1 s2.0 S0020169305004287 Main PDF

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 9

Inorganica Chimica Acta 359 (2006) 81–89

www.elsevier.com/locate/ica

Synthesis and characterization of thiocyanato and


chlorodioxo tungsten(VI) compounds: Comparative oxygen
atom transfer capability with molybdenum analogs
Henri Arzoumanian a,*, Giuseppe Agrifoglio b,*, Mario V. Capparelli c,
Reinaldo Atencio b, Alexander Briceño b, Angel Alvarez-Larena d
a
UMR 6180 CNRS, Université Paul Cézanne, Chirotechnologies: Catalyse et biocatalyse, Faculte des Sciences St. Jerome, Aix Marseille III, Case A62
Ave Escadrille Normandie Niemen, 13397 Marseille, France
b
Centro de Quı́mica, Instituto Venezolano de Investigaciones Cientı́ficas (IVIC) Apartado 21827, Caracas 1020-A, Venezuela
c
Escuela de Quı́mica, Facultad de Ciencias, Universidad Central de Venezuela, Apartado 47074, Caracas 1041-A, Venezuela
d
Servicio de Difracción de Rayos X, Universidad Autónoma de Barcelona, 08193 Bellaterra, Spain

Received 20 April 2005; received in revised form 26 July 2005; accepted 3 August 2005
Available online 30 September 2005

Abstract

Four new tungsten-oxo(VI) complexes have been synthesized, characterized spectroscopically and their molecular structure estab-
lished by X-ray diffraction analysis. These bear identical environment as previously reported molybdenum-oxo(VI) complexes, which
allowed direct comparison of their spectroscopic properties. Their capability as oxygen atom transfer agents was found to be significantly
lower than their molybdenum analogs.
 2005 Elsevier B.V. All rights reserved.

Keywords: Tungsten oxo complexes; X-ray structures; Oxygen atom transfer

1. Introduction [11–13] and many models have been synthesized and stud-
ied [14,15]; furthermore, tungsten-oxo or peroxo complexes
The chemistry of transition metal-oxo compounds is an are also known to be active in oxidation reactions specially
area of particular interest with potential relevance to oxy- in aqueous medium [16]. We, thus, considered worthy of
gen atom transfer processes in a wide range of catalytic interest the synthesis of the tungsten analogs of the molyb-
reactions. Molybdenum-oxo compounds hold an impor- denum complexes found to be performant oxygen atom
tant place in this field in systems going from metalloen- transfer agents in order to compare directly their proper-
zymes to solid state metal oxide surfaces [1–6]. In this ties. We chose four oxo compounds of different nature,
context, we have reported several molybdenum oxo, dioxo one anionic dioxo, two neutral monomeric dioxo and one
or l-oxo complexes bearing various other ligands among dioxo-l-oxo dimer. This report concerns the synthesis
which some exhibited exceptional oxygen atom transfer and characterization of these four new tungsten com-
capability either under stoichiometric or catalytic condi- pounds and their direct comparison with the corresponding
tions [7–10]. Another important transition metal-oxo of molybdenum complex.1
group VI is tungsten. Several tungsten enzymes are known

*
Corresponding authors. Tel.: +33 491288256; fax: +33 491027776.
1
E-mail addresses: henri.arzoumanian@univ.u-3mrs.fr (H. Arzouma- Similar observations have been reported on other molybdenum and
nian), gagrifog@cantv.net (G. Agrifoglio). tungsten analog complexes [17,18].

0020-1693/$ - see front matter  2005 Elsevier B.V. All rights reserved.
doi:10.1016/j.ica.2005.08.003
82 H. Arzoumanian et al. / Inorganica Chimica Acta 359 (2006) 81–89

2. Experimental 2.2.3. Dichloro(dioxo)-4,4 0 -tert-butyl-2,2 0 -


bipyridyltungsten(VI) (3)
2.1. General material and procedure A 50 mL CH3CN suspension containing 1.37 g (5 mmol)
of 4,4 0 -tert-butyl-2,2 0 -bipyridine was added 1.37 g
All materials were commercial products and used with- (5.1 mmol) of anhydrous Na2WO4. The suspension was
out further purification unless otherwise noted. All sol- stirred 30 min at room temperature and to it was added
vents, stored over 4 Å molecular sieves, were thoroughly 1.15 g (10.5 mmol) of trimethylchlorosilane. The mixture
degassed prior to use by repeated evacuation followed by was refluxed for 4 h and reduced to dryness under vacuum
admission of dry nitrogen or argon. 4,4 0 -tert-butyl-2,2 0 - to give a colorless solid residue. Recrystallization from
bipyridine was prepared by the method described by Ben CH2Cl2/n-hexane (2 /1) yielded a colorless crystalline solid
Hadda and Le Bozec [19]. NMR spectra were recorded (85%). IR (KBr); m(W@O), 944(s), 911(s) cm1. 1H NMR
on a Bruker AMX 400 spectrometer. IR spectra were re- (CDCl3, TMS) d 7.72 (dd, J = 6 Hz, 1.6 Hz, H5), 8.16 (d,
corded on a Perkin–Elmer 1720 XFT spectrometer. UV 1.6 Hz, H3), 9.10 (d, J = 6 Hz, H6). Anal. Calc. for
spectra were recorded on an HP 8452A diode array spec- C18H24N2O2Cl2W: C, 38.94; H, 4.35; N, 5.04. Found: C,
trometer with 89531 A operating software. 39.18; H, 4.42; N, 5.19%.

2.2.4. l-oxobis[cis-dioxo(thiocyanato-N)-4,4 0 -di-tert-butyl-


2.2. Synthesis of the complexes
2,2 0 -bipyridyltungsten(VI)] (4)
To an aqueous solution (30 mL) containing Na2WO4 Æ 2-
2.2.1. Bis(tetraphenylphosphonium)dioxo(tetrathiocyanato-
H2O (1.2 g, 3.6 mmol) and NaSCN (1.2 g, 15 mmol) was
N) tungstate(VI) (1)
slowly added 15 mL of 1 N HCl. After stirring at 25 C for
To a 30 mL aqueous solution of Na2WO4 Æ 2H2O (1.2 g,
15 min, a CH2Cl2 solution (60 mL) containing 0.96 g
3.6 mmol) and NaSCN (1.2 g, 15 mmol) was slowly added
(3.6 mmol) of 4,4 0 -tert-butyl-2,2 0 -bipyridine was added and
15 mL of 1 N HCl. To the resulting yellow solution was
vigorous stirring continued for 30 min. The organic layer
quickly added a 50 mL CH2Cl2 solution containing tetra-
was separated, dried over MgSO4 and evaporated to give a
phenylphosphonium chloride (2.7 g, 3.6 mmol). After
colorless crystalline solid. It was recrystallized from
5 min of vigorous stirring the organic layer was separated
CH2Cl2/n-hexane (2/1) at 15 C. IR(KBr): m(SCN) =
and dried over anhydrous MgSO4. Evaporation and recrys-
2055(s) cm1; m(W@O) = 959(s), 909(s) cm1; m(W–O–
tallization of the resulting solid residue (CH2Cl2/n-hexane,
W) = 812(s) cm1. 1H NMR (400 MHz, CDCl3, TMS) d
2/1)gave the desired product in 47% yield (m.p. > 200 C).
1.45(s), 1.47(s), 1.48(s), 1.49(s), 1.50(s), 6.49 (dd, J = 6 Hz,
IR (KBr): m(NCS) = 2104(s), 2057(s) cm1; m(W@O) =
1.6 Hz, H5), 7.51 (dd, J = 6 Hz, 1.6 Hz, H5),7.63 (dd,
941(s), 895(s) cm1. Anal. Calc. for C52H40N4P2S4O2W:
J = 6 Hz, 1.6 Hz, H5), 7.77 (m, H3), 7.80 (dd, J = 6 Hz,
C, 55.41; H, 3.57; N, 4.97. Found: C, 52.94; H, 3.57; N,
1.6 Hz, H5), 8.10 (d, 1.6 Hz, H3), 8.15 (d, 1.6 Hz, H3), 8.19
4.67%.
(d, 1.6 Hz, H3), 8.23 (d, 1.6 Hz, H3), 8.25 (d, 1.6 Hz, H3),
8.75 (d, J = 6 Hz, H6), 8.97 (d, J = 6 Hz, H6), 9.46 (d,
2.2.2. Dioxo(dithiocyanato-N)-4,4 0 -tert-butyl-2,2 0 - J = 6 Hz, H6).
bipyridyltungsten(VI) (2)
A CH2Cl2 solution (100 mL) containing 2.7 g 2.3. Crystal structure determination
(2.4 mmol) of 1 and 0.64 g (2.4 mmol) of freshly sub-
limed 4,4 0 -tert-butyl-2,2 0 -bipyridine was stirred at room Experimental details on unit cell and intensity measure-
temperature for 24 h with no apparent color change. ments can be found in the CIF files deposited with the
Diethyl ether (100 mL) was added to the mixture and Cambridge Crystallographic Data Centre. Crystal data,
the resulting solution kept at 5 C for 8 h. This resulted intensity data collection parameters and final refinement
in the precipitation of PPh4SCN. The solution obtained results are summarized in Table 1.
upon filtration was evaporated and the solid residue The systematic absences for 2 (h k 0: h = 2n + 1, 0 k l:
was checked by IR for the presence of residual phospho- k + l = 2n + 1), were compatible with space groups Pn21a
nium salt (if positive – three characteristic bands at 751, (No. 33) and Pnma (No. 62); the former was discarded
723 and 689 cm1 – the precipitation procedure was re- on the basis of refinement (non-positive definite Us, large
peated). The solid product was washed with diethyl correlation between pseudo symmetry-related atoms). The
ether, dried under vacuum and recrystallized (CH2Cl2/ systematic absences for 1 (h 0 0: h = 2n + 1; 0 k 0:
n-hexane, 2/1) to yield a very slightly pink crystalline so- k = 2n + 1; 0 l 0: l = 4n + 1), were compatible with either
lid. IR (KBr): m(SCN) = 2016(s) cm1; m(W@O) = 952(s), of the enantiomorphous space groups P41212 (No. 92) or
912(s) cm1. 1H NMR (CDCl3, TMS) d 1.51 (s, CH3), P43212 (No. 96). Refinement of the Flack parameter [20]
7.80 (dd, J = 6 Hz, 1.6 Hz, H5), 8.21 (d, J = 1.6 Hz, indicated that the latter was the correct one. For 4 the
H3), 9.13 (d, J = 6 Hz, H6). Anal. Calc. for space group P 1 ðNo. 2Þ was chosen on the basis of the Z
C20H24N4O2S2W: C, 40.01; H, 4.03; N, 9.32; S, 10.68. value. The space group of 3, P212121, was uniquely deter-
Found: C, 40.94; H, 4.33; N, 8.80; S, 10.70%. mined by the systematic absences.
H. Arzoumanian et al. / Inorganica Chimica Acta 359 (2006) 81–89 83

Table 1
Crystal data, intensity data collection parameters and final refinement results
Compound 2 1 4 3
CCDC deposit No. CCDC 268191 CCDC 268192 CCDC 268193 CCDC 268194
Crystal data
Formula C20H24N4O2S2W C52.64H41.28N4O2P2S4Cl1.28W C40H51N6O5S2W2 C18H24Cl2N2O2W
MW 600.40 1181.31 1141.70 555.14
Color colorless yellow colorless pale pink
Morphology irregular prism prism prism
Specimen size (mm) 0.44 · 0.36 · 0.34 0.42 · 0.36 · 0.34 0.20 · 0.11 · 0.07 0.22 · 0.20 · 0.18
T (K) 298(2) 298(2) 298(2) 298(2)
a (Å) 15.747(3) 12.838(6) 13.524(3) 8.3834(17)
b (Å) 13.947(2) 12.838(6) 13.642(3) 13.484(3)
c (Å) 10.674(2) 31.947(9) 15.488(6) 18.195(4)
a () 92.84(3)
b () 110.23(2)
c () 118.53(2)
V (Å3) 2344.3(7) 5265(2) 2274.3(13) 2056.8(7)
Crystal system orthorhombic tetragonal triclinic orthorhombic
Space group (No.) Pnma (62) P43212 (96) P 1 ð2Þ P212121 (19)
Z 4 8 2 4
Dc (g cm3) 1.701 1.490 1.667 1.793
F(0 0 0) 1176 2363.6 1120 1080
l (Mo Ka) (mm1) 5.128 2.521 5.193 5.889
h Range () for cell 10.4–15.0 11.3–14.1 8.5–15.4 1.9–28.9
Number of reflections for cell 23 18 25 12 376
Data collection
h Range () 2.3–25.0 1.7–24.9 1.5–25.0 1.9–28.1
h Range 0, 18 0, 15 16, 15 7, 9
k Range 0, 16 0, 10 16, 16 15, 15
l Range 12, 0 0, 38 0, 18 21, 20
Mean DI for checks (%) 3.1 0.3 <1.0 <1.0
Number of reflections measured 3718 2719 8301 19 047
Number of reflections unique 2168 2719 7974 3842
Number of reflections I > 2r(I) 1372 2098 5242 3675
Rint 0.1594 0.0325 0.0352
Refinement (last cycle)
Weighting scheme (a,b) 0.0352, 0.0 0.0828, 0.0 0.0469, 0.0 0.0438, 2.94
Transfer coefficient (Tmin, Tmax) 0.126, 0.175 0.370, 0.420 0.503, 0.694
Number of parameters refined 145 305 501 226
Number of restraints 0 0 36 0
Flack parameters (x) 0.06(2) 0.010(15)
R1 [I > 2r(I)] 0.0611 0.0486 0.0390 0.0388
R1 (all data) 0.1070 0.0640 0.0981 0.0415
wR2 [I > 2r(I)] 0.1005 0.1086 0.0817 0.0899
wR2 (all data) 0.1170 0.1155 0.0949 0.0917
S (Goodness-of-fit) (all data) 0.977 1.003 1.012 1.182
D/r maximum <0.0005 <0.0005 <0.0005 <0.0005
D/r mean <0.0005 <0.0005 <0.0005 <0.0005
Dqr (minimum, maximum) (e Å3) 3.76, 3.25 0.35, 0.89 1.50, 1.40 1.09, 1.47

The structures were solved by Patterson methods and re- methyl group. Two tert-butyl moieties of 4 (C37, C47) were
fined on F2 by full-matrix least-squares using all reflections found to have rotational disorder; each of the six methyl
with I > 0, anisotropic displacement parameters and weights groups were placed in two alternative positions with comple-
2 1
w ¼ ½r2 ðF 2o Þ þ ðaP Þ þ bP  , with P ¼ ðF 2o þ 2F 2c Þ=3. The mentary occupancies and refined with restraints in the
hydrogen atoms were placed in calculated positions using a C(tert)–C(Me) and C(Me)–C(Me) distances. The final occu-
riding atom model with fixed C–H distances [0.93 Å for pancies were 0.61(2) for C38, C39, C40, and 0.51(2) for C48,
C(sp2), 0.96 Å for C(sp3, CH3), 0.97 for C(sp3, CH2)] and C49, C50. In 3, the displacement ellipsoid of C19 shows a
Uiso = pUeq(parent atom) [p = 1.2 for C(sp2) and 1.5 for very elongated prolate shape, probably due to rotational dis-
C(sp3)]. A single orientation parameter was refined for each order about C14–C17, which could not be satisfactorily
84 H. Arzoumanian et al. / Inorganica Chimica Acta 359 (2006) 81–89

modeled. The final difference Fourier syntheses were feature-


less. The large residual electronic density observed in 2 can be
attributed to a poor absorption correction, due to the large
value of l and the unfavorable specimen morphology.
Software used: (a) data collection and processing: MSC/
AFC Software [21], CAD-4 Software [22], CRYSTAN [23], CRYS-
TALCLEAR [24]; (b) program interfacing: WINGX [25]; (c)
structure solution and refinement: SHELX97 [26]; (d) data
analysis: WINGX [25], PLATON [27]; (e) molecular graphics:
ORTEP-3 [28].
The relevant geometric parameters of the molecules are
given in Table 1 and the structures are shown in Figs. 1–4.
Features common to all compounds:

 In general, the values of bond lengths and angles are


within the expected values [29,30].
 The ranges of the cis [68.9(2)–107.0(4)] and trans
[155.9(3)–169.0(4)] angles reveal a significant distortion
of the ideal octahedral geometry.
 The W–N distances trans to the oxo groups are longer
than expected, due to the strong trans effect of the oxo
groups [31] (this is evident in 2, where the axial W1–
N1 is significantly longer than the equatorial W1–N2).
 The O–W–O is the largest cis angle, due to the mutual
repulsion of the oxygen atoms.
 The byte angle N(L)–W–N(L) is the narrowest cis angle.
 The axial ligands are bent away from the oxo ligands,
which exert the stronger ligand–ligand repulsion.
 The L ligand (L = 4,4 0 -di(tert-butyl)-2,2 0 -bipyridyl)
shows a significant deviation from planarity, as usually
found in bipyridine metal complexes [32].
 The crystal packing is entirely due to van der Waals Fig. 2. Molecular structure of the anionic moiety of 1 showing the atomic
forces. numbering (only for the atoms in the asymmetric unit; symmetry-
generated atoms are not labelled). The displacement parameters are
drawn at 30% probability.

3. Results and discussion

3.1. Synthesis and properties

3.1.1. Bis(tetraphenylphosphonium)dioxo(tetrathiocyanato-
N) tungstate(VI) (1)
The dioxotungstate anion was obtained by the same
procedure reported for the molybdenum analog and iso-
lated as its tetraphenylphosphonium salt by a metathetic
cation exchange under phase transfer conditions [7]. It
exhibits in infrared spectroscopy two strong absorption
bands at 941 and 895 cm1 attributed, respectively, to
masym W@O and msym W@O of the cis-WO2 unit. When
compared to the corresponding molybdenum complex
the slight shift to higher frequencies could be indicative
of a tungsten-oxo function having some ‘‘triple’’ bond
Fig. 1. Molecular structure of 2 showing the atomic numbering (only for character. The bands at 2104 and 2057 cm1 assigned
the atoms in the asymmetric unit; symmetry-generated atoms are not to the thiocyanato ligand are comparable to those ob-
labeled). The displacement parameters are drawn at 30% probability. served for the molybdenum analog and suggest a N-
H. Arzoumanian et al. / Inorganica Chimica Acta 359 (2006) 81–89 85

Fig. 3. Molecular structure of 4 showing the atomic numbering. The displacement parameters are drawn at 30% probability.

bound species. This was confirmed by an X-ray analysis. cleus. The greater O–W–O angle could also be a conse-
The crystals of 1 contain [PPh4]+ and [WO2(NCS)4]2 quence of the bigger tungsten atom.
ions and CH2Cl2 molecules of solvation. The crystals
are isostructural with the analogous Mo complex [7] 3.1.2. Dioxo(dithiocyanato-N)-4,4 0 -tert-butyl-2,2 0 -
(an enantiomorphous specimen was reported in that bipyridyltungsten(VI) (2)
study). The cation lies on a general position, but the In a manner comparable to the molybdenum analog, the
W atom lies on a twofold axis (along [1 1 0], Wyckoff site bipyridyltungsten(VI) complex was obtained by reacting 1
a) and, therefore, the anion displays overall 2 (C2) crys- with the bidentate ligand. The colorless crystalline solid
tallographic point symmetry and only on half of it is exhibited in infrared spectroscopy two strong absorption
crystallographically independent. The C atom of the sol- bands at 952 and 912 cm1, which correspond to a shift
vent molecule also lies on a Wyckoff site a. The solvent to higher frequencies when compared to the molybdenum
molecule was assigned occupancy 0.64 to account for the complex (936 and 904 cm1) indicating, as in the case of
large displacement parameters. The experimental details 1, a more M„O character plausibly due to the higher elec-
for the analysis of 1 are given in Table 1. The final tropositivity of tungsten. The 1H NMR spectrum for the
atomic coordinates and selected bond distances and an- bipyridyl moiety showed, as expected, a shift at lower field
gles are presented in Table 2. The perspective drawing when compared to the free bipyridine compound. How-
in Fig. 1 shows indeed the ligand to be N-bound in a ever, this downshift, attributed to an electron flux from
somewhat distorted octahedral structure. The W@O the ligand to the metal, was less extensive than the one ob-
(1.71 Å) is slightly longer than the analogous Mo@O served for the molybdenum analog, which might seem
(1.69 Å). This is contrary to what one would expect from unexpected for a more electropositive tungsten atom. A
the infrared data, and also if one considers the higher plausible rationalization could be the more pronounced
electropositivity of tungsten, but the relative longer W– electron donating contribution of the oxo moiety in the
O could simply be due to the intrinsic size of tungsten case of 2, rendering the participation of the nitrogen ligand
since bond distances are established from nucleus to nu- less important.
86 H. Arzoumanian et al. / Inorganica Chimica Acta 359 (2006) 81–89

Fig. 4. Molecular structure of 3 showing the atomic numbering. The displacement parameters are drawn at 30% probability.

The X-ray analysis showed a structure quite similar to to higher frequency compared to the molybdenum analog
the molybdenum analog with a N-bound thiocyanato li- (936 and 904 cm1) indicates as for 2 a higher M„O char-
gand. The crystals of 2 contain only molecules of acter. The extend of this shift is, however, lower for 3,
[WO2(NCS)2L] and are isostructural with the analogous which is in accord with the higher e donor character for
Mo complex [10]. The W atom and the NCS groups lie Cl compared to SCN. The 1H NMR indicates a shift at
on the mirror plane (perpendicular to [0 1 0], Wyckoff site lower field for the bipyridyl ligand, but, as for 2, this shift
c). Therefore, the complex displays m (Cs) crystallographic is less important than expected for a tungsten complex. The
point symmetry, and the asymmetric unit consists of half a plausible higher contribution of the oxo moiety as an e
molecule. The experimental details are given in Table 1 and donor in 3 might be the reason of the small difference ob-
the final atomic coordinates and selected bond distances served in the NMR spectrum of 3 when compared to its
and angles are presented in Table 2. The main features con- molybdenum analog.
cern the W@O (1.715 Å) which is slightly longer than The crystals of 3 contain only molecules of [WO2Cl2L].
Mo@O (1.70 Å) and somewhat unexpected for a more elec- The crystal structure is similar to that of [MoO2Br2L] [8],
tropositive tungsten atom. As mentioned above, this could but the Mo complex crystallizes in space group Pnma con-
be due to the size of the heavier metal. On the other hand, trary to a similar compound bearing an unsubstituted
the W–N (bipy) distance (2.27 Å), compared to the Mo– bipyridine ligand and reported as crystallizing in spacial
N(bipy) (2.29 Å) is in accord with the 1H NMR results. group P 1 [33].
This emphasizes the good r-donor and poor p-donor char-
acter of the nitrogen heterocycle [17]. 3.1.4. l-oxobis[cis-dioxo(thiocyanato-N)-4,4 0 -di-tert-butyl-
2,2 0 -bipyridyltungsten(VI)] (4)
3.1.3. Dichloro(dioxo)-4,4 0 -tert-butyl-2,2 0 - The l-oxo dimeric tungsten complex was obtained using
bipyridyltungsten(VI) (3) the same procedure as for the molybdenum analog [34]. It
The dichloro complex analogous to 2 was prepared di- exhibits in infrared spectroscopy two strong bands at 959
rectly from sodium tungstate and trimethylchlorosilane and 909 cm1 attributed to terminal W@O. These values
[18]. The white crystalline solid exhibited in infrared spec- at higher frequencies when compared to the analogous
troscopy two strong absorbtion bands at 944 and molybdenum complex (940 and 905 cm1) follow the same
911 cm1 attributed to the tungsten-oxo function. The shift trend as 1–3 and argues for a W@O function with higher
H. Arzoumanian et al. / Inorganica Chimica Acta 359 (2006) 81–89 87

Table 2 Table 2 (continued)


Selected bond lengths (Å) and angles ()
O4–W2–O5 101.2(3)
Compound 2 N31–W2–N41 69.8(2)
W1–O1 1.712(6) N31–W2–N2 80.8(3)
W1–N1 2.020(12) N31–W2–O5 79.4(2)
W1–N11 2.282(7) N41–W2–N2 75.7(3)
W1–N2 2.043(13) N41–W2–O5 87.5(2)
O3–W2–N41 159.4(3)
O1–W1–O1i 107.0(4)
O4–W2–N31 159.8(3)
O1–W1–N1 96.0(3)
N2–W2–O5 157.5(3)
O1–W1–N2 96.8(3)
O1–W1–N11 92.1(3)
Compound 3
O1–W1–N11i 160.9(3)
W1–O1 1.701(7)
N11–W1–N11i 68.9(3)
W1–N11 2.291(6)
N11–W1–N1 80.9(3)
W1–Cl1 2.359(2)
N11–W1–N2 81.4(3)
W1–O2 1.705(6)
N1–W1–N2 158.4(5)
W1–N21 2.285(7)
W1–Cl2 2.352(2)
Compound 1
W1–O1 1.710(7) O1–W1–O2 106.6(3)
W1–N1 2.075(11) O1–W1–N11 92.0(3)
W1–N2 2.224(11) O1–W1–Cl1 96.4(3)
O1–W1–Cl2 94.9(3)
O1–W1–O1ii 101.2(6) N11–W1–Cl1 82.51(17)
O1–W1–N1 95.7(4) N11–W1–Cl2 81.14(16)
O1–W1–N1ii 95.2(4) O1–W1–N21 160.8(3)
O1–W1–N2ii 89.8(4) Cl1–W1–Cl2 160.36(8)
O1–W1–N2 169.0(4) N11–W1–N21 68.9(2)
N2–W1–N2ii 79.3(5) O2–W1–N21 92.5(3)
N2–W1–N1 83.8(3) O2–W1–Cl1 95.3(2)
N2–W1–N1ii 83.0(3) O2–W1–Cl2 96.8(2)
N1–W1–N1ii 162.9(5) N21–W1–Cl1 82.04(17)
N21–W1–Cl2 81.98(16)
Compound 4 O2–W1–N11 161.4(3)
W1–O1 1.710(5)
Symmetry codes: (i) x, 0.5  y, z; (ii) 1  y, 1  x, 0.5  z.
W1–O2 1.711(6)
W1–N11 2.279(6)
W1–N21 2.291(6)
W1–N1 2.087(8)
triple bond character. The absorption band at 813 cm1 as-
W1–O5 1.881(6) signed to a l-oxo function is also shifted to higher fre-
W2–O3 1.703(6) quency (m Mo–O–Mo 775 cm1) and supports clearly the
W2–O4 1.707(6) dimeric nature of the complex. The 1H NMR analysis con-
W2–N31 2.307(6) firmed it and furthermore showed the presence in solution
W2–N41 2.319(7)
W2–N2 2.096(8)
of the same l-oxo conformational equilibrium as observed
W2–O5 1.875(6) for the molybdenum analog [34] (Fig. 5). Indeed the spec-
trum exhibited in the aromatic region one ‘‘set’’ of bipyr-
O1–W1–O2 106.0(3)
O1–W1–N1 93.7(3)
idyl signals integrating for 75% and four separate and
O1–W1–O5 101.2(3) distinct ‘‘sets’’ integrating for 25%. Five signals were also
O1–W1–N11 91.4(3) observed and attributed to tertiary butyl groups, one inte-
O2–W1–N21 92.7(2) grating for 75% and the four others of equal intensity for a
O2–W1–N1 93.8(3) total of 25%. The interpretation is founded upon analogy
O2–W1–O5 100.1(3)
N11–W1–N21 69.3(2)
with the molybdenum complex; the ‘‘set’’ integrating for
N11–W1–N1 78.9(3) 75% corresponds to a species having four equivalent nitro-
N11–W1–O5 81.9(3) gen heterocycles (with a linear W–O–W) and the others
N21–W1–N1 78.1(3) integrating for 25% to the conformer having all four pyr-
N21–W1–O5 81.7(2) idyl ring in an anisotropically different environment (with
O1–W1–N21 160.0(3)
O2–W1–N11 161.6(2)
a bent W–O–W). The difference with the molybdenum
N1–W1–O5 155.9(3) complex resides mainly on the equilibrium mixture in solu-
W1–O5–W2 172.5(3) tion(75/25 versus 44/56) but also on the difference of aniso-
O3–W2–O4 106.7(3) tropic environment for each hydrogen atom in the ‘‘bent’’
O3–W2–N2 91.0(3) configuration. For example in the case of tungsten the three
O3–W2–O5 100.5(3)
O3–W2–N31 92.9(3)
a-hydrogens of the bipyridyl ligand resonates at 9.46 ppm
O4–W2–N41 90.1(3) and the fourth one is observed at 8.75 ppm whereas for
O4–W2–N2 93.7(3) molybdenum these four a-hydrogens resonate as four
88 H. Arzoumanian et al. / Inorganica Chimica Acta 359 (2006) 81–89

Fig. 5. 1H NMR of compound 4.

distinct bands (9.40, 9.36, 8.78, 8.51). Any rationalization and the dimeric complexes, however, the difference in reac-
of these observations would be at present for the least tivity within the set of tungsten compounds was much less
speculative. pronounced. What is more noteworthy, on the other hand,
The crystals of 4 contain molecules of [WO(NCS)L]2O is the very large difference in oxygen atom transfer capabil-
and NCMe molecules of solvation. Both metal atoms dis- ity between molybdenum and tungsten analogs. Whereas a
play similar environments and are linked by a non-linear complex such l-oxobis[cis-dioxo(thiocyanato-N)-4,4 0 -di-
symmetrical (within experimental error) oxo bridge. tert-butyl-2,2 0 -bipyridylmolybdenum(VI)] would necessi-
tate stop flow methods for an accurate rate measurement
4. Reactivity towards triphenylphosphine for triphenylphosphine oxide formation, complexes 1–4
were found to be extremely less reactive. The difference in
The properties of complexes 1–4 as oxygen atom trans- reactivity was so great that an estimation of the order of
fer agents towards triphenylphosphine were compared with magnitude could only be done. Thus, the monomeric and
the molybdenum analogs. the dimeric tungsten complexes were estimated as being
For the anionic complex 1, the O-transfer was followed 104 to 105 times less reactive than their molybdenum
spectrophotometrically in CH2Cl2 solution. Pseudo first or- analog.
der conditions were used at room temperature, monitoring
the increase in absorbance at 405 nm. An isobestic point
was observed at 368 nm. The results illustrated in Fig. 6
indicate compound 1 to be about 200 times less reactive
than its molybdenum analog (kobs: 104 versus 17 · 102)
[9]. As in the case of the isostructural molybdenum com-
plex [7], the stoichiometric oxidation of PPh3 to OPPh3
by 1 produces a monomeric W(IV)@O reduced complex
indicating that the comproportionation reaction to W(V)-
l-oxo dimer is inhibited for electronic reasons [35].
The oxygen-atom transfer was also studied under cata-
lytic conditions for complexes 1–4. The experiments were
run in d6DMSO in the presence of a tenfold excess of tri-
phenylphosphine and followed by 31P NMR spectroscopy,
under the same conditions as those reported for the molyb-
denum analog [4,7]. All four complexes were found to be
less active than their molybdenum analog.
As in the case of molybdenum, the anionic complex was Fig. 6. Time evolution of the reaction of 1 with PPh3 (W/PPh3 = 1/200) in
found to be less active than both the neutral monomeric dichloromethane at 24 C monitored spectrophotometrically.
H. Arzoumanian et al. / Inorganica Chimica Acta 359 (2006) 81–89 89

5. Supplementary material [14] C. Lorber, J.P. Donahue, C.A. Goddard, E. Nordlander, R.H.
Holm, J. Am. Chem. Soc. 120 (1998) 8102.
[15] K.M. Sung, R.H. Holm, Inorg. Chem. 39 (2000) 1275, and leading
Supplementary crystallographic data for the structural references therein.
analyses have been deposited with the Cambridge Crystal- [16] R. Noyori, M. Aoki, K. Sato, J. Chem. Soc., Chem. Commun. (2003)
lographic Data Centre under the numbers: CCDC No. 1977, and references therein.
268192 for 1, CCDC No. 268191 for 2, CCDC No. [17] A. Thapper, J.P. Donahue, K.B. Musgrave, M.W. Willer, E.
268194 for 3 and CCDC No. 268193 for 4. Nordlander, B. Hedman, K.O. Hodgson, R.H. Holm, Inorg. Chem.
38 (1999) 4104.
[18] M. Miao, M.W. Willer, R.H. Holm, Inorg. Chem. 39 (2000) 2843.
Acknowledgments [19] T. Ben Hadda, H. Le Bozec, Polyhedron 7 (1988) 575.
[20] H.D. Flack, Acta Crystallogr. A39 (1983) 876.
This work was done under the auspices of the project [21] Molecular Structure Corporation, MSC/AFC, Diffractometer Con-
ECOS-Nord France-Venezuela and Fundayacucho. Finan- trol Software, Molecular Structure Corporation, The Woodlands,
TX, USA, 1993.
cial support is kindly acknowledged. [22] Enraf-Nonius, CAD-4 Software, Version 5, Enraf-Nonius, Delft, The
Netherlands, 1989.
[23] H. Burzalff, R. Böhme, M.C. Gomm, A Crystallographic Program
References System for Minicomputers, University of Erlangen, Germany, 1997.
[24] Rigaku/MSC, Inc., CRYSTALCLEAR, Software Users Guide, version
[1] R.H. Holm, Chem. Rev. (1987) 1401. 1.3.6, The Woodlands, TX, USA, 2000.
[2] F. Bottomley, L. Sutin, Adv. Organomet. Chem. 28 (1998) 339. [25] L.J. Farrugia, J. Appl. Crystallogr. 32 (1999) 837.
[3] R.H. Holm, Coord. Chem. Rev. 100 (1990) 183. [26] G.M. Sheldrick, SHELX97. Programs for Crystal Structure Analysis,
[4] H. Arzoumanian, Coord. Chem. Rev. 178 (part 1) (1998) 191. Release 97-2, University of Göttingen, Germany, 1997.
[5] I.A. Weinstock, Chem. Rev. 98 (1998) 113. [27] A.L. Spek, J. Appl. Crystallogr. 36 (2003) 7.
[6] I.V. Kozhevnikov, Chem. Rev. 98 (1998) 171. [28] L.J. Farrugia, J. Appl. Crystallogr. 30 (1997) 565.
[7] H. Arzoumanian, R. Lopez, G. Agrifoglio, Inorg. Chem. 33 (1994) 3177. [29] F.H. Allen, O. Kennard, D.G. Watson, L. Brammer, A.G. Orpen,
[8] H. Arzoumanian, G. Agrifoglio, H. Krentzien, M. Capparelli, J. R. Taylor, J. Chem. Soc., Perkin Trans. II (1987) S1.
Chem. Soc., Chem. Commun. (1995) 655. [30] A.G. Orpen, L. Brammer, F.H. Allen, O. Kennard, D.G. Watson,
[9] H. Arzoumanian, G. Agrifoglio, H. Krentzien, New J. Chem. 20 R. Taylor, J. Chem. Soc. Dalton (1989) S1–S83.
(1996) 699. [31] B.J. Coe, S.J. Glenwright, Coord. Chem. Rev. 203 (2000) 5.
[10] H. Arzoumanian, L. Maurino, G. Agrifoglio, J. Mol. Catal. A: [32] A. Hazell, Polyhedron 23 (2004) 2081.
Chem. 117 (1997) 471. [33] W.A. Herrmann, W.R. Thiel, E. Herdtweck, Chem. Ber. 123 (1990) 271.
[11] A. Kletzin, M.W.W. Adams, FEMS Microbiol. Rev. 18 (1996) 5. [34] H. Arzoumanian, R. Bakhtchadjian, G. Agrifoglio, H. Krentzien,
[12] M.K. Johnson, D.C. Rees, M.W.W. Adams, Chem. Rev. 96 (1996) J.C. Daran, Eur. J. Inorg. Chem. (1999) 2255.
2817. [35] J.H. Enemark, J.A. Coney, J.J. Wang, R.H. Holm, Chem. Rev. 104
[13] W.R. Hagen, A.F. Arendsen, Struct. Bond. 90 (1998) 161. (2004) 1175.

You might also like