).

The distance from the W atom to the O donor atom of the solvent molecule in 1a, 1b and 3a represents the longest bond within the distorted octahedron (Table 1). The C7–N1 bond of the L3OMe ligand in complexes 1a and 1b is 1.275(4) and 1.284(5) Å, respectively and is not significantly different than in the free ligands (1.2791(18) Å in H2L3OMe and 1.276(2) Å in H2L3OMe·H2O).16 The N2–C1 bond length is shortened in the complexes 1a and 1b (1.290(5) Å in 1a and 1.270(5) Å in 1b) in comparison to the free ligands (1.3552(18) Å and 1.339(2) Å in H2L3OMe and H2L3OMe·H2O, respectively), while the N1–N2 bond is lengthened (1.401(4) Å in 1a and 1.395(3) Å in 1b) in comparison to the free ligands (1.3791(19) Å and 1.3724(15) Å in H2L3OMe and H2L3OMe·H2O, respectively), due to electron delocalization.

Table 1 Selected bond lengths (Å) and angles (°) for compounds 1a, 1b and 3a
  1a 1b 3a
W–O1 2.001(2) 2.000(3) 1.997(2)
W–O2 1.922(2) 1.927(3) 1.927(3)
W–O3 1.706(2) 1.704(3) 1.704(2)
W–O4 1.718(2) 1.721(3) 1.719(2)
W–O5 2.292(2) 2.327(3) 2.306(3)
W–N1 2.249(2) 2.217(3) 2.248(3)
C7–N1 1.275(4) 1.284(5) 1.282(4)
N1–N2 1.401(4) 1.395(4) 1.399(4)
N2–C1 1.290(5) 1.270(5) 1.291(4)
O1–W–O2 149.17(9) 150.99(11) 148.58(9)
O1–W–O3 97.60(10) 96.44(14) 97.96(11)
O1–W–O4 97.71(10) 98.32(13) 98.93(11)
O1–W–O5 78.35(9) 80.63(11) 79.48(9)
O1–W–N1 71.81(9) 72.09(11) 71.91(9)
O2–W–O3 98.56(10) 98.56(14) 99.16(11)
O2–W–O4 103.40(10) 101.59(13) 102.06(11)
O2–W–O5 80.73(9) 80.73(12) 78.98(10)
O2–W–N1 81.42(9) 81.72(11) 81.17(9)
O3–W–O4 104.48(11) 104.90(14) 104.37(12)
O3–W–O5 168.71(10) 171.06(12) 169.69(11)
O3–W–N1 92.41(10) 96.10(12) 92.58(10)
O4–W–O5 86.59(9) 83.92(12) 85.92(10)
O4–W–N1 161.35(10) 157.88(13) 161.87(10)
O5–W–N1 76.32(8) 74.97(11) 77.13(9)


The C7–N1 bond length in 3a is 1.282(4) Å and is not significantly different than in the free H2LH ligand, 1.2756(16) Å.17 The N2–C1 bond is shortened in 3a (1.291(4) Å) in comparison to the free H2LH ligand (1.3558(17) Å), whereas the N1–N2 bond is lengthened in the complexes (1.399(4) Å) in comparison to the free H2LH ligand (1.3699(15) Å) due to the electron delocalization.

In 1a, 1b and 3a the ligand is not planar with the largest deviation from planarity being that of the phenyl and the pyridyl moieties in 3a (φ = 8.94(12)°, Table S2, see ESI). The largest dihedral angle between the planes of the five- and six-membered chelate rings is also observed in 3a (ψ = 16.87(16)°, Table S2).

A common feature of 1a, 1b and 3a is the association of the molecules by hydrogen bonds (O5–H50⋯N3) into centrosymmetric dimers, graph set R22(18), Fig. 2. The shortest hydrogen bond distance is in complex 1b, d(O5⋯N3[−x, −y, 1−z]) = 2.692(4) Å (Table S3, see ESI).


image file: c6ra05067k-f2.tif
Fig. 2 Discrete dimer formed through intermolecular O5–H50⋯N3[−x, −y, 1−z] hydrogen bonds in 1b. Hydrogen bonds are shown by dotted lines.

In all three complexes (1a, 1b and 3a), the dimers are also held together by π⋯π interactions between the two pyridyl rings. Neighboring dimer molecules are further connected by π⋯π interactions between: the five-membered chelate rings and phenyl rings in 1a, between the five-membered chelate rings and pyridyl rings in 1b, and between the two phenyl rings in 3a (Fig. 3 and Table S4, see ESI).


image file: c6ra05067k-f3.tif
Fig. 3 Partial structural motif of endless chains in 1a, 1b and 3a. Hydrogen-bonds are shown by blue dotted lines. Green dashed lines indicate π⋯π interactions (values are given in Å). The centroids of the five-membered chelate rings are shown as yellow spheres, the centroids of pyridyl rings as green spheres and the centroids of the phenyl rings as red spheres.

Complex 1b is isostructural with the dioxomolybdenum(VI) complex [MoO2(L3OMe)(MeOH)] (CSD code KANYIJ18 and KANYIJ01).19 The powder X-ray diffraction patterns of 2a and 2b indicate that they are isostructural (Fig. S3). The dioxotungsten(VI) complex 3a is isostructural with the corresponding dioxomolybdenum(VI) complexes, [MoO2(LH)(EtOH)] (CSD code UVAQOY)2b and [MoO2(LH)(MeOH)] (CSD code PAGCEF,20 PAGCEF01).21 PXRD of 3a and 3b indicate that they are also isostructural (Fig. S3).

IR spectroscopy

The IR spectroscopic analysis confirms the doubly deprotonated state of the ONO ligands for all [WO2(LR)(D)] complexes (Scheme 2). The bands found in the IR spectra of H2LR, characteristic for the N–H and C[double bond, length as m-dash]O groups (at ca. 3150 cm−1 and 1645 cm−1, respectively) are absent in the IR spectra of the tungsten complexes, suggesting tautomerism (C[double bond, length as m-dash]N–NH–(C[double bond, length as m-dash]O)– → = N–NC[double bond, length as m-dash](C–OH)–), deprotonation and coordination through the oxygen atom. This is additionally supported by the presence of a new band at ca. 1315 cm−1 assigned to the stretching vibrations of the C–O bond.22 Two stretching vibration bands belonging to the C[double bond, length as m-dash]Nimine and C–Ophenolic bonds appeared in the spectra of the complexes at ca. 1600 cm−1 and 1340 cm−1.22 They indicate that coordination of the ligand L2− to the cis-{WO2}2+ core occurs through the ONO donor atoms (Scheme 2). A band around 1050 cm−1, seen in the IR spectra of 1a–3a and 1b–3b, is assigned to the C–O stretching vibration of the coordinated alcohol molecule.23

The presence of the {WO2}2+ core in 1a–3a and 1b–3b was identified by the appearance of the strong absorption bands found in the range 959–949 cm−1 and 920–896 cm−1, characteristic for νasym(WO2) and for the antisymmetric combination of W[double bond, length as m-dash]O and W–OEtOH (or W–OMeOH) stretchings, respectively.24 The corresponding symmetric stretching bands νsym(WO2) having significantly lower intensities appeared in the same region but they either overlap with the asymmetric ones or appear as shoulders.

UV-Vis spectroscopy

The electronic absorption spectrum of the ligands exhibit several absorption bands with distinct maxima at 310 and 225 nm (H2L3OMe), at 338 and 219 nm (H2L4OMe), and at 294 and 325 nm (H2LH). These transitions are assigned to the intra-ligand charge transfer transitions.25,26 Electronic spectra of the dioxotungsten(VI) complexes 1a–3a and 1b–3b recorded in methanol exhibited N(pπ)–W(dπ) LMCT and O(pπ)–W(dπ) LMCT bands in the 410–404 nm and 337–309 nm regions.27 Each complex displays additional bands in the higher energy region characteristic for intraligand transitions.

NMR spectroscopy

The structures of all compounds in solution have been confirmed by one- and two-dimensional NMR techniques. The proton and carbon resonances were unambiguously assigned by using 1H, APT, COSY, HSQC and HMBC experiments and are displayed in the ESI (Tables S5 and S6, Scheme S1, see ESI). In the 1H NMR spectra of the ligand molecules the most deshielded protons were those belonging to the NH and OH groups resonating at around 12 and 11 ppm, respectively (Fig. S4–S6, see ESI). The signals were somewhat broadened indicating their involvement in hydrogen bonding interactions. In the complexes 1a, 2a and 3a those signals disappeared reflecting their deprotonation upon coordination of the ligands to tungsten.

Deprotonation was accompanied by shielding and deshielding of the neighbouring hydrogen and carbon atoms (see Tables S1 and S2), as a result of the electron redistribution upon complexation. Most of the neighbouring carbon atoms were deshielded, i.e. they underwent upfield shifts upon complexation up to 10.34 ppm as observed for carbon C-1 in 1a. The atoms C-4, C-12, C-14 and C-15 were also deshielded but to a smaller extent, while C-5 experienced a shielding effect up to 3.15 ppm as observed in 1a. These findings are in accordance with the crystal structures obtained for these compounds.

Catalytic studies

The catalytic reactions were performed under different experimental conditions in order to better understand the role of solvent that delivers TBHP, water (procedure A) and decane (procedure C) as well as the role of added organic solvent (herein MeCN) with aqueous TBHP (procedure B) (Scheme 3).
image file: c6ra05067k-s3.tif
Scheme 3 The different conditions used for the epoxidation.

As seen previously,6a TBHP in water without catalyst does react poorly with cyclooctene under organic solvent free conditions, as well as in the presence of organic solvent.28

Organic solvent-free epoxidation (catalytic procedure A)

All prepared mononuclear tungsten complexes [WO2(LR)(D)] were employed as (pre)catalysts using aqueous TBHP as oxidant, no additional organic solvent, and a 0.25 mol% catalyst charge. After the TBHP addition at 80 °C, the tungsten complexes remained partially undissolved at the beginning of the reaction, but dissolved completely as the reactions progressed. Under these conditions, it is known that TBHP is mostly transferred to the organic phase.6b The substrate conversion and the desired epoxide formation were monitored by analyzing only the organic layer. The analyses of the aqueous layer were not carried out because the desired epoxide is confined in the organic layer and only diol by-product may be partially lost in the aqueous layer. The obtained results are compiled in Table 2 and the kinetic profiles are presented in Fig. 4. The complexes with coordinated ethanol (1a–3a) are slightly more active than those with methanol (1b–3b). The profiles show similar evolution, with good activity in the first 20 minutes followed by a slowdown. The two (pre)catalysts 3a and 2a, however, seemed to regain activity after 150 min.
Table 2 Relevant catalytic results for the cyclooctene epoxidation by aqueous TBHPa
Cat. Conversionb/% Selectivityc/% TOF20 mind/h−1 TONe
a Reaction conditions: time, 5 h; temperature, 80 °C; [W]/cyclooctene/TBHP molar ratio: 0.25/100/200.b Calculated after 5 h.c Formed epoxide per converted olefin after 5 h.d n(cyclooctene transformed)/n(catalyst)/time at 20 minutes.e n(cyclooctene transformed)/n(catalyst at 5 h).
1a 20 3 73 63
1b 17 12 171 70
2a 28 33 81 117
2b 23 8 111 93
3a 31 11 204 129
3b 22 6 168 90



image file: c6ra05067k-f4.tif
Fig. 4 Kinetic monitoring of cis-cyclooctene epoxidation with TBHP–H2O in the presence of tungsten complexes: 1a – black circle, 1b – white circle, 2a – black triangle, 2b – white triangle, 3a – black square, 3b – white square. Experimental conditions are in Table 1.

The cyclooctene conversion for all tested compounds after 5 h is between 17 and 31%, following the order 3a > 2a > 2b3b > 1a > 1b. The cycloctene oxide selectivity is rather poor, varying from 8 to 33%, certainly implying that the organic solvent-free conditions and water from TBHP promote the cyclooctene epoxide ring opening into the corresponding cyclooctanediol.6 The highest selectivity was observed for 2a and the lowest one for 1a. The OMe substituent position has a marked influence on the reaction selectivity. Poor results in terms of selectivity and conversion could also be caused by the inhibition of the tungsten catalyst in water, as stated elsewhere with other tungsten-containing complexes.29–31 The highest turnover frequency (TOF) is observed for 3a.

We previously reported results of epoxidation catalysis with [WO2(LR*)(EtOH)] complexes (where LR* are the corresponding 4-hydroxy-benzhydrazide-related ligands).12 When comparing with the isoniazide performance, it can be concluded that the ligand hydrazone moiety does not influence the activity whereas the aldehyde moiety does, the 4OMe derivative always giving higher conversion than the corresponding 3OMe derivative (Tables S7 and Scheme S2, see ESI). This could be related to a steric impediment of the 3OMe substituent toward the approach and coordination of TBHP to the metal centre. As discussed earlier for the catalysts with molybdenum complexes having a similar coordination sphere, the catalytic cycle probably starts from the pentacoordinated mononuclear compound [WO2(LR)]. The sixth coordination site is available for the TBHP coordination and activation through the assistance of an O–H⋯O hydrogen bond.6b

Epoxidation of cyclooctene by aqueous TBHP in MeCN (catalytic procedure B)

Acetonitrile is a commonly used solvent for metal-catalyzed epoxidation.30 Addition of MeCN to the reaction mixture positively affected the 1–3 catalyst activity, see Table 3. In all cases, significant improvement of cyclooctene conversion was observed (35–63%) after only 20 min of the reaction progress. Like under organic solvent-free conditions, the ethanol-stabilized complexes were more active than those containing methanol. The higher conversion for the complexes containing (L4OMe)2− and (L3OMe)2− ligands than for those having the (LH)2− ligand is more pronounced in MeCN. After 20 min, the conversion reached again a plateau, implying possible catalyst deactivation. All catalysts were poorly selective towards the corresponding epoxide, which clearly indicates that the addition of organic solvent does not promote epoxide formation. The acetonitrile effect seems complex, since many parameters affect the catalyst action: polarity, reactant and product solubility, diffusion effects, solvent coordination to the metal center, side reactions, etc.32,33 It has been suggested that MeCN addition to alcohol-stabilized mononuclear tungsten complexes leads to oligomerization.12 Certain polynuclear complexes were found more active than the corresponding alcohol-stabilized mononuclear ones,5a,5c possibly explaining the better conversion in presence of MeCN with the tested catalysts.
Table 3 Comparison of the catalytic parameters (conversion and selectivity) for the cyclooctene epoxidation by TBHP (procedures A, B and C)a
a Reaction conditions: time, 5 h; temperature, 80 °C; [W]/cyclooctene/TBHP molar ratio: 0.25/100/200 for all compounds.
Procedure A B C
Solvent MeCN
TBHP in Water Water Decane

Catalyst Conv. (selec.) % Conv. (selec.) % Conv. (selec.) %
1a 20 (3) 66 (1) 20 (31)
1b 17 (12) 51 (2) 22 (12)
2a 28 (33) 78 (3) 20 (27)
2b 23 (8) 63 (2) 25 (19)
3a 31 (11) 47 (2) 29 (31)
3b 22 (6) 35 (1) 23 (13)


Epoxidation of cyclooctene by TBHP/decane under organic solvent-free conditions (catalytic procedure C)

The use of TBHP in decane did not result in any dramatic change of the conversion parameters compared to procedure A, yielding very similar results for all catalysts. This contrasts with the positive contribution of decane observed with other complexes.30 The solvent that delivers TBHP to the cyclooctene solution (water or decane) has no influence on the conversion rate. On the other hand, the epoxide selectivity was the same or better than using aqueous TBHP. This has certainly to be linked to the protective hydrophobic effect of decane towards the epoxide opening with water.34

One blank reaction was performed using TBHP in decane and no solvent. After 5 h cyclooctene conversion of 17% was observed with selectivity of 49% towards epoxide. The catalysts have a very low conversion when using 0.25 mol% catalyst charge. However, when adding the W complexes, the lower selectivity showed that another transformation occurred in addition to the epoxidation, but the nature of the other compounds was not determined.

Conclusions

Replacement of the acetylacetonate ligands in [WO2(acac)2] by the corresponding aroylhydrazone ligand H2LR in ethanol or methanol results in mononuclear complexes [WO2(LR)(EtOH)] (1a–3a) and [WO2(LR)(MeOH)] (1b–3b), respectively. Reactions in the presence of a small amount of DMF are effective in producing crystals suitable for single crystal X-ray diffraction experiment.

The association of the molecules by hydrogen bonds into centrosymmetric dimers is a common feature of 1a, 1b and 3a. Furthermore, the dimers are held together by π⋯π interactions involving the pyridyl rings.

All the dioxotungsten(VI) complexes led to greater cyclooctene conversion by oxidation with TBHP in an acetonitrile solution than in an organic solvent-free process but with very poor selectivity towards the desired epoxide. Under organic solvent-free conditions, use of TBHP in water or decane has no impact to the cyclooctene conversion. Nevertheless, the use of TBHP in decane has a positive impact on the selectivity in comparison to the use of aqueous TBHP.

In general, the catalytic behaviour of the complexes depends on the aroylhydrazone ligand, more precisely on the presence and position of the substituent present in the salicylaldehyde moiety. Compounds [WO2(L4OMe)(D)] achieved higher conversions than the corresponding [WO2(L3OMe)(D)], indicating the importance of the ring substituent position. The complexes coordinated with ethanol showed higher conversions than those with methanol.

Experimental section

Preparative part

Ligands H2L3OMe, H2L4OMe and H2LH were prepared by Schiff base condensation according to the procedure described in the literature.35 The tungsten complex WO2Cl2 was prepared according to literature procedure.36 Toluene and acetylacetone were dried by refluxing over P2O5 and then distilled. Diethyl-ether was dried over sodium wire and distilled. Methanol and ethanol were dried using magnesium turnings and iodine and then distilled. For the catalytic investigations cis-cyclooctene (98% Aldrich), cyclooctene oxide (Aldrich), cis-1,2-cyclooctanediol (99% Aldrich), TBHP (70% in water, Aldrich), H2O2 (35% in water, Aldrich) and organic solvents (acetonitrile, diethyl ether from ACROS) were commercially available and used as received.

Synthesis of [WO2(acac)2]·0.5C6H5Me

A suspension of WO2Cl2 (2 g, 7 mmol) and 20 mL of acetylacetone in 120 mL of toluene was refluxed under a dry argon atmosphere for 9 h and left overnight. The obtained mixture was refluxed additionally for 2 hours and filtered in a glove box while still hot to remove insoluble impurities. The solvent was evaporated by vacuum distillation to approximately 5–10 mL and the resulting solution was then kept overnight at 10 °C to give a white solid. The obtained precipitate was filtered and washed with dry diethyl-ether. Yield: 1.9 g; 59%. Anal. calcd for C13.5H18O6W (460.12): C, 35.2; H, 3.9. Found: C, 35.4; H, 3.8%. TG: calc. for C6H5CH3 10.01%, found 9.85% calc. for WO3 50.4%, found 50.35%. Selected IR data (cm−1): 1558, 1515, 954, 937, 909 cm−1.

Synthesis of dioxotungsten(VI) complexes with hydrazone ligands

A mixture of [WO2(acac)2]·0.5C6H5Me (0.22 mmol) and H2LR (0.22 mmol) and 20 μL DMF in dry methanol or ethanol (35 mL) was refluxed for five hours under a dry argon atmosphere. The obtained solution was cooled down to room temperature and then concentrated under vacuum to one third of its volume. The resulting solution was left at room temperature for a few days. The obtained yellow precipitate was filtered, rinsed with cold alcohol and dried in a desiccator up to the constant weight.
[WO2(L3OMe)(EtOH)] (1a). Yield: 0.05 g; 43%. Anal. calcd for C16H17N3O6W (531.17): C, 36.2; H, 3.2; N, 7.9. Found: C, 35.9; H, 3.2; N, 7.6%. TG: calc. for EtOH 8.7%, found 8.35% calc. for WO3 43.6%, found 43.9%. Selected IR data (cm−1): 1622 (C[double bond, length as m-dash]N)py, 1600 (C[double bond, length as m-dash]N), 1341 (C–Ophenolate), 1315 (C–O), 1194 (C–OEtOH), 959, 948 (WO2), 920, 910 (O[double bond, length as m-dash]W–OEtOH). UV-Vis (methanol): λ/nm (ε/dm3 mol−1 cm−1): 225 (42[thin space (1/6-em)]500), 309 (27[thin space (1/6-em)]350) and 410 (16[thin space (1/6-em)]314).
[WO2(L4OMe)(EtOH)] (2a). Yield: 0.05 g, 43%. Calc. for C16H17N3O6W (531.14): C, 36.2; H, 3.2; N, 7.9. Found: C, 36.0; H, 3.0; N, 7.6%. TG: calc. for EtOH 8.7%, found 8.4%, calc. for WO3 43.65%, found 43.1%. IR data (cm−1): 1616 (C[double bond, length as m-dash]N)py, 1600 (C[double bond, length as m-dash]N), 1334 (C–Ophenolate), 1315 (C–Oenolate), 1199 (C–OEtOH), 949, 943 (WO2), 896 (O[double bond, length as m-dash]W–OEtOH). UV-Vis (methanol): λ/nm (ε/dm3 mol−1 cm−1): 217 (38[thin space (1/6-em)]300), 337 (24[thin space (1/6-em)]800) and 407 (6414).
[WO2(LH)(EtOH)] (3a). Yield: 0.06 g, 55%. Calc. for C15H15N3O5W (501.14): C, 35.95; H, 3.0; N, 8.4. Found: C, 35.65; H, 2.8; N, 8.2%. TG: calc. for EtOH 9.2%, found 8.9%, calc. for WO3 46.3%, found 45.95%. Selected IR data (cm−1): 1621 (C[double bond, length as m-dash]N)py, 1600 (C[double bond, length as m-dash]N), 1339 (C–Ophenolate), 1317 (C–O), 954 (WO2), 913, 905 (O[double bond, length as m-dash]W–OEtOH). UV-Vis (methanol): λ/nm (ε/dm3 mol−1 cm−1): 216 (27[thin space (1/6-em)]500), 294 (15[thin space (1/6-em)]650), 333 (12[thin space (1/6-em)]950) and 404 (1507).
[WO2(L3OMe)(MeOH)] (1b). Yield: 0.07 g; 62%. Anal. calcd for C15H17N3O6W (517.142): C, 34.8; H, 2.9; N, 8.1. Found: C, 35.9; H, 3.2; N, 7.6%. TG: calc. for MeOH 6.2%, found 6.2%, calc. for WO3 44.8%, found 43.45%. Selected IR data (cm−1): 1621 (C[double bond, length as m-dash]N)py, 1600 (C[double bond, length as m-dash]N), 1345 (C–Ophenolate), 1315 (C–O), 1194 (C–OEtOH), 949 (WO2), 920, 910 (O[double bond, length as m-dash]W–OEtOH). UV-Vis (methanol): λ/nm (ε/dm3 mol−1 cm−1): 225 (42[thin space (1/6-em)]500), 309 (27[thin space (1/6-em)]350) and 410 (16[thin space (1/6-em)]314).
[WO2(L4OMe)(MeOH)] (2b). Yield: 0.06 g, 53%. Calc. for C15H15N3O6W (517.14): C, 34.8; H, 2.9; N, 8.1. Found: C, 38.5; H, 3.1; N, 7.9%. TG: calc. for MeOH 6.2%, found 6.4%, calc. for WO3 44.8%, found 44.4%. IR data (cm−1): 1615 (C[double bond, length as m-dash]N)py, 1600 (C[double bond, length as m-dash]N), 1334 (C–Ophenolate), 1314 (C–O), 1199 (C–OEtOH), 949, 943 (WO2), 896 (O[double bond, length as m-dash]W–OEtOH). UV-Vis (methanol): λ/nm (ε/dm3 mol−1 cm−1): 217 (38[thin space (1/6-em)]300), 337 (24[thin space (1/6-em)]800) and 407 (6414).
[WO2(LH)(MeOH)] (3b). Yield: 0.08 g, 75%. Calc. for C14H13N3O5W (487.1): C, 34.5; H, 2.7; N, 8.6. Found: C, 34.7; H, 2.8; N, 8.7%. TG: calc. for MeOH 6.6%, found 6.3%, calc. for WO3 47.6%, found 47.4%. Selected IR data (cm−1): 1622 (C[double bond, length as m-dash]N)py, 1602 (C[double bond, length as m-dash]N), 1340 (C–Ophenolate), 1317 (C–O), 955 (WO2_asym), 914, 906 (WO2_sym). UV-Vis (methanol): λ/nm (ε/dm3 mol−1 cm−1): 216 (27[thin space (1/6-em)]500), 294 (15[thin space (1/6-em)]650), 333 (12[thin space (1/6-em)]950) and 404 (1507).

General procedure for the epoxidation of cyclooctene by aqueous TBHP

Procedure A. A mixture of cyclooctene (2.76 mL, 20 mmol), acetophenone (as an internal reference) and W (pre)catalyst 1a–3a and 1b–3b (0.05 mmol) was stirred and heated up to 80 °C before addition of aqueous TBHP (70% w/w, 5.48 mL, 40 mmol). The mixture is initially an emulsion, but two phases become clearly visible as the reaction progresses, a colourless aqueous one and a yellowish organic one. The reaction was monitored for 5 h with withdrawal and analysis of organic phase aliquots (0.1 mL) at required times. Each withdrawn sample was mixed with 2 mL of Et2O, treated with a small quantity of MnO2 and then filtered through silica and analyzed by GC.
Procedure B. The procedure was performed in the same manner as procedure A with addition of 2 mL of MeCN before the addition of aqueous TBHP.
Procedure C. The same as procedure A with using TBHP in decane (40 mmol, 6.5 mL) as oxidizing agent.
Physical methods. Elemental analyses were provided by the Analytical Services Laboratory of the Ruđer Bošković Institute, Zagreb. Thermogravimetric (TG) analysis was carried out with a Mettler-Toledo TGA/SDTA851e thermobalance using aluminum crucibles. All experiments were recorded in a dynamic atmosphere with a flow rate of 200 cm3 min−1. Heating rates of 5 K min−1 were used for all investigations. Fourier Transform Infrared spectra (FT-IR) were recorded in KBr pellets with a Perkin-Elmer 502 spectrophotometer. Spectra were recorded in the spectral range between 4500 and 450 cm−1. Electronic absorption spectra were recorded at 25 °C on a Cary 100 UV-Vis Spectrophotometer.

All NMR spectra were recorded on Bruker Avance III HD 400 spectrometer operating at 400 MHz equipped with a broadband observed (BBO) Prodigy cryoprobe and z-gradient accessory. Compounds were dissolved in DMSO-d6 and measured in 5 mm NMR tubes at 298 K with TMS as an internal standard. The sample concentration was 10 mg mL−1.

Proton spectra with spectral width of 6002 Hz and a digital resolution of 0.36 Hz per point were measured with 64 and 128 scans. In the gCOSY experiment, 2048 points in the f2 dimension and 128 increments in the f1 dimension were used. For each increment, 16 scans and the spectral width of 6010 Hz were applied. Digital resolution was 5.87 and 93.78 Hz per point in f2 and f1 dimensions, respectively. The gHSQC and gHMBC spectra were acquired with 16 and 32 scans, respectively. Spectral width was 6009 Hz in f2 and 20[thin space (1/6-em)]124 Hz in f1 dimension for both experiments. 1k data points were applied in the time domains and for each data set 256 increments were collected. The resulting digital resolution was 5.86 Hz per point in f2 dimension and 314.4 Hz per point in f1 dimension.

Catalytic reactions were followed by gas chromatography on an Agilent 6890A chromatograph equipped with FID detector, a HP5-MS capillary column (30 m × 0.25 mm × 0.25 μ) and automatic sampling, or on a Fisons GC 8000 chromatograph equipped with FID detector and with a SPB-5 capillary column (30 m × 0.32 mm × 0.25 μ). The GC parameters were quantified with authentic samples of the reactants and products. The conversion of cis-cyclooctene and the formation of cyclooctene oxide were calculated from calibration curves (r2 = 0.999) relatively to an internal standard.

X-ray crystallography. Powder diffraction. The powder X-ray diffraction data were collected by the Panalytical X'Change powder diffractometer in the Bragg–Brentano geometry using CuKα radiation. The sample was contained on a Si sample holder. Patterns were collected in the range of 2θ = 5–40° with the step size of 0.03° and at 1.5 s per step. The data were collected and visualized using the X'Pert programs Suite.37
X-ray crystallography. Single crystal diffraction. The X-ray single-crystal structure data obtained for [WO2(acac)2]·0.5C6H5Me, 1a, 1b and 3a are tabulated in Table 4. The single crystal X-ray diffraction data were collected by ω-scans on an Oxford Diffraction Xcalibur 3 CCD diffractometer with graphite-monochromated Mo Kα radiation (λ = 0.71073 Å). Each single crystal was glued onto a glass fibber and the data were collected at room temperature. The data reduction was performed using the CrysAlis software package.38 Solution, refinement and analysis of the structures were done using the programs integrated in the WinGX system.39
Table 4 Crystallographic data for compounds [WO2(C5H7O2)2]·0.5C6H5Me, 1a, 1b and 3a
  [WO2(acac)2]·0.5C6H5Me 1a 1b 3a
a R = Σ∣∣Fo∣ − ∣Fc∣∣/Σ∣Fo∣.b wR = [Σ(Fo2Fc2)2/Σw(Fo2)2]1/2.c S = Σ[w(Fo2Fc2)2/(NobsNparam)]1/2
Empirical formula C13.5H18O6W C16H17N3O6W C15H15N3O6W C15H15N3O5W
Mr 460.12 531.17 517.14 501.14
Crystal colour, habit Colourless, plate Yellow, plate Yellow, plate Yellow, plate
Crystal size (mm3) 0.56 × 0.36 × 0.18 0.56 × 0.38 × 0.18 0.48 × 0.34 × 0.18 0.56 × 0.20 × 0.12
Crystal system Monoclinic Triclinic Monoclinic Monoclinic
Space group P21/n P[1 with combining macron] P21/n P21/n
[thin space (1/6-em)]
unit cell parameters
a (Å) 7.3569(4) 7.2563(4) 6.7820(3) 8.1744(2)
b (Å) 16.2269(7) 10.4696(5) 30.3344(16) 12.5034(2)
c (Å) 13.5020(11) 12.6154(5) 7.9111(5) 15.3736(3)
α (°) 90 113.702(4) 90 90
β (°) 95.619(6) 96.023(4) 93.335(5) 97.032(2)
γ (°) 90 97.909(4) 90 90
V3) 1604.12(17) 855.54(8) 1624.78(15) 1559.48(6)
Z 4 2 4 4
Dcalc (g cm−3) 1.905 2.062 2.114 2.135
Temperature (K) 295 295 295 295
μ (mm−1) 7.220 6.791 7.148 7.439
F(000) 884 512 992 960
Number of unique data 3639 5682 3835 3735
Number of data [Fo ≥ 4σ(Fo)] 2646 5228 3053 2805
Number of parameters 172 236 227 221
R1a, [Fo ≥ 4σ(Fo)] 0.0327 0.0269 0.0251 0.0235
wR2b 0.0841 0.0585 0.0579 0.0438
Goodness of fit on F2, Sc 1.00 1.00 1.10 0.94
Min. and max. electron density (e Å−3) −2.47, 0.76 −1.35, 1.67 −1.80, 1.33 −0.61, 0.86


The structures were solved by using the SHELXS program and the Patterson method. The refinement procedure was performed by the full-matrix least-squares method based on F2 against all reflections using SHELXL.40

In [WO2(acac)2]·0.5C6H5Me the non-hydrogen atoms were refined anisotropically with exception of the toluene solvent molecule atoms which were refined isotropically. The toluene solvent molecule is situated near the centre of symmetry and is disordered over the two centrosymmetrically related positions. Restraints on the C–C distances and angles, and planarity (FLAT) were used for the toluene molecule. The centre of symmetry is within the toluene molecule and is located between atoms C12 and its centrosymmetrically related pair C12[−x, −y, −z] so placement of H-atoms at C12[−x, −y, −z] and the neighbouring C15 was difficult and their positions had to be fixed (Fig. S1). All other hydrogen atoms were placed in calculated positions and refined using the riding model.

The non-hydrogen atoms in 1a, 1b and 3a were refined anisotropically. The hydrogen atoms in 1a, 1b and 3a were placed in calculated positions and refined using the riding model, exceptions were hydrogen atoms of the coordinated solvent molecule hydroxyl group.

Acknowledgements

Financial support for this research was provided by Ministry of Science and Technology of the Republic of Croatia. We acknowledge the Centre National de la Recherche Scientifique (CNRS) for financial support, the Université Paul Sabatier and its Institut Universitaire Technologique for the facilities. All authors want to thank Prof. R. Poli for suggestions and advices.

Notes and references

  1. (a) M. C. Rodríguez-Argüelles, S. Mosquera-Vázquez, P. Tourón-Touceda, J. Sanmartín-Matalobos, A. M. García-Deibe, M. Belicchi Ferrari, G. Pelosi, C. Pelizzi and F. Zani, J. Inorg. Biochem., 2007, 101, 138 CrossRef PubMed; (b) D. S. Kalinowski, P. C. Sharpe, P. V. Bernhardt and D. R. Richardson, J. Med. Chem., 2008, 51, 331 CrossRef CAS PubMed.
  2. (a) V. Vrdoljak, B. Prugovečki, D. Matković-Čalogović, R. Dreos, P. Siega and C. Tavagnacco, Cryst. Growth Des., 2010, 10, 1373 CrossRef CAS; (b) V. Vrdoljak, B. Prugovečki, D. Matković-Čalogović, J. Pisk, R. Dreos and P. Siega, Cryst. Growth Des., 2011, 11, 1244 CrossRef CAS; (c) V. Vrdoljak, B. Prugovečki, D. Matković-Čalogović, T. Hrenar, R. Dreos and P. Siega, Cryst. Growth Des., 2013, 13, 3773 CrossRef CAS; (d) W.-X. Xu, Y.-M. Yuan and W. H. Li, J. Coord. Chem., 2013, 66, 2726 CrossRef CAS; (e) M. R. Maurya, L. Rana and F. Avecilla, Inorg. Chim. Acta, 2015, 429, 138 CrossRef CAS.
  3. (a) V. Vrdoljak, B. Prugovečki, I. Pulić, M. Cigler, D. Sviben, J. Parlov Vuković, P. Novak, D. Matković-Čalogović and M. Cindrić, New J. Chem., 2015, 39, 7322 RSC; (b) J. Pisk, B. Prugovečki, D. Matković-Čalogović, T. Jednačak, P. Novak, D. Agustin and V. Vrdoljak, RSC Adv., 2014, 4, 39000 RSC; (c) S.-B. Ding and W.-H. Li, J. Coord. Chem., 2013, 66, 2023 CrossRef CAS; (d) M. Nandy, S. Shit, C. Razzoli, G. Pilet and S. Mitra, Polyhedron, 2015, 88, 63 CrossRef CAS.
  4. A. Dolbecq, P. Mialane, L. Lisnard, J. Marrot and F. Sécheresse, Chem.–Eur. J., 2003, 9, 2914 CrossRef CAS PubMed.
  5. (a) J. Pisk, D. Agustin, V. Vrdoljak and R. Poli, Adv. Synth. Catal., 2011, 353, 2910 CrossRef CAS; (b) M. E. Judmaier, C. Holzer, M. Volpe and N. C. Mösch-Zanetti, Inorg. Chem., 2012, 51, 9956 CrossRef CAS PubMed; (c) J. Pisk, B. Prugovečki, D. Matković-Čalogović, R. Poli, D. Agustin and V. Vrdoljak, Polyhedron, 2012, 33, 441 CrossRef CAS; (d) M. Bagherzadeh, M. Amini, H. Parastar, M. Jalali-Heravi, A. Ellern and L. K. Woo, Inorg. Chem. Commun., 2012, 20, 86 CrossRef CAS; (e) M. Bagherzadeh, M. Zare, V. Amani, A. Ellern and L. K. Woo, Polyhedron, 2013, 55, 223 CrossRef.
  6. (a) B. Guérin, D. Mesquita-Fernandes, J.-C. Daran, D. Agustin and R. Poli, New J. Chem., 2013, 37, 3466 RSC; (b) J. Morlot, N. Uyttebroeck, D. Agustin and R. Poli, ChemCatChem, 2013, 5, 601 CrossRef CAS.
  7. (a) A. Rezaeifard, I. Sheikhshoaie, N. Monadi and M. Alipour, Polyhedron, 2010, 29, 2703 CrossRef CAS; (b) M. Loubidi and D. Agustin, C. R. Chim., 2014, 17, 549 CrossRef CAS; (c) W. Wang, T. Vanderbeeken, D. Agustin and R. Poli, Catal. Commun., 2015, 63, 26 CrossRef CAS; (d) W. Wang, T. Guerrero, S. R. Merecias, H. García-Ortega, R. Santillan, J.-C. Daran, N. Farfán, D. Agustin and R. Poli, Inorg. Chim. Acta, 2015, 431, 176 CrossRef CAS.
  8. M. A. Katkar, S. N. Rao and H. D. Juneja, RSC Adv., 2012, 2, 8071 RSC.
  9. X. Zhou, J. Zhao, A. M. Santos and F. E. Kühn, Z. Naturforsch., B: J. Chem. Sci., 2004, 59, 1223 CAS.
  10. A. Dupé, M. K. Hossain, J. A. Schachner, F. Belaj, A. Lehtonen, E. Nordlander and N. C. Mösch-Zanetti, Eur. J. Inorg. Chem., 2015, 3572 CrossRef.
  11. D. D. Agarwal, R. Rastogi, V. K. Gupta and S. Singh, Indian J. Chem., Sect. A: Inorg., Bio-inorg., Phys., Theor. Anal. Chem., 1999, 38, 369 Search PubMed.
  12. V. Vrdoljak, J. Pisk, D. Agustin, P. Novak, J. Parlov Vuković and D. Matković-Čalogović, New J. Chem., 2014, 38, 6176 RSC.
  13. (a) S.-B. Yu and R. H. Holm, Inorg. Chem., 1989, 28, 4385 CrossRef CAS; (b) W. A. Herrmann, J. Fridgen, G. M. Lobmaier and M. Spiegler, New J. Chem., 1999, 23, 5 RSC.
  14. D. Agustin, C. Bibal, B. Neveux, J.-C. Daran and R. Poli, Z. Anorg. Allg. Chem., 2009, 635, 2120 CrossRef CAS.
  15. Cambridge Structural Database V5.37, Cambridge Crystallographic Data Centre, 12 Union Road, Cambridge, England, 2015 Search PubMed.
  16. (a) F. Zhi and R. Wang, Acta Crystallogr., Sect. E: Struct. Rep. Online, 2010, 66, o892 CAS; (b) S.-J. Peng and H.-Y. Hou, Acta Crystallogr., Sect. E: Struct. Rep. Online, 2008, 64, o1996 CAS.
  17. Y.-J. Xu, S. Zhao and S. Bi, Acta Crystallogr., Sect. E: Struct. Rep. Online, 2007, 63, o4633 CAS.
  18. Y. Lei and C. Fu, Synth. React. Inorg., Met.-Org., Nano-Met. Chem., 2011, 41, 704 CrossRef CAS.
  19. S. Alghool and C. Slebodnick, Polyhedron, 2014, 67, 11 CrossRef CAS.
  20. Y.-L. Zhai, X.-X. Xu and Y. Wang, Polyhedron, 1992, 11, 415 CrossRef CAS.
  21. V. S. Sergienko, V. L. Abramenko, L. K. Minacheva, M. A. Porai-Koshits and V. G. Sakharova, Koord. Khim., 1993, 19, 28 CAS.
  22. M. R. Maurya, S. Agarwal, C. Bader and D. Rehder, Eur. J. Inorg. Chem., 2005, 147 CrossRef CAS.
  23. M. R. Maurya, S. Gopinathan, C. Gopinathan and R. C. Maurya, Polyhedron, 1993, 12, 159 CrossRef CAS.
  24. (a) Y.-L. Wong, J.-F. Ma, W.-F. Law, Y. Yan, W.-T. Wong, Z.-Y. Zhang, T. C. W. Mak and D. K. P. Ng, Eur. J. Inorg. Chem., 1999, 313 CrossRef CAS; (b) Y.-L. Wong, Y. Yan, E. S. H. Chan, Q. Yang, T. C. W. Mak and D. K. P. Ng, J. Chem. Soc., Dalton Trans., 1998, 3057 RSC.
  25. K. Chichak, U. Jacquenard and N. R. Branda, Eur. J. Inorg. Chem., 2002, 357 CrossRef CAS.
  26. P. Krishnamoorthy, P. Satyadevi, P. T. Mutiah and N. Dharmaraj, RSC Adv., 2012, 2, 12190 RSC.
  27. S. Gupta, A. Kumar Barik, S. Pal, A. Hazra, S. Roy, R. J. Butcher and S. Kumar Kar, Polyhedron, 2007, 26, 133 CrossRef CAS.
  28. (a) M. Lashanizadegan, S. Shayegan and M. Sarkheil, J. Serb. Chem. Soc., 2016, 81, 153 CrossRef; (b) M. Ghorbanloo, R. Bikas and G. Małecki, Inorg. Chim. Acta, 2016, 445, 8 CrossRef CAS; (c) C. I. Fernandes, M. D. Carvalho, L. P. Ferreira, C. D. Nunes and P. D. Vaz, J. Organomet. Chem., 2014, 760, 2 CrossRef CAS.
  29. M. Ciclosi, C. Dinoi, L. Gonsalvi, M. Peruzzini, E. Manoury and R. Poli, Organometallics, 2008, 27, 2281 CrossRef CAS.
  30. C. Dinoi, M. Ciclosi, E. Manoury, L. Maron, L. Perrin and R. Poli, Chem.–Eur. J., 2010, 16, 9572 CrossRef CAS PubMed.
  31. P. Sőzen-Aktas, E. Manoury, F. Demirhan and R. Poli, Eur. J. Inorg. Chem., 2013, 2728 CrossRef.
  32. A. Corma, P. Esteve and A. Martínez, J. Catal., 1996, 161, 11 CrossRef CAS.
  33. Z. Opre, T. Mallat and A. Baiker, J. Catal., 2007, 245, 482 CrossRef CAS.
  34. S. Bonollo, D. Lanari and L. Vaccaro, Eur. J. Org. Chem., 2011, 2587 CrossRef CAS.
  35. (a) J.-F. Lu, Acta Crystallogr., Sect. E: Struct. Rep. Online, 2008, 64, o2032 CAS; (b) H. H. Rassem, A. Salhin, B. Bin Salleh, M. M. Rosli and H.-K. Fun, Acta Crystallogr., Sect. E: Struct. Rep. Online, 2012, 68, o2279 CAS; (c) A. Salhin, A. A. Tameem, B. Saad, S.-L. Ng and H.-K. Fun, Acta Crystallogr., Sect. E: Struct. Rep. Online, 2007, 63, o2880 CAS.
  36. W. C. Gibson, T. P. Kee and A. Shaw, Polyhedron, 1988, 7, 579 CrossRef.
  37. X'Pert Software Suite, Version 1.3e, Panalytical B. V., Almelo, The Netherlands, 2001 Search PubMed.
  38. CrysAlisPro, Version 1.171.37.33, Agilent Technologies, release 27-03-2014 CrysAlis171.NET Search PubMed.
  39. L. J. Farrugia, J. Appl. Crystallogr., 1999, 32, 837 CrossRef CAS.
  40. G. M. Sheldrick, Acta Crystallogr., Sect. A: Found. Crystallogr., 2008, 64, 112 CrossRef CAS PubMed.

Footnote

Electronic supplementary information (ESI) available: (1) Spectral data, (2) XRPD patterns, (3) tables and (4) figures for compounds. Crystallographic data sets for the structures 1a, 1b, 3a, and [WO2(acac)2]·0.5C6H5Me. CCDC 1454679–1454682. For ESI and crystallographic data in CIF or other electronic format see DOI: 10.1039/c6ra05067k

This journal is © The Royal Society of Chemistry 2016
Click here to see how this site uses Cookies. View our privacy policy here.