Você está na página 1de 18

Coordination Chemistry Reviews 268 (2014) 83–100

Contents lists available at ScienceDirect

Coordination Chemistry Reviews


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

Review

Monomeric and dimeric oxido–peroxido tungsten(VI) complexes in


catalytic and stoichiometric epoxidation
Mojtaba Amini a,∗ , Mohammad Mehdi Haghdoost b , Mojtaba Bagherzadeh b,∗
a
Department of Chemistry, Faculty of Science, University of Maragheh, P.O. Box 55181-83111, Maragheh, Iran
b
Chemistry Department, Sharif University of Technology, P.O. Box 11155-3615, Tehran, Iran

Contents

1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 83
2. Homogeneous epoxidation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 84
3. Heterogeneous epoxidation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 91
3.1. Polymer-anchored catalysts . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 91
3.2. Silica-anchored catalysts . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 94
3.3. Magnetically recyclable catalysts . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 97
4. Conclusions and outlook . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 98
Acknowledgements . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 99
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 99

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

Article history: Epoxides are important materials in the production of fine chemicals. Selective oxidation of alkenes to
Received 12 November 2013 epoxides remains a challenge in synthetic chemistry. There has been considerable progress in tungsten-
Accepted 2 January 2014 catalyzed epoxidation of alkenes in recent years. This review highlights the use of monomeric and dimeric
Available online 15 February 2014
oxido–peroxido tungsten(VI) complexes as catalysts or oxidants in homogeneous or heterogeneous cat-
alytic systems for epoxidation of olefins.
Keywords:
© 2014 Elsevier B.V. All rights reserved.
Oxido–peroxido
Tungsten(VI)
Epoxidation
Homogeneous
Heterogeneous

1. Introduction

Olefin epoxidation is an important reaction for the production


of epoxides as chemical intermediates to afford useful chemicals
[1–5]. The use of transition metal complexes as catalysts in organic
synthesis is an exciting field of research, and such catalysts are
Abbreviations: 2-pybmz, 2-(2-pyridyl)benzimidazole; 3-pybmz, 2-(3- extensively utilized in a wide range of areas [5–10]. In general, in
pyridyl)benzimidazole; DCE, dichloroethane; DMF, dimethylformamide; ee, the presence of hydrogen peroxide, metals form oxido–peroxido
enantiomeric excess; dppmO2 , bis(diphenylphosphineoxide)methane; HA, complexes, which can be used in the synthesis of a number of
hydroxyapatite; HBPT, hexabuthylphosphor triamide; HMPT, hexamethylphos-
fine chemicals [11–13]. Most oxidation reactions involving molec-
phor triamide; HMS, hexagonal mesoporous silica; HPEOH, 1-(2 -hydroxyphenyl)
ethanone oxime; IL, ionic liquid; MO-LCAO, molecular orbital-linear combination ular oxygen and metal complexes proceed via inner-sphere or
of atomic orbitals; PMA, polymethacrylate; PS, polystyrene; PTC, phase-transfer addition–elimination mechanisms. Oxido–peroxido complexes of
catalyst; PW2 , [HPO4 {WO(O2 )2 }2 ]2− ; QO, 8-quinolinol; TBA, tetrabutylammonium transition metal are important intermediates for this type of acti-
cation; TBHP, tert-butylhydroperoxide; THA, tetrahexylammonium; TOF, turnover
vation, so interest in them as catalysts in oxidation reactions has
frequency; TON, turnover number; MBO, 3-methyl-3-buten-1-ol.
∗ Corresponding authors. Tel.: +98 421 2278900; fax: +98 421 2276066. increased [14–17].
E-mail addresses: mamini@maragheh.ac.ir (M. Amini), bagherzadeh@sharif.edu The chemistry of transition metal oxido–peroxido complexes
(M. Bagherzadeh). has attracted special attention owing to their importance in a

http://dx.doi.org/10.1016/j.ccr.2014.01.035
0010-8545/© 2014 Elsevier B.V. All rights reserved.
84 M. Amini et al. / Coordination Chemistry Reviews 268 (2014) 83–100

O O 1a; L= H2O
1b; L= DMF
W
1c; L= HMPT
1d; L=Pyridine
O O
1e; L=HBPT
L

Fig. 1. W(VI) oxido–bisperoxido complexes developed by Mimoun et al. [31].


Fig. 2. Intermediates in epoxidation of allylic alcohols by (a) hydroperoxovanadium
and titanium, and (b) 1c complexes.
variety of industrial, pharmaceutical, and biological processes
[18,19]. Coordination to a metal center activates peroxido groups
for oxidation of a variety of substrates, so oxido–peroxido metal different rates. In fact, 1c behaves as a mono-oxygenating reagent
complexes are important as intermediates in biological and syn- toward olefins. Transfer of the first peroxide oxygen is faster than
thetic catalysis [20–23]. that of the second, contrary to behavior observed for the molybde-
Oxido–peroxido complexes of molybdenum and tungsten ions num analog.
[Mo(VI) and W(VI)] with a closed-shell electronic structure and d0 Di Furia et al. reported that epoxidation of the allylic alcohols
configuration are of particular interest. Like molybdenum, tungsten geraniol 3a and linalool 3b with 1c at ambient temperature is
displays high activity in oxidation reactions, and oxido–peroxido remarkably faster than for simple olefins and is regiospecific [33].
complexes of tungsten can be used for stoichiometric oxygen trans- Under the experimental conditions used, highly selective and quan-
fer reactions, including epoxidation of alkenes and oxidation of titative yields of the corresponding monoepoxides were obtained,
organic and inorganic substrates, and can be regarded as an envi- but the regioselectivity was opposite between geraniol and linalool.
ronmentally friendly alternative to traditional oxidation catalysts. Geraniol 3a affords the 2,3-epoxygeraniol 4a, whereas the 6,7-
Structural studies on W(VI) complexes have shown that the epoxy derivative 4b was obtained from linalool 3b (Scheme 1).
metal nucleus is more shielded in oxido–peroxido than in the Interestingly, the opposite regioselectivity was not observed in
corresponding dioxido complex for similar ligands. Accordingly, the same epoxidation reactions of 3a and 3b using similar vanadium
bisperoxido should be more shielded than monoperoxido com- and titanium complexes as oxidant. The hydroperoxo vanadium
plexes [24]. A number of reviews have focused on synthetic and and titanium complexes regiospecifically epoxidize both 3a and 3b
catalytic aspects of polyoxotungstates [25–28]. In 2005, Mizuno at the allylic double bond (Scheme 1) [34,35]. Clearly there must
et al. reviewed H2 O2 -based epoxidation of olefins catalyzed by be some significant difference between the two classes of oxidants.
polyoxotungstates [5]. They concluded that polyoxotungstates are In fact, epoxidation with hydroperoxo V and Ti complexes starts
the best among simple soluble salts of metal oxides for epoxidation, with coordination of the substrate to the metal center to form an
but are sometimes unsuitable for the production of acid sensitive alkoxohydroperoxo complex (Fig. 2a), which favors oxygen transfer
and/or water-soluble shorter-chain epoxides owing to the low pH to the allylic double bond. By contrast, epoxidation with Mo and
in the aqueous phase. We recently reviewed catalytic and stoichio- W peroxido complexes should not involve a coordinated substrate
metric oxidations of oxido–peroxido molybdenum(VI) complexes and the accelerating effect of the hydroxo moiety should be related
[29]. In many ways the behavior of oxido–peroxido tungsten(VI) to intramolecular hydrogen bonding between the OH group of the
complexes is very similar to that of molybdenum complexes [30]. allylic alcohol and the peroxido groups of the tungsten complex
However, unlike molybdenum complexes, the reactivity of tung- (Fig. 2b). This conclusion was supported by the observation that
sten complexes has not been studied much because of their low when alcohols such as CH3 OH and CF3 CH2 OH were added to the
solubility. reaction mixture at low concentration, detectable enhancement of
Here we review monomeric and dimeric oxido–peroxido tung- the epoxidation rate of a simple olefin occurred, thus confirming
sten(VI) catalysis of olefin epoxidation. As mentioned above, olefin that an accelerating effect by a protic species, likely via hydrogen
epoxidation by polyoxotungstates has recently been reviewed [5] bonding of the peroxido oxygen in the transition state, is involved
and is not covered here. The review is organized into two gen- [33].
eral sections, homogeneous catalysis and heterogeneous catalysis. Mimoun-type tungsten complexes act not only as stoichiomet-
Future directions for research in this area are also discussed. ric oxidants but also as catalysts [36]. Oxidation of cyclohexene
and 1-methylcyclohexene by H2 O2 in the presence of cat-
2. Homogeneous epoxidation alytic amounts of 1c was studied at 40 ◦ C. The results showed
quantitative formation of epoxide based on H2 O2 consumed,
In 1969, Mimoun et al. were the first to prepare a series and indicated that under the appropriate experimental condi-
of oxido–bisperoxido W(VI) complexes of the general formula tions, W(VI) is by far a better catalyst of oxidation by H2 O2
[WO(O2 )2 Ln ] 1 (n = 1–2; L = H2 O 1a, DMF 1b, HMPT 1c, pyridine than other metals studied so far [36]. This is because of its
1d, HBPT 1e) (Fig. 1) and evaluate their activity as a stoichiomet- high reactivity for oxidation and inherent poor activity for
ric oxidant in the epoxidation of alkenes [31]. Since then, several decomposition of H2 O2 . Since the reaction was carried out in
molybdenum and tungsten complexes with this general formula, acid medium, the epoxide initially formed undergoes quantita-
known as Mimoun-type complexes, have been investigated. tive acid-catalyzed ethanolysis, yielding 2-ethoxycyclohexan-l-ol
The use of Mimoun-type tungsten complexes as stoichiometric and 2-ethoxy-2-methyl-cyclohexan-l-ol from cyclohexene and
oxidants was the subject of much research in the 1970s and 1980s. 1-methylcyclohexene, respectively [36]. The kinetic studies con-
In 1986, the kinetics of stoichiometric epoxidation of olefins with firmed that similar to Mo(VI) oxido-bisperoxido catalytic systems
1c in DCE was studied [32]. Although the molecular structure of 1c [37], 1-methylcyclohexene (more nucleophilic substrate) reacts
and its molybdenum analog, [MoO(O2 )2 (HMPT)] 2, are very similar, significantly faster, by a factor of approximately 2.5, than does
the oxygen transfer rates are faster from 1c than from 2 by factors cyclohexene.
ranging from five for cyclooctene to 40 for cis-stilbene. In addition, The nature of the species formed in aqueous solutions of
the two peroxido groups in 1c appear to react with the substrates at [WO4 ]2− in the presence of H2 O2 was studied using Raman, IR, and
M. Amini et al. / Coordination Chemistry Reviews 268 (2014) 83–100 85

C3 or hydroperoxovanadium and titanium O


OH OH

3a 4a

HO HO HO

hyd roperoxovanadium and titanium C3


O

4c 3b 4b

Scheme 1. Oxidation of geraniol and linalool by hydroperoxovanadium and titanium and 1c complexes [33–35].

NMR spectroscopy [37,38]. At pH values from 1 to 9 in solutions simple olefins. An alternative approach is to run the epoxidation
containing H2 O2 /W at a ratio of 5:1, various anionic tungsten oxido- reaction under phase transfer conditions. Venturello and cowork-
peroxido species exist. At high pH (>7), [W(O2 )4 ]2− 5 predominates; ers described the catalytic epoxidation of simple cyclic and acyclic
at intermediate pH (1.3–7), dimer [W2 O3 (O2 )4 (H2 O)2 ]2− 6 is alkenes in the presence of a phase-transfer catalyst (PTC) in biphasic
the major species; and at pH < 1.3 in the presence of HCl, cis- 8% H2 O2 /DCE [53]. For this catalytic system, 1-octene was oxidized
[WO(O2 )Cl4 ]2− 7 is formed (Fig. 3). Long before the nature of these to 1,2-octeneoxide in 87% yield. In spite of the acidic conditions,
tungsten oxido–peroxido species was known, it was observed that no hydroxylation product was detected. The reason for this is the
[WO4 ]2− /H2 O2 systems catalyze epoxidation reactions. As early as protective effect of the biphasic system. Furthermore, a significant
1959, an ␣,␤-unsaturated carboxylic acid, maleic acid, was epox- improvement in catalysis can be achieved by simultaneous use of
idized in quantitative yield using H2 O2 and Na2 WO4 at pH 4–5.5 both Na2 WO4 and H2 WO4 as catalysts [54].
[39]. At 65 ◦ C, the epoxidation reactions required approximately The efficiency of biphasic epoxidation by [WO4 ]2− /H2 O2
1.5 h. Under similar reaction conditions, simple olefin substrates strongly depends on the type of phase transfer agent. As an
showed lower reactivity. For example, 2-heptene underwent epox- example, when a quaternary ammonium salt was immobilized on
idation to afford the corresponding epoxide in 55% yield in 6 h. The polymer as PTC, an increase in reaction selectivity compared to a
Na2 WO4 /H2 O2 catalytic system can be applied successfully to all system with non-immobilized salt was observed [55]. Changing
di- and trisubstituted unsaturated acids by slight modification of the PTC can even change the reaction product in some cases. For
the reaction conditions [40]. In addition, [WO4 ]2− -catalyzed epox- oxidation of cyclohexene using 30% H2 O2 , Na2 WO4 , and [CH3 (n-
idation of ␣,␤-unsaturated carboxylic acids to the corresponding C8 H17 )3 N]HSO4 as the PTC, adipic acid was isolated in 93% yield
epoxides was faster than for molybdate [MoO4 ]2− and vanadate instead of epoxide [56,57].
[VO4 ]3− catalysis [41,42]. The dimer [W2 O3 (O2 )4 (H2 O)2 ]2− 6 (Fig. 3) was the first tungsten
Epoxidation of allylic alcohols with aqueous H2 O2 in the pres- oxido–peroxido complex to be characterized by X-ray crystallog-
ence of tungstic acid (H2 WO4 ) as a catalyst has also been studied raphy [45]. In the binuclear structure of [W2 O3 (O2 )4 (H2 O)2 ]2− , the
[46–50]. In contrast to Na2 WO4 , H2 WO4 is a strong Brønsted tungsten oxido–bisperoxido units are linked through an oxo ligand
acid. Oxidation of olefins in acidic solution of H2 WO4 with H2 O2 [45]. In the presence of the non-anionic monodentate and biden-
leads to the formation of diols [51–53]. For selective epoxide for- tate ligands, the W O W bonds cleave to form complexes with
mation, the catalytic reaction should be carried out in buffered separate WO5 units. Gelbard and coworkers prepared a series of
conditions. Epoxidation of olefinic alcohols was achieved in high oxido–bisperoxido tungsten(VI) complexes (8b–11b) with several
yields in buffered water–substrate mixtures with hydrogen perox- neutral organophosphorus(V) and diamine ligands (8a–11a, Fig. 4).
ide and H2 WO4 [48]. Under these reaction conditions, epoxidation Complexes 8b–11b, either preformed or formed in situ, were cat-
occurred with complete retention of the stereochemistry around alytically active in the epoxidation reaction [58,59] (Fig. 5).
the carbon–carbon double bond, and no isomerization, cleavage of Among the catalysts investigated, the highest turnover
the olefinic bond, or overoxidation to diols was observed. frequency was observed for epoxidation of allylic alcohols cat-
Application of the [WO4 ]2− /H2 O2 catalytic system is limited to alyzed by 11b (TOF = 58). Interestingly, the reaction of 6 with
substrates that are soluble in water or water–alcohol mixtures. mono-anionic phosphonic acid ligands 12a and 13a gave the perox-
Consequently, this method is not suitable for the epoxidation of oditungstic complexes 12b and 13b, in which two WO(O2 )2 groups

2- 2- 2-
H2O
O O O
O O O
O
O O
O W O O Cl O
W
O W
O O W Cl Cl
O O
O O Cl
O
O
H2O
5 6 7

Fig. 3. W(VI) oxido–peroxido complexes 5 [43,44], 6 [45], and 7.


86 M. Amini et al. / Coordination Chemistry Reviews 268 (2014) 83–100

Fig. 4. Ligands 8a–14a and W(VI) oxido–bisperoxido complexes 8b–14b [58–60].

are connected by the phosphonic acid ligand (Fig. 4). Under similar 12b and 14b reveals that the latter exhibits lower activi-
reaction conditions, the TOF obtained with 12b (TOF = 88) was even ties. Obviously, the presence of the phenyl rings decreases
higher than that with 11b. The authors attributed this finding to the the catalytic ability of the complex but the reason is not yet
hydrophilic and hydrophobic (amphiphilic) behavior of 12b. Com- clear.
plex 12b has both ionic character and a hydrophobic counter-cation Complexes 15a–c were active species in olefin oxidation and
(Q+ ). During epoxidation it is essentially concentrated in the organic formed epoxides as primary products. The tungsten complex 15a
phase, where the catalytic reaction takes place. After each catalytic reacted much faster than the corresponding molybdenum com-
cycle, it returns to the aqueous H2 O2 phase to regenerate the cat- plex [61]. The exact reaction conditions and epoxidation results
alytically active bisperoxido complex. However, as mentioned later, were not described by the authors, but they mentioned that
it seems to be more appropriate to relate the high activity of 12b the yield and turnover number for epoxidation depend on the
directly to the Lewis acidity of the metal centers. [W]/[ligand]/[PPh4 ] ratio [62]. Noyori et al. published a detailed
In the same year, the activity of 14b was examined in cat- review of the use of H2 O2 , [WO4 ]2− , and WO3 precursors and an
alytic epoxidation of a number of cyclic and linear alkenes in onium salt as a PTC in the oxidation of various organic compounds
a biphasic H2 O2 –benzene mixture [60]. Comparison between [63], so this is not repeated here.
M. Amini et al. / Coordination Chemistry Reviews 268 (2014) 83–100 87

R3 OH

R4 R1

R2 H2O
H2O
O
W
O R3 OH
I
H2O
R4 R1
R4 R1
2 3 R2
R R
a
O H
HOOH O
+
HO 50 <a<70o
o
W

O
W

R3 OH

O
R4 R1

Fig. 5. W(VI) oxido–bisperoxido complexes 15a–c [61,62]. R2

Scheme 2. Mechanism for epoxidation of allylic alcohols by K2 [W2 O3 (O2 )4 (H2 O)2 ]
dimer [68].
For non-surfactant amphiphilic sugars, such as sucrose, l-
arabinose, and methyl or ethyl ␤-d-fructopyranoside, instead of
onium salts, epoxidation of allylic alcohols by H2 O2 in the presence
of [WO4 ]2− salts occurred in neutral aqueous solution [64]. In this epoxidation of 1,3-allylic strained alcohols proceeded diastereos-
catalytic system, an amphiphilic carbohydrate enhances the solu- electively to give the threo diastereoisomer preferentially, but the
bility of the allylic alcohol in water. For example, both geraniol and erythro selectivity was higher for allylic alcohols without 1,3-allylic
nerol substrates have low solubility in pure water but are highly strain. The stereochemical results suggest that the dihedral angle
soluble in 1 M sucrose solution. Using [WO4 ]2− salt as a catalyst between the ␲ plane of the C C double bond and the hydroxyl
and H2 O2 as a terminal oxidant, complete conversion of isophorol group of the allylic alcohol in the transition-state geometry is
(3,5,5-trimethyl-2-cyclohexen-1-ol) to its corresponding epoxide 50–70◦ (Scheme 2). On the basis of these results, in addition to
was achieved for reactions in the presence of carbohydrates. In the kinetic and spectroscopic results, the authors proposed a mech-
absence of carbohydrates, epoxidation of acyclic alcohols gave mix- anism for catalytic epoxidation of allylic alcohols by 6 [18]. In
tures of erythro and threo epoxides, but epoxidation in glycosidic this mechanism the allylic alcohol substrate coordinates to the
aqueous media proceeded with remarkable stereo-discrimination tungsten center by dislodging water that is already coordinated
in favor of erythro-enriched epoxy alcohols. Hydrophobic interac- (metal–alcoholate binding mechanism) (Scheme 2). This mecha-
tions between the most hydrophobic face of the substrate and that nism is incompatible with the mechanisms proposed by Di Furia
of the sugar additive are responsible for the stereo-discrimination et al. for epoxidation of allylic alcohols by [WO(O2 )2 (HMPT)] (1c,
observed. Fig. 2b) [33]. The mechanism suggested by Di Furia et al. involves
In 1986, Mimoun et al. prepared [Ph3 PCH2 Ph]2 [6] (Fig. 3) and hydrogen bond formation between the substrate and the peroxido
used it as an effective catalyst for epoxidation of alkenes in a groups the of tungsten complex (hydrogen-bonding mechanism).
water–organic mixture [65]. In this catalytic system, epoxidation We surmise from the results that the distinction between these
proceeds in the organic phase, where the tungsten complex resides two mechanisms lies in the existence or absence of an easily
because of the lipophilicity of the Ph3 PCH2 Ph cation. Similar to replaceable ligand. The water ligand of the K2 [W2 O3 (O2 )4 (H2 O)2 ]
other reports [33], allylic alcohols were rapidly epoxidized but con- complex can be easily replaced by allylic alcohols, whereas the
jugated double bonds were not very reactive toward epoxidation. HMPT ligand of the [MoO(O2 )2 (HMPT)] complex cannot be replaced
Mizuno and coworkers showed that the potassium salt of dimer by other monodentate ligands. Competitive epoxidation of homoal-
6 (K2 [6], Fig. 3) could act as an effective catalyst for chemo-, regio- lylic and allylic alcohols catalyzed by [SeO4 {WO(O2 )2 }2 ]2− 16 also
, and diastereo-selective and stereospecific epoxidation of allylic revealed that a metal–alcoholate species is not formed in the tran-
alcohols using H2 O2 in water without PTCs such as onium salts sition state for 16-catalyzed epoxidation of allylic alcohols because
[66,67]. Oxidation of 1-propen-3-ol in water using complex K2 [6] the W atoms in 16 does not contain an easily replaceable ligand
as a catalyst proceeds smoothly with a high rate of conversion (hydrogen-bonding mechanism) [69].
and selectivity (TON = 4900). Retention of the dimer structure of Preliminary coordination of the substrate to the W(VI)
6 was suggested by IR, UV/Vis, and 183 W NMR spectra of recovered bisperoxido complex was also proposed for epoxidation of cis-1-
catalysts [66]. propenylphosphonic acid [70]. However, 13 C NMR showed that
The stereochemistry observed for catalytic epoxidation of 1,3- ␣,␤-unsaturated carboxylic acids cannot coordinate to the W(VI)
allylic alcohols by K2 [6] was rather surprising [68]. Catalytic center of the bisperoxido complex prior to oxygen transfer. This
88 M. Amini et al. / Coordination Chemistry Reviews 268 (2014) 83–100

O O O O
O O O O O
O
W W O W
O O W
O O O O
O
O O O O
O
25 26
Fig. 7. Dinuclear oxido–peroxido W(VI) complexes 25 and 26 [20].

activity of peroxido W(VI) complexes is not limited only to epoxida-


tion reactions. In 1994, Reynolds et al. reported that the bisperoxido
tungsten complex [WO(O2 )2 (C2 O4 )]2− 18 oxidizes bromide ions at
Fig. 6. W(VI) oxido–bisperoxido complexes 17a and 17b [74]. a faster rate than the molybdenum(V1) counterpart does [79].
Various W(VI) complexes (19–24) containing oxido, peroxido,
and 8-quinolinol ligands were prepared by Bhattacharyya et al.
indicates the role of the functional group conjugated to the olefinic [80,81]. The preparation procedure is shown in Scheme 3. Stoi-
double bond in the epoxidation mechanism. chiometric epoxidation of alkenes was carried out using 20 and
Deprotonation of the OH bond of an allylic alcohol followed 22 complexes in acetonitrile. When 20 was allowed to react stoi-
by proton transfer to the peroxido ligand occurs when the allylic chiometrically with cyclopentene in a 1:1 ratio, complex 21 formed
alcohol coordinates to the tungsten ion (Scheme 2). The epoxy alco- in addition to cyclopentene oxide. Under similar reaction condi-
hol and tungsten dioxido-monoperoxido complex are formed via tions, 22 reaction with cyclopentene in molar ratios of 1:1 and
an oxygen transfer reaction. Then the H2 O2 oxidizes the dioxido 1:2 led to significant amount of epoxide and complexes 23 and 24,
species back to bisperoxido, which reenters the cycle and contin- respectively.
ues the process [68]. The same mechanism was also identified in the All complexes 19–24 were catalytically active for olefin epoxida-
oxidation of sulfides [71,72] and bromide ions [73] by monomeric tion. Comparative studies for epoxidation of cyclohexene revealed
and dimeric oxido-peroxido tungsten(VI) complexes. that the catalytic activity increases in the order 23 < 20 < 22 < 19.
In 2000, the chiral oxido–bisperoxido W(VI) complexes 17a Compared to monoperoxido tungsten catalysts (20 and 23), the
and 17b (Fig. 6) were successfully used by Yoon et al. in bisperoxido complexes (19 and 22) showed significantly higher
the catalytic epoxidation of styrene derivatives using TBHP as catalytic activity. Using complex 22 as a catalyst, epoxidation of
a terminal oxidant [74]. In particular, catalytic epoxidation of cyclohexene at room temperature in the presence of H2 O2 as oxi-
trans-␤-methylstyrene at room temperature using 17b afforded dant and NaHCO3 as promoter yielded cyclohexene oxide with TOF
(1S,2S)-trans-methylstyrene oxide in 41% yield and 63% ee. How- as high as 5460 h−1 . Catalyst 22 without NaHCO3 as co-catalyst
ever, for styrene as substrate the chemical yield (29%) and showed very low activity. It is known that H2 O2 and bicarbonate
enantioselectivity (4%) were quite low. The authors noted that react in an equilibrium process to produce peroxymonocarbon-
tungsten complexes 17a and 17b were more active than molyb- ate (HCO4 − ), which is a more reactive nucleophile than H2 O2 and
denum(VI) analogs in the epoxidation reaction. This finding seems speeds up epoxidation reactions [82,83]. The authors noted that the
to conflict with many studies that have reported higher catalytic epoxidation rate depends strongly on the substrate concentration,
activity of tungsten compared to molybdenum peroxido complexes bicarbonate concentration, and reaction temperature.
[29]. In 2006, Jimtaisong and Luck reported that W-based diox- Two dinuclear oxido–peroxido W(VI) com-
ido and oxido–peroxido catalysts gave higher epoxide conversions plexes, (TBA)2 [{WO(O2 )2 }2 (␮-O)] ((TBA)2 [25]) and
than Mo-based catalysts in H2 O2 systems. By contrast, Mo-based (TBA)2 [{WO(O2 )2 }2 (␮-O2 )] ((TBA)2 [26]; Fig. 7), were synthe-
catalysts were more reactive than the W counterparts in TBHP sized by Hazarika et al. [73]. In principle, the structure of anion 25
systems [75]. Such considerable differences may result from the is similar to the well-known dinuclear oxido–peroxido W(VI) com-
different ability of the molybdenum and tungsten catalysts to acti- plex 6 (Fig. 3), except that in 25 the aqua ligands are eliminated.
vate TBHP and H2 O2 . W(VI) is better able to activate H2 O2 than Reaction of 25 with H2 O2 gives 26, which contains three different
Mo(VI) is, whereas Mo(VI) is more active in epoxidation with TBHP peroxido ligands (␩2 , ␩1 :␩2 , and ␩1 :␩1 ). In acetonitrile media,
than W(VI) is. stoichiometric epoxidation of various alkenes with (TBA)2 [26]
Almost all of the tungsten peroxido complexes show higher produced quantitative amounts of the corresponding epoxides,
reactivity for stoichiometric epoxidation than chromium and but stoichiometric epoxidation with (TBA)2 [25] hardly proceeded.
molybdenum counterparts. The activation energy calculated for Complex 25 alone was only able to epoxidize allylic alcohols.
ethylene epoxidation with Mimoun-type [MO(O2 )2 (OP(CH3 )3 )] IR, Raman, and UV–vis spectroscopic studies revealed that 26
complexes decreases from Cr (23.1 kcal/mol) to Mo (17.9 kcal/mol) is converted to 25 by reaction with olefin and 26 can easily be
to W (15.4 kcal/mol). Density functional theory (DFT) analysis regenerated by reaction of 25 with H2 O2 . Taken together, these
reveals that the donation/backdonation (d/b) ratio, which is a mea- results are consistent with 26 as the active oxidant and 25 as the
sure of the electronic character of the oxidants, should be a measure precursor. The oxidation ability of 26 is interesting, since (␩1 :␩1 -
of the epoxidation activity in such complexes [76]. The d/b ratio and peroxido)dimetal complexes are usually inactive for oxygenation
therefore the electrophilicity increases in the order Cr < Mo < W. of organic substances.
Since oxido–peroxido complexes of VIb transition metals attack Although the dinuclear oxido–peroxido W(VI) anion 25 is not an
olefins in an electrophilic manner [77], the epoxidation reactivity active oxidant, the THA salt of this anion is an efficient catalyst for
of these complexes correlates with their electrophilicity. oxidation of alkenes with H2 O2 in acetonitrile [85]. Under the con-
In another study, Rösch and co-workers suggested a linear ditions used, cis-cyclooctene was completely transformed to the
correlation between the O O  * -orbital energy level and the activa- corresponding epoxide after 4 h of reaction at reflux temperature
tion barrier for ethylene epoxidation with [MO(O2 )2 (NH3 )] (M = Cr, (TON = 600). Catalytic epoxidation by anion 25 was dramatically
Mo, W) complexes: the higher the energy of the ␴* (O O) orbital, accelerated by a factor of approximately 60 on addition of HNO3
the higher is the activation barrier [78]. The superior catalytic or HClO4 [86]. Mizuno et al. showed that a tetranuclear W(VI)
M. Amini et al. / Coordination Chemistry Reviews 268 (2014) 83–100 89

O
O O
O O O
H W crystallization N
O O H from acetonitrile O W
excess H2O2 O O
H2WO4.2H2O N N
CH3COOH (4 M) O
N

19 20

O O O
O
O O O
O
N
O W O W O excess H2O2 O W
PPh4 O PPh4 O
N N O
O N

21
23 22

O O
W
PPh4 O
N

24

Scheme 3. Stepwise synthesis of catalysts 19–24 [80].

peroxido complex was synthesized by reaction of 25 with an acid.


protic aprotic
The tetranuclear complex has three equivalents of active oxygen
for epoxidation of cyclooctene and its TOF is much greater than for
other di- and tetranuclear peroxotungstates [87].
Immobilization of catalysts in a room-temperature (RT) ionic + N () [W2O11] N () [W2O11]
5 +
liquid (IL) is a promising approach to achieve recyclable catalysts HN 5
N
[88,89]. Using the anionic peroxotungstate complexes described
2 2
above as an anion component of IL, it is possible to obtain a recy- [HHIm]2[25] [HMIm]2[25]
clable catalyst for epoxidation of alkenes. With this approach, two
protic and two aprotic ionic liquid containing [W2 O11 ]2− anion 25
(Fig. 8) were used as catalysts for epoxidation of cyclooctene in dif-
ferent media [90]. An interesting point is that the solvents had a
significant effect on the reaction efficiency and repeatability. The
catalysts exhibited better catalytic performance in polar organic + N () [W2O11] + N () [W2O11]
solvents, which makes H2 O2 utilization quite advantageous, but the 11 11
HN N
conversion rapidly decreased after the first cycle in organic solvents
2 2
(catalyst leaching). Use of mixed solvents addressed the problem of [HDIm]2[25] [DMIm]2[25]
catalyst recovery; for CH3 OH–H2 O, cyclooctene conversion remain
as high as 99.6% with > 99% selectivity, even after six consecutive Fig. 8. Ionic liquids containing anion 25 [90].
cycles.
90 M. Amini et al. / Coordination Chemistry Reviews 268 (2014) 83–100

O O
N O N O O
O W
W O
O O N
O N
O O
O O
R

27a; R=H 27e


27b; R=Me (ortho)
27c; R=Me (meta)
27d; R=Me (para)

O
O O
O W
O O O
N O
O O O
W
O
O N
PPh4

28

29

O O
O M
O O
PPh4
N
OH

M=W 30
M=Mo 31

Fig. 9. Oxido–bisperoxido W(VI) complexes 27–31 studied by Bhattacharyya and coworkers [91,92].

Complexes 27a–e showed very similar catalytic activity for reactivity toward less reactive substrates such as linear olefins in
H2 O2 epoxidation of olefins, but 27e was the most active. Thus, comparison to the molybdenum compound. In the case of alcohol-
the electron-donating methyl group did not change the catalyst functionalized olefins, this difference was more marked [92]. Both
efficiency. Using 3–4 equivalents of H2 O2 as a terminal oxidant complexes were capable of furnishing stoichiometric epoxidation.
and in the presence of NaHCO3 as a promoter, 27e showed high Mizuno et al. reported on synthesis of a novel selenium-
catalytic activity and excellent selectivity for cyclohexene epox- containing binuclear peroxido tungsten(VI) complex,
idation (TON = 4300). A unique feature of this catalytic system is (TBA)2 [SeO4 {WO(O2 )2 }2 ] ((TBA)2 [16]) (Fig. 10) [69]. The complex
the very high catalytic selectivity. For almost all olefinic substrates, obtained was not characterized by X-ray analysis but was used
epoxide selectivity was 100%. Reaction of an aqueous solution of in H2 O2 oxidation of olefins to epoxides [69,93]. Oxidation of
Na2 WO4 and an ethanolic solution of hydroxamic acid ligand pro- 3-methyl-3-buten-1-ol (a homoallylic alcohol) with only one
duced dioxido W(VI) complex 28 (Fig. 9), which showed lower equivalent of H2 O2 in the presence of (TBA)2 [16] was carried
catalytic efficiency (TON = 3700) than 27a–e. An attempt to reach out at ambient temperature and the corresponding epoxide was
the bisperoxido [WO(O2 )2 (hydroxamato)] complex failed; instead, obtained in excellent yield. Catalyst (TBA)2 [16] showed TOF as
the catalytically active complex, [PPh4 ][WO(O2 )2 (C6 H5 COO)] 29, high as 150 h–1 for 10-mmol-scale epoxidation of homoallylic
was formed (Fig. 9). Under identical conditions, this anionic oxido- alcohols. The catalytic activity of (TBA)2 [16] was also compared
bisperoxido complex displayed the higher catalytic performance to that of similar complexes containing XO4 n– ligands (X = S, P, As,
for the epoxidation than did 27a-e catalysts (TON = 4450). There is and Si). Among the catalysts tested, the most promising results
an important conclusion, that the bisperoxo tungsten complexes were obtained for (TBA)2 [16]. The epoxide yield decreased with
generally give better results than the monoperoxido complexes in the bond length between the W atom and the O atom of the
the epoxidation reaction. XO4 n− ligand (Se > S > As > P > Si). The same trend was previously
Using the same approach, Bhattacharyya et al. reported that observed for stoichiometric epoxidation of (R)-(+)-limonene [94].
PPh4 [MO(O2 )2 (HPEOH)] complexes (M = W 30, M = Mo 31) (Fig. 9) DFT calculations suggested that the long SeO W bond length
with 1-(2 -hydroxyphenyl) ethanone oxime as ligand are effective results in weak interaction between the SeO4 n− ligand and the
olefin epoxidation catalysts with H2 O2 as the oxidizing agent [92]. dimeric WO(O2 )2 unit, low Lewis acidity of the W centers, and high
The rate of catalytic epoxidation of cyclic olefins was almost the catalytic activity for epoxidation. The (TBA)2 [16] catalyst could be
same for both complexes. The tungsten compound showed higher recycled without any decrease in catalytic activity.
M. Amini et al. / Coordination Chemistry Reviews 268 (2014) 83–100 91

Fig. 10. Proposed structure of the anion part of 16, 32, and 33 [69,95].

The same group reported on a series of DFT studies to iden- [96]. In a recent study, two possible active structures including
tify the exact mechanism for (TBA)2 [16] catalysis of epoxidation peroxo ring and hydroperoxo structures have been investigated
(Scheme 4) [93]. According to their suggested mechanism, an using DFT [97], but the results could not give a clear-cut answer to
alkene is oxidized to the corresponding epoxide via nucleophilic the question of which is the true active structure. In our opinion,
attack of the C C bond by one oxygen atom of the peroxido ligand. there is a strong possibility that both active structures can be
The catalytic cycle is then started again by regenerated anion 16 involved simultaneously in the oxygen transfer step.
with reaction of the peroxido complex and H2 O2 . The activation The Brégeault group showed that the [H2 PO4 {WO(O2 )2 }2 ]−
barriers calculated indicate that the oxygen atoms of the bridging complex 32 (Fig. 10) is an effective catalyst and a stoichiometric
peroxido groups, in particular the oxygen atom bound to a sin- oxidant for epoxidation of (R)-(+)-limonene at room temperature
gle tungsten atom (O1) (Fig. 10), are primarily involved in olefin [98]. With this complex (0.25 mmol), stoichiometric oxidation of
epoxidation. In an ab initio MO-LCAO study in 1997, Fantucci et al. (R)-(+)-limonene (1.5 mmol) in DCM led to the formation of monoe-
showed that ethylene oxidation by [H2 PO4 {WO(O2 )2 }2 ]− complex poxides (0.45 mmol). This means that half the peroxido ligand can
32 (Fig. 10) occurs via the same mechanism [95]. The tendency transfer an oxygen atom to the C C bond. It seems that only bridg-
of the oxygen atoms of the peroxido groups of 32 to react with ing peroxido ligands are involved in the oxygen transfer reaction.
ethylene increases in the order O2 < O3, 4 < O1. The same research group found that a very similar anionic com-
Although the above-mentioned mechanism is generally plex, [HPO4 {WO(O2 )2 }2 ]2− 33 (Fig. 10), was an efficient catalyst for
accepted, studies have revealed contradictory results regarding epoxidation with H2 O2 in the presence of a phase–transfer agent
the active structure of the tungsten peroxido complexes. While in a biphasic H2 O2 –CH3 Cl mixture [94,99].
many studies postulate that the W(O2 ) three-membered ring is In contrast to extensive studies on the catalytic properties
the active group [68,93,95], a few propose that the WOOH group, of bisperoxido complexes of W(VI), there are few studies on
which forms via reaction of W(O2 ) with H2 O2 , is the active group the catalytic activity of W(VI) monoperoxido complexes. The
[WO(O2 )L(CH3 OH)] complex 34 (Fig. 11) was the first Schiff base-
W(O2 ) complex used as a catalyst for epoxidation [100]. Complex
34 showed high activity for a wide range of olefins in the presence
of H2 O2 /NaHCO3 as oxidant, and catalyzed selective epoxidation
of different linear and cyclic alkenes at room temperature. The
catalytic activity of complex 34 for epoxidation was almost iden-
tical to that of the molybdenum analog [MoO(O2 )L(CH3 OH)], and
depended strongly on the reaction conditions [12].
Table 1 summarizes some of the catalytic properties of
monomeric and dimeric oxido–peroxido W(VI) complexes. More
details on the reaction conditions can be found in the relevant
references.

3. Heterogeneous epoxidation

A major drawback of homogeneous catalysts is difficult recovery


from the reaction medium. Precipitation with subsequent recov-
ery, for example by distillation of the reaction products, which
is an energy intensive process, is typically needed for reuse of
homogeneous catalysts. Recognition of this inherent limitation of
homogeneous catalyst led to a surge in activity to develop het-
erogeneous catalysts. The aim is to combine the advantages of
homogenous catalysts (activity and selectivity) with the facility of
heterogeneous systems for catalyst recycling.

3.1. Polymer-anchored catalysts

Scheme 4. Mechanism for epoxidation in the presence of W(VI) complex 15 as Several heterogeneous tungsten(VI) peroxido catalysts have
reported by Mizuno et al. [93]. shown good performance in epoxidation reactions. In 1992,
92
Table 1
Summary of results for stoichiometric and catalytic oxidation of olefins by W(VI) oxido–peroxido complexes.

W complex Substrate Reaction conditions Epoxide Ref. Remarks


yield (%)

1c Cyclooctene DCE, 20 ◦ C, 30 min (stoichiometric 80 [32] Superior oxidizing ability of W(VI) over Mo(VI)
oxidation)
1c Cyclohexene H2 O2 /EtOH:H2 O, 40 ◦ C, CH3 SO3 H 98 [36] Subsequent ethanolysis of the epoxide ring occurs
(0.05 M)
8b Cyclohexene H2 O2 /dioxane, 70 ◦ C, 1.3 h 88 [59] Complex generated in situ
8b 2-Propen-1-ol H2 O2 /DCE:H2 O,70 ◦ C, 3 h 12 [58] Highest TOF observed for 12b
11b 2-Propen-1-ol H2 O2 /DCE:H2 O, 70 ◦ C, 1 h 53 [58]
11b Cinnamyl alcohol H2 O2 /dioxane, 70 ◦ C, 24 h 34 [58]
12b 2-Propen-1-ol H2 O2 /DCE:H2 O, 70 ◦ C, 0.5 h 40 [58]
12b Cinnamyl alcohol H2 O2 /dioxane, 70 ◦ C, 23 h 28 [58]
13b 2-Propen-1-ol H2 O2 /DCE:H2 O, 70 ◦ C, 9 h 50 [58]
14b Cyclohexene H2 O2 /benzene:H2 O, 75 ◦ C, 3 h 8 [60] Low epoxide yield.
[Ph3 PCH2 Ph]2 [6] Cyclooctene H2 O2 /DCE:H2 O, 50 ◦ C, 15 h 88 [65] Reaction is run in nearly neutral conditions and no need to buffer the aqueous

M. Amini et al. / Coordination Chemistry Reviews 268 (2014) 83–100


solution; allylic alcohols very rapidly epoxidized
K2 [6] 2-Propen-1-ol H2 O2 /H2 O, 32 ◦ C, 10 h 95 [66] Efficient epoxidation in water without using phase transfer catalysts
K2 [6] 2-Hexen-1-ol H2 O2 /H2 O, 32 ◦ C, 5 h 98 [67]
(TBA)2 [25] MBO H2 O2 /CD3 CN, 32 ◦ C, 8 h 16 [69] H2 O2 efficiency only 23%
(THA)2 [25] Cyclooctene H2 O2 /CH3 CN, reflux, 4 h 100 [85] Same results obtained under N2 atmosphere
(TBA)2 [25] Cyclooctene CH3 CN, RT, 42 h (stoichiometric 2 [84] Stoichiometric epoxidation with 26 gave the corresponding epoxides and 25
oxidation)
(TBA)2 [26] Cyclooctene CH3 CN, RT, 42 h (stoichiometric 97 [84]
oxidation)
[HHIm]2 [25] Cyclooctene H2 O2 /CH3 OH:H2 O, 55 ◦ C, 4 h 54.4 [90] Catalysts can be easily recovered
[HDIm]2 [25] Cyclooctene H2 O2 /CH3 OH:H2 O, 55 ◦ C, 4 h 100 [90] and reused
[HMIm]2 [25] Cyclooctene H2 O2 /CH3 OH:H2 O, 55 ◦ C, 4 h 26.6 [90]
[DMIm]2 [25] Cyclooctene H2 O2 /CH3 OH:H2 O, 55 ◦ C, 4 h 100 [90]
(TBA)2 [SO4 {WO(O2 )2 }2 ] MBO H2 O2 /CD3 CN, 32 ◦ C, 8 h 69 [69] Catalytic activity of the dinuclear peroxo-tungstates with
(TBA)2 [HAsO4 {WO(O2 )2 }2 ] MBO H2 O2 /CD3 CN, 32 ◦ C, 8 h 12 [69] XO4 n− ligands: Se > S > As > P > Si; the low Lewis acidity of
(TBA)2 [33] MBO H2 O2 /CD3 CN, 32 ◦ C, 8 h 6 [69] W in 16 results in high catalytic activity for epoxidation
(TBA)2 [15a] MBO H2 O2 /CD3 CN, 32 ◦ C, 8 h 1 [69]
(TBA)2 [16] cyclohexene H2 O2 /CH3 CN, 50 ◦ C, 40 min 88 [93]
(TBA)2 [SO4 {WO(O2 )2 }2 ] cyclohexene H2 O2 /CH3 CN, 50 ◦ C, 40 min 18 [93]
(TBA)2 [HAsO4 {WO(O2 )2 }2 ] cyclohexene H2 O2 /CH3 CN, 50 ◦ C, 40 min 4 [93]
(TBA)2 [33] cyclohexene H2 O2 /CH3 CN, 50 ◦ C, 40 min 2 [93]
(TBA)2 [15a] cyclohexene H2 O2 /CH3 CN, 50 ◦ C, 40 min 1 [93]
17a styrene TBHP/CCl4 , 60 ◦ C, 5 h 28 [74] Small enantiomeric excesses
17b styrene TBHP/CCl4 , 60 ◦ C, 5 h 29 [74]
WCl2 (O)(O2 )dppmO2 Cyclooctene H2 O2 /EtOH, 70 ◦ C, 6 h 98 [75] Lower reactivity for W- than for Mo-based catalysts in the presence of TBHP; the
W(O)(O2 )2 dppmO2 Cyclooctene H2 O2 /EtOH, 70 ◦ C, 6 h 90 [76] former showed superior reactivity to the latter for H2 O2 as the oxidant
19 Cyclohexene H2 O2 :NaHCO3 /CH3 CN, RT, 50 min 93 [80] High yield, high TON, short reaction time
20 Cyclohexene H2 O2 :NaHCO3 /CH3 CN, RT, 50 min 72 [80]
22 Cyclohexene H2 O2 :NaHCO3 /CH3 CN, RT, 50 min 90 [80]
23 Cyclohexene H2 O2 :NaHCO3 /CH3 CN, RT, 50 min 68 [80]
27a Cyclohexene H2 O2 :NaHCO3 /CH3 CN, RT, 1 h 76 [91] Significant decrease in catalytic efficiency in
27b Cyclohexene H2 O2 :NaHCO3 /CH3 CN, RT, 1 h 80 [91] the absence of NaHCO3 ; anionic complexes
27c Cyclohexene H2 O2 :NaHCO3 /CH3 CN, RT, 1 h 82 [91] showed the highest catalytic activity
27d Cyclohexene H2 O2 :NaHCO3 /CH3 CN, RT, 1 h 78 [91]
27e Cyclohexene H2 O2 :NaHCO3 /CH3 CN, RT, 1 h 86 [91]
28 Cyclohexene H2 O2 :NaHCO3 /CH3 CN, RT, 1 h 74 [91]
29 Cyclohexene H2 O2 :NaHCO3 /CH3 CN, RT, 1 h 89 [91]
30 Cyclohexene H2 O2 :NaHCO3 /CH3 CN, RT, 1 h 96 [92] Mo and W complexes almost equally effective as epoxidation catalysts; for alcohol,
31 Cyclohexene H2 O2 :NaHCO3 /CH3 CN, RT, 1 h 99 [92] sulfide, and amine oxidation, W slightly more potent than Mo catalyst
34 Cyclohexene H2 O2 :NaHCO3 /CH2 Cl2 :CH3 OH, RT, 93 [100] Efficient catalysts in regard to epoxide yield and TON.
74.5 min
M. Amini et al. / Coordination Chemistry Reviews 268 (2014) 83–100 93

Fig. 11. Molecular structure of complex 34. Figure reproduced from reference [100] with the permission of the copyright holders.

tungsten(VI) peroxido complexes were immobilized for the first to be accounted for by the lipophilic–hydrophilic balance inside
time on functionalized gel and macroporous-type PS through pyri- the polymer, allowing good diffusion of both polar H2 O2 and non-
dine, pyridine-N-oxide, phosphine oxide (35a–e), phosphonic acid polar olefin [103]. In this case the support is not inert in catalysis
(36a–d), and phosphonamide (37a,b and 38a,b) ligands (Fig. 12) and its properties can play an important role. The characteristics
[101,102]. Cyclohexene epoxidation by H2 O2 was successful with can be tuned by appropriate modification of the support surface
all five catalyst types. Catalyst screening revealed that the best TOF to obtain an optimal hydrophilic–hydrophobic balance for efficient
was obtained with 38a. Cyclohexene epoxidation in dioxane solu- mass transfer of reagents.
tion at 70 ◦ C catalyzed by heterogeneous 38a gave cyclohexene In 2000, Gelbard briefly reviewed work on alkene epox-
oxide as the main product, as well as the corresponding diol (∼20%). idation with peroxido tungsten(VI) catalysts immobilized on
Catalyst 38a could be recycled three times without a significant organophosphoryl-functionalized polymers [104]. Gelbard et al.
decrease in catalytic activity. also reported on cyclohexene epoxidation with peroxotungstic
The same research group reported on epoxidation with PMA- species supported on PMA ammonium resins 40a–e (Fig. 13) [103].
based peroxido W(VI) catalysts (Fig. 13) [103]. PMA, which is more Compared to 39a and 39b, catalysts 40a–e exhibited poorer activity
polar than PS, changes the relative diffusion rate of the olefin and and selectivity toward the epoxy product.
H2 O2 inside the beads and the hydrophilic–hydrophobic balance In an identical approach, dimer 6 (Fig. 3) was electrostatic-
around the catalytic site. Under reaction conditions similar to those ally immobilized on amberlite IRA-900 (catalyst 41) [105,106].
used for catalysts 35–38 [102], the heterogeneous catalyst 39a
(Fig. 13) led to efficient catalysis of cyclohexene epoxidation and
TOF values >1000 were obtained. In the case of catalyst 39b, 13% diol
formation was observed owing to the acidity of the macroporous
PMA support. Under similar reaction conditions, the homogeneous
catalyst WO(O2 )2 (HMPT) 1c (Fig. 1) led to a lower TOF value of
75. The higher catalytic activity of heterogeneous catalysts seems

R
PS P
R
O O PS PO3(WO5)2 (Q+)2
O
W
O O
O
35a; PS= BioBeads, R= Ph 36a; Q=Oct3MeN
35b; PS= BioBeads, R= octyl 36b; Q=Bu4N
35c; PS= XAD4, R= octyl
35d; PS= XAD4, R= Ph
35e; PS= XAD16, R= Ph

NMe2 Me NMe2

PS P PS CH2N P
NMe2 NMe2
O O O O
O O
W W
O O O O
O O
37a; PS= Phosphonic resin LES63 38a; PS= BioBeads SX1
37b; PS= BioBeads SX1 38b; PS= Duolite A369

Fig. 12. Peroxido tungsten(VI) catalysts immobilized on organophosphoryl-


functionalized PS [102]. Fig. 13. Polymethacrylate-based peroxido W(VI) catalysts [103,104].
94 M. Amini et al. / Coordination Chemistry Reviews 268 (2014) 83–100

Fig. 14. W(VI), Mo(VI), and V(V) peroxido complexes immobilized on PS through the chelating ligands 2-pybmz and 3-pybmz [107].

Amberlite IRA-900 resin is a macroreticular PS containing quater- that 33 was immobilized on the silica surface without significant
nary ammonium groups. The catalyst was used for cyclooctene deformation, but different immobilized species were observed for
oxidation by H2 O2 and provided a moderate yield of the corre- Si-MCM-41. For the complex supported on silica beads, H2 O2 epox-
sponding epoxide (19%, 16 h). Catalyst screening studies revealed idation of cyclooctene in H2 O/t-BuOH at RT led to the formation of
that 41 did not consume the oxidant in a very efficient manner. cyclooctene oxide in 99% yield.
Peroxido complexes of W(VI), Mo(VI), and V(V) were immo- Like silica beads, the titanium silicate (TS-1) surface can anchor
bilized on PS crossed-linked with 5% divinylbenzene using the anionic oxido–peroxido W(VI) complexes. Brégeault et al. tested
chelating ligands 2-pybmz and 3-pybmz to give PS-[MO(O2 )2 (L)] the [HPO4 {WO(O2 )2 }2 ]2− anion 33 supported on TS-1 (catalyst
(M = W, Mo, and V) (Fig. 14) [107]. For styrene epoxidation by H2 O2 , 46a) in the epoxidation of (R)-(+)-limonene [109]. The substrate
the vanadium catalysts showed the best results in terms of activ- conversion obtained was rather good (75%), but the epoxide selec-
ity and selectivity toward the desired epoxide product. Use of 42a tivity was not high (53%) and substantial amounts of diol were
and 43a as catalysts for styrene oxidation in refluxing acetonitrile also produced. Overoxidation to diol was attributed to the Brønsted
resulted in poor catalytic activity (8–9.5% conversion). Selectivity to acidity of the support surface. For anion 33 grafted onto phosphate-
the epoxide was also poor, and benzaldehyde was identified as the modified TS-1 (catalyst 46b) complete conversion of substrate was
main product. When TBHP was used as the oxygen source, the for- achieved within 6 h, but only a trace of epoxide was observed. 1,2-
mation of epoxide predominated, but the styrene conversion was Diol was obtained as the major product in 53% yield.
still less than 10%. A heterogeneous W(VI) peroxido catalyst was synthesized by
The catalytic activity of the polybenzimidazole grafted W(VI) covalent anchoring of WO5 on phosphoramide-grafted silica, K10
peroxido complexes 44a and 44b (Fig. 15) was measured for cyclo- montmorillonite, and hectorite (Fig. 16) [110]. In this study, the
hexene epoxidation with H2 O2 in dioxane [103]. The epoxidation ligands and tungsten complex are covalently bonded to the sur-
activity and selectivity of the catalysts were poor. Catalyst 44a, for face. The catalysts were treated with a solution of HCl and a hot
instance, showed TOF as low as 44 h–1 and 44b led to slightly better solution of H2 O2 to remove electrostatically bound W(VI) species.
results (TOF = 74 h–1 ). The catalytic activity and selectivity of the catalysts depended on
the nature of both the support and the ligand. The catalytic activ-
3.2. Silica-anchored catalysts ity of 47a in the oxidation of cyclohexene in water/dioxane solvent
was the highest (TOF = 140 h−1 ). In a similar process, catalysts 47b
Immobilization of W(VI) peroxido complexes on silica supports and 47c had TOF of 42 h−1 and 124 h−1 , respectively. In general
is another approach to obtain novel heterogeneous catalysts for TOF for the supports varied in the order bead-shaped silica > K10
olefin epoxidation. Brégeault et al. demonstrated this approach montmorillonite > sol–gel-prepared amorphous silica > hectorite.
in the preparation of the heterogeneous olefin epoxidation cata- The neutrality of the surface and the bond strength between the
lysts via electrostatic anchoring of the [HPO4 {WO(O2 )2 }2 ]2− anion ligand and the support may contribute to this behavior. Under sim-
33 on siliceous supports (pure beads, 45a; and mesoporous Si- ilar reaction conditions, all catalysts showed poor performance in
MCM-41, 45b) [108]. Here, the anionic complex is electrostatically the oxidation of conjugated aromatic alkenes such as styrene and
fixed to Si-OH2 + groups on the silica surface. 31 P NMR revealed ␤-methylstyrene (TOF < 8 h−1 ). Recycling experiments for catalysts
recovered by hot filtration revealed significant decreases in cat-
alytic activity and selectivity. Elemental analyses for the recovered
materials revealed that the tungsten content decreased by 3–10%
after each use.
In 2004, Liu et al. elegantly synthesized a hybrid cat-
alyst by immobilizing the oxido–peroxido W(VI) compound
{HPO4 [W(O)(O2 )2 ]2 } 33 on the surface of HMS and exchanging Pd
ions into the HMS channels (catalyst 50) (Fig. 17) [111]. The hybrid
catalyst showed 34.1% propylene conversion and 83.2% selectiv-
ity for propylene oxide for oxidation of propylene using molecular
oxygen as an oxidant at 100 ◦ C. Acetone, acrolein, propionaldehyde,
propane, and C4 –C6 hydrocarbons were obtained as by-products.
Remarkably, HMS-immobilized {HPO4 [W(O)(O2 )2 ]2 } (without Pd
ions) displayed high selectivity for epoxide product but the sub-
strate conversion was only 1.7%. This highlights the importance of
Pd ions in the catalyst structure. Interestingly, oxygenated com-
pounds formed from methanol co-oxidation (such as formic acid
Fig. 15. Two polybenzimidazole-grafted W(VI) peroxido complexes [104]. and formaldehyde) were detected in small amounts after catalytic
M. Amini et al. / Coordination Chemistry Reviews 268 (2014) 83–100 95

Fig. 16. Heterogeneous W(VI) oxido–bisperoxido catalysts used by Gelbard et al. [110].

reaction, indicating that a proportion of the methanol solvent was corresponding diol was the main reaction product. Even catalyst
co-oxidized during propylene oxidation. This finding is consistent 51b, which exhibited the highest catalytic activity and selectivity,
with a mechanism in which Pd2+ reduction by methanol generates only showed 30% epoxide selectivity. The authors noted a slight
Pd0 , which then activates molecular oxygen to oxidize methanol to increase in epoxide selectivity and TOF in the second catalytic
a hydroxymethyl hydroperoxide (HOCH2 OOH) intermediate [112]. cycle. In addition, catalysts 51a, 51b, and 51e were lipophilized
The HOCH2 OOH intermediate probably plays a role in regenerating and buffered by partial exchange of the remaining H+ with Bu4 N+ .
the O O bonds of the immobilized W(VI) complex. The catalytic activity and selectivity of the buffered catalysts were
Although the homogeneous catalyst lower than for 51a–e.
(Pd(OAc)2 + {HPO4 [W(O)(O2 )2 ]2 }2− 33) showed higher con- A tungsten(VI) oxido–peroxido complex was also deposited on
version than that over the hybrid catalyst but the hybrid catalyst a HA support using a suspension of HA in a DCM solution of
could be reused five times by a simple filtration method without [AliQ]2 [W2 O11 ] (catalyst 51f) [113]. Oxidation of cyclohexene with
loss of the catalytic activity and selectivity. catalyst 51f and H2 O2 as oxidant did not reach completion (20%
In the same year, Parvulescu et al. reported successful immo- after 10 h) and the epoxide selectivity was only 4%.
bilization of WO(O2 )2 units on HA [113]. To this end, HA was Fraissard et al. designed heterogeneous catalysts 52–60 for
suspended in different concentrations of H2 [25] solution either epoxidation of cyclooctene and (R)-limonene based on function-
at room temperature or in an autoclave at 90 or 120 ◦ C (Table 2). alized silica Si(CH2 )3 Q+ [Q = NH3 , NEt3 , NC5 H5 , PPh3 ] with
The shift in 31 P CP/MAS NMR spectrum and FT-IR data for the HA- different surface lipophilicity and [HPO4 {WO(O2 )2 }2 ]2− anion 33
supported tungsten catalysts suggest that the tungsten complex (Fig. 18) [114]. Using 52 or 57 as a catalyst and H2 O2 as a terminal
is anchored via OH groups of the HA. The catalysts were tested oxidant, complete conversion of cyclooctene to cyclooctene oxide
in heterogeneous epoxidation of cyclohexene. In all cases, the was achieved in t-BuOH (TOF 35 h−1 for 52 and 8.8 h−1 for 57).
Catalysts 54 and 55 showed conversion of 50% and 66%, respec-
tively, within 24 h (TOF 2.6 h−1 for both). Other solids were less
active. Comparison of results for heterogeneous catalysts 52–60
revealed that the structure of the onium salt has only a minor

Table 2
Preparation conditions and tungsten content for catalysts 51a–e [113].

Catalyst Preparation temperature (◦ C) W content (%)

51a 90 3.54
51b Room temperature 3.65
51c 120 3.30
51d 120 3.80
51e 120 2.95
Fig. 17. Model of hybrid catalyst 50 [111].
96 M. Amini et al. / Coordination Chemistry Reviews 268 (2014) 83–100

Fig. 18. Heterogeneous W(VI) catalysts used by Fraissard et al. [114].

impact on catalyst performance, but the preparation method and accessibility of the active sites to some degree and lead to less active
lipophilicity of the support have considerable effects on catalyst catalyst.
activity and reusability. Catalysts with a hydrophilic support sur- A mesoporous SBA-15 material containing oxido–bisperoxido
face (52, 55, 57, and 59) displayed high catalytic activity in the W(VI) complexes of the type [WO(O2 )2 L] (L = pyrazolylpyridine
first run, but were rapidly deactivated. Conversely, supports with a 61a, ethylenediamine 61b, imidazole 61c, and 4,4 -bipyridine 61d)
lipophilic surface gave catalysts that were less active but relatively was synthesized via a post-grafting route (Fig. 19) [115]. All four W-
more stable in the consecutive runs. Catalysts 54, and 56 based containing solids were active catalysts for oxidation of cyclooctene
on these supports were quite stable in three consecutive runs. The using hydrogen peroxide as an oxidant, although only 61a showed
hydrophilic surface of catalyst can contribute to the poisoning by acceptable reusability. For 61b–d, a sharp decrease in catalytic
adsorption of polar epoxides and deactivated species and lead to activity was noted after the first run, probably due to a change in
faster catalyst deactivation. The end-capping SiMe3 fragments the coordination environment of the tungsten active sites during
on the catalyst surface should result in greater stability of PW2 the catalytic reaction. Comparison of the results showed that the
anions toward poisoning. Nevertheless, end-capping reduce the catalytic activity varied in the order, 61a (TOF 143 h−1 ) > 61d (TOF

Fig. 19. The heterogeneous W(VI) oxido-bisperoxido catalysts 61a–d [115].


M. Amini et al. / Coordination Chemistry Reviews 268 (2014) 83–100 97

Fig. 20. Dimer 6 immobilized on silica modified with ionic liquid [119,120].

109 h−1 ) > 61c (TOF 54 h−1 ) > 61b (TOF 17 h−1 ). The heterogeneous
catalyst 61a could be recycled six times without any loss of activity
because of the strong binding between WO(O2 )2 and the chelating
ligand [115].
For selective oxidation of olefins to epoxides via heterogeneous
W(VI) catalysis, the acidity of the support should first be con-
sidered [102,110]. If a highly acidic support is used, the epoxide
can undergo ring opening and further oxidation to give various
products. Vafaeezadeh et al. reported on catalytic oxidation of
cyclohexene to adipic acid with hydrogen peroxide over 1-butyl-
3-methylimidazolium tungstate ([BMIm]2 WO4 ) IL supported on
silica sulfamic acid, without any formation of epoxide or diol prod-
ucts [116]. In such a catalytic system, the epoxide is formed first
and then reacts with the silica-supported acid catalyst in solution
to form adipic acid [117,118].
Mizuno et al. used dimer 6 (Fig. 3) immobilized on IL-modified
SiO2 (62) (Fig. 20) as an efficient heterogeneous epoxidation cata-
lyst with hydrogen peroxide [119,120]. The catalytic activity of 62
for epoxidation is comparable to that of the homogeneous analog
[n-C12 H25 N(CH3 )3 ]2 [6] under the same conditions. This result indi-
Fig. 21. Oxido–bisperoxido W(VI) catalysts 63a and 63b [121].
cates that a homogeneous W(VI) catalyst can be immobilized with
retention of catalytic activity on IL-modified silica as a support.

3.3. Magnetically recyclable catalysts


oxidation of allylic alcohols and obtained the corresponding epox-
ides with excellent chemoselectivity. The catalyst could be recycled
Recent interest has focused on the use of magnetic nanoparticles
up to five times with just a minimum decrease in catalytic activity,
as for immobilization of homogeneous catalysts. The active cat-
after which the chitosan was destroyed, the catalyst was lost in the
alytic species, an IL-type oxido–bisperoxide tungsten(VI) complex,
reaction solution, and the recovery efficiency gradually decreased.
was immobilized via either hydrogen bonding (63a) or covalent
Table 3 summarizes results for catalytic epoxidation of alkenes
Si O linkage (63b) on silica-coated Fe3 O4 nanoparticles (Fig. 21)
by heterogeneous W(VI) oxido–peroxido catalysts. Interested read-
[121]. Immobilization of the IL via hydrogen bonding led to a sharp
ers are referred to the original articles for details on the reaction
increase in the tungsten complex loading. Both 63a and 63b were
conditions.
used as heterogeneous catalysts for epoxidation of a wide vari-
The main difference between heterogeneous and homogeneous
ety of olefins with H2 O2 as the oxygen source. For catalyst 63b,
catalysts is that every single catalytic entity in a homogeneous cata-
cyclooctene was completely converted to the cyclooctene oxide
lyst can act as a single active site. Therefore, homogeneous catalysts
within 3 h in H2 O/CH3 OH (1:10). Similarly, catalyst 63a displayed
are intrinsically more active and selective compared to heteroge-
excellent activity for alkene epoxidation and could be reused in
neous catalysts. The main advantages of heterogeneous catalysts
water/methanol up to a remarkable ten times with consistent activ-
are their simple separation and reutilization, but a major drawback
ity. Interestingly, the solvent has a significant effect on recovery of
is leaching of the catalyst during use and recycling, which lead to
the catalyst. The conversion dramatically decreased after the first
deactivation.
cycle when pure methanol or ethanol was used as the solvent. It
seems that water molecules help to stabilize hydrogen bond inter-
actions between the terminal Si OH groups of the silica layer and
silanol functionalities in the IL [122]. A catalytic system with 63a
could be used for epoxidation of several olefins and allylic alcohols
under mild conditions.
In an alternative approach, Lu et al. prepared and character-
ized a novel magnetically recyclable catalyst in which protonated
[W2 O11 ]2− was immobilized in a network of cross-linked chitosan
with a Fe3 O4 magnetic core (64, Fig. 22) [123]. This heteroge-
neous catalyst showed high conversion of 72–88% for the oxidation
of cyclic and linear alkenes when H2 O2 was used as the oxi-
dant (71–86% selectivity). The authors also used the catalyst for Fig. 22. Magnetically recyclable catalyst 64 [123].
98 M. Amini et al. / Coordination Chemistry Reviews 268 (2014) 83–100

Table 3
Summary of results for catalytic epoxidation of alkenes by heterogeneous oxido–peroxido W(VI) catalysts.

W complex Substrate Reaction conditions Epoxide Ref. Remarks


yield (%)

35a Cyclohexene H2 O2 /dioxane, 70 ◦ C, 1 h 54 [102] Best TOF was obtained with 38a
35b Cyclohexene H2 O2 /dioxane, 70 ◦ C, 1 h 56 [102]
35c Cyclohexene H2 O2 /dioxane, 70 ◦ C, 3 h 44 [102]
35d Cyclohexene H2 O2 /dioxane, 70 ◦ C, 3 h 60 [102]
35e Cyclohexene H2 O2 /dioxane, 70 ◦ C, 4 h 77.5 [102]
36a Cyclohexene H2 O2 /dioxane, 70 ◦ C, 1 h 88 [102]
36b Cyclohexene H2 O2 /dioxane, 70 ◦ C, 23 h 71 [102]
37a Cyclohexene H2 O2 /dioxane, 70 ◦ C, 3 h 55 [102]
37b Cyclohexene H2 O2 /dioxane, 70 ◦ C, 1.5 h 15 [102]
38a Cyclohexene H2 O2 /dioxane, 70 ◦ C, 0.5 h 59 [102]
38b Cyclohexene H2 O2 /dioxane, 70 ◦ C, 1 h 65 [102]
40a Cyclohexene H2 O2 /dioxane, 70 ◦ C, 50 min 38 [103] Lipophilic–hydrophilic balance inside the
40b Cyclohexene H2 O2 /dioxane, 70 ◦ C, 4 h 84 [103] polymeric beads allows good diffusion of all
40c Cyclohexene H2 O2 /dioxane, 70 ◦ C, 4 h 65 [103] reactants and product
40d Cyclohexene H2 O2 /dioxane, 70 ◦ C, 4 h 68 [103]
40e Cyclohexene H2 O2 /dioxane, 70 ◦ C, 3 h 54 [103]
41 Cyclooctene H2 O2 /CH3 CN, 50 ◦ C, 16 h 19 [105] Low-efficiency catalyst
42a Styrene H2 O2 /CH3 CN, 80 ◦ C, 6 h Trace [107] Oxidative cleavage of double bond
42a Styrene TBHP/CH3 CN, 80 ◦ C, 6 h 3 [107] and formation of benzaldehyde
43a Styrene H2 O2 /CH3 CN, 80 ◦ C, 6 h Trace [107]
43a Styrene TBHP/CH3 CN, 80 ◦ C, 6 h 2 [107]
44a Cyclohexene H2 O2 /dioxane, 70 ◦ C 27 [107]
44b Cyclohexene H2 O2 /dioxane, 70 ◦ C 42 [104] Catalyst can be reused five times without any loss of reactivity
45a Cyclooctene H2 O2 /t-BuOH, RT, 24 h 99 [108] High leaching of W complex, low catalyst stability
46a (R)-Limonene H2 O2 /acetone, RT, 24 h 40 [109] Brønsted acidity of the support surface converted the
46b (R)-Limonene H2 O2 /acetone, RT, 24 h Trace [109] epoxide to diol
47a Cyclohexene H2 O2 /dioxane, 60 ◦ C, 6 h 96 [110] Nature of the support has a significant effect on the
47b Cyclohexene H2 O2 /dioxane, 60 ◦ C, 6 h 97 [110] epoxidation reaction efficiency
47c Cyclohexene H2 O2 /dioxane, 60 ◦ C, 6 h 86 [110]
47d Cyclohexene H2 O2 /dioxane, 60 ◦ C, 6 h 25 [110]
48 Cyclohexene H2 O2 /dioxane, 60 ◦ C, 6 h 92 [110]
49 Cyclohexene H2 O2 /dioxane, 60 ◦ C, 6 h 91 [110]
50 Propylene O2 /CH3 OH, 100 ◦ C, 6 h 28 [111] Selectivity for epoxide over the solid catalyst was similar to
that over a homogeneous catalyst, but conversion was lower
over the former than over the latter
52 Cyclooctene H2 O2 /t-BuOH, 70 ◦ C, 6 h 100 [114] Catalyst efficiency and product yield depend on the
53 Cyclooctene H2 O2 /t-BuOH, 70 ◦ C, 24 h 12 [114] preparation method and lipophilicity of the support
54 Cyclooctene H2 O2 /t-BuOH, 70 ◦ C, 24 h 49 [114]
55 Cyclooctene H2 O2 /t-BuOH, 70 ◦ C, 24 h 66 [114]
56 Cyclooctene H2 O2 /t-BuOH, 70 ◦ C, 5 h 100 [114]
57 Cyclooctene H2 O2 /t-BuOH, 70 ◦ C, 24 h 60 [114]
58 Cyclooctene H2 O2 /t-BuOH, 70 ◦ C, 24 h 37 [114]
59 Cyclooctene H2 O2 /t-BuOH, 22 ◦ C, 24 h 24 [114]
54 (R)-Limonene H2 O2 /t-BuOH, 22 ◦ C, 24 h Trace [114]
56 (R)-Limonene H2 O2 /t-BuOH, 22 ◦ C, 24 h 70 [114]
60 (R)-Limonene H2 O2 /t-BuOH, 22 ◦ C, 0.3 h 1 [114]
61a Cyclooctene H2 O2 /CH3 CN, 55 ◦ C, 10 h 76 [115] Only 61a showed good recoverability and high stability
61b Cyclooctene H2 O2 /CH3 CN, 55 ◦ C, 3 h 83 [115] against leaching
61c Cyclooctene H2 O2 /CH3 CN, 55 ◦ C, 6 h 84 [115]
61d Cyclooctene H2 O2 /CH3 CN, 55 ◦ C, 10 h 65 [115]
62 Cyclooctene H2 O2 /CH3 CN, 55 ◦ C, 1 h 99 [119] Excellent epoxide yields, short reaction time, green approach
63a Cyclooctene H2 O2 /CH3 OH:H2 O, 60 ◦ C, 3 h 99 [121] Catalysts can be recycled 10 times with no loss of catalytic
63b Cyclooctene H2 O2 /CH3 OH:H2 O, 60 ◦ C, 3 h 99 [121] properties by simple magnetic decantation
64 Cyclooctene H2 O2 /CH3 CN, 40 ◦ C, 12 h 86 [123] For equimolar olefin:alcohol, the olefin was the first to be
oxidized

4. Conclusions and outlook We have described progress made in the area of W(VI)-catalyzed
epoxidation. Comparison of the activity of the various catalysts
The chemistry of Mo(VI) and W(VI) is quite similar in many considered in this review suggests the following conclusions:
cases [30]. However, some deviations in physicochemical prop-
erties (such as Pauling electronegativity, ionic radius, and Lewis
acid strength of the cations) exist between these two species (a) W(VI) oxido–bisperoxido complexes act not only as epoxidation
[124]. These differences may in turn result in different cat- catalysts but also as stoichiometric reagents in the epoxidation
alytic activity of Mo and W complexes. Mo(VI) oxido–peroxido of olefins. Oxygen transfer rates are much greater from W(VI)
complexes have been known as among the most efficient cata- bisperoxido complexes to alkenes than from the corresponding
lysts for epoxidation [29]. W(VI) oxido–peroxido complexes are Mo(VI) species [32,61].
even more efficient catalysts than the corresponding Mo(VI) (b) In the presence of W(VI) peroxido complexes, oxidation of allylic
species, and the ability of these complexes to catalyze epox- alcohols is several orders of magnitude faster than for sim-
idation has attracted major research interest during the last ple olefins. The mechanism of this epoxidation depends on the
20 years. existence (or absence) of an easily replaceable ligand on W.
M. Amini et al. / Coordination Chemistry Reviews 268 (2014) 83–100 99

(c) Monoperoxido W(VI) complexes always show lower activity [21] M. Jia, A. Seifert, W.R. Thiel, Chem. Mater. 15 (2003) 2174.
for catalytic epoxidation of olefins than bisperoxido complexes [22] V. Conte, F. Di Furia, S. Moro, J. Phys. Org. Chem. 9 (1996) 329.
[23] A.J. Burke, Coord. Chem. Rev. 252 (2008) 170.
with the same ligands, but still display good reactivity [80,91]. [24] M.L. Ramos, L.L.G. Justino, H.D. Burrows, Dalton Trans. 40 (2011) 4374.
(d) In the presence of H2 O2 as the oxidant, W(VI) peroxido catalysts [25] N. Mizuno, K. Yamaguchi, K. Kamata, Y. Nakagawa, in: T. Oyama (Ed.),
show higher reactivity toward olefins than the correspond- Activation of Hydrogen Peroxide by Polyoxometalates (Mechanism in Homo-
geneous and Heterogeneous Catalytic Epoxidation), Elsevier, Amsterdam,
ing Mo species do [61,75,92]. The presence of NaHCO3 as a 2008, pp. 155–176.
co-catalyst is often required for activation of H2 O2 by Mo(VI) [26] B.S. Lane, K. Burgess, Chem. Rev. 103 (2003) 2457.
peroxido complexes [29], whereas epoxidation with W(VI) per- [27] N. Mizuno, K. Kamata, Coord. Chem. Rev. 255 (2011) 2358.
[28] J.-M. Brégeault, M. Vennat, L. Salles, J.-Y. Piquemal, Y. Mahha, E. Briot, P.C.
oxido catalysts does not require any additive or co-catalyst.
Bakala, A. Atlamsani, R. Thouvenot, J. Mol. Catal. A: Chem. 250 (2006) 177.
Conversely, Mo-based catalysts usually show superior reactivity [29] M. Amini, M.M. Haghdoost, M. Bagherzadeh, Coord. Chem. Rev. 257 (2013)
to W-based catalysts when TBHP is the oxygen source [74,75]. 1093.
[30] H. Mimoun, in: S. Patai (Ed.), The Chemistry of Peroxides, Wiley, New York,
(e) DFT studies have revealed correlation between the epoxidation
1983, p. 463.
reactivity of W(VI) peroxido complexes and the electrophilic- [31] H. Mimoun, I.S. de Roch, L. Sajus, Bull. Soc. Chim. Fr (1969) 1481.
ity of the complex, as well as the O O  * -orbital energy level [32] G. Amato, A. Arcoria, F.P. Ballistrei, G.A. Tomaselli, J. Mol. Catal. 37 (1986)
[76–78]. 165.
[33] A. Arcoria, F.P. Ballistreri, G.A. Tomaselli, F. Di Furia, G. Modena, J. Org. Chem.
(f) The neutrality [109,110] and lipophilicity or hydrophilicity 51 (1986) 2374.
[103,113,114] of the support can have a major impact on the cat- [34] K.B. Sharpless, R.C. Michaelson, J. Am. Chem. Soc. 95 (1973) 6136.
alytic performance and reusability of supported W(VI) peroxido [35] K.B. Sharpless, T.R. Verhoven, Aldrichim. Acta 12 (1979) 63.
[36] A. Arcoria, F.P. Ballistreri, G.A. Tomaselli, F. Di Furia, G. Modena, J. Mol. Catal.
complexes. 18 (1983) 177.
[37] H. Mimoun, I.S. de Roch, L. Sajus, Tetrahedron 26 (1970) 37.
In the catalytic epoxidation of alkenes on W(VI) peroxido [38] N.J. Campbell, A.C. Dengel, C.J. Edwards, W.P. Griffith, J. Chem. Soc. Dalton
Trans. (1989) 1203.
catalysts, substrate conversion and epoxide selectivity are very sen- [39] G.B. Payne, P.H. Williams, J. Org. Chem. 24 (1959) 54.
sitive not only to the catalyst structure but also to the reaction [40] K.S. Kirshenbaum, K.B. Sharpless, J. Org. Chem. 50 (1985) 1979.
conditions. In the future, research directions in this area should [41] G.G. Allan, A.N. Neogi, J. Catal. 16 (1970) 197.
[42] M.A. Beg, I. Ahmad, J. Catal. 39 (1975) 260.
include the development of new ligands and W(VI) catalysts and [43] R. Stomberg, Acta Chem. Scand. Ser. A 39 (1985) 507.
identification of superior reaction conditions to obtain catalytic [44] R. Stromberg, J. Less Common Met. 143 (1988) 363.
systems with enhanced activity and acceptable epoxide selectivity. [45] F.W.B. Einstein, B.R. Penfold, Acta Crystallogr. 17 (1964) 1127.
[46] H.C. Stevens, A.J. Kaman, J. Am. Chem. Soc. 87 (1965) 734.
Although much effort has been devoted to studying the mecha-
[47] Z. Raciszewski, J. Am. Chem. Soc. 82 (1960) 1267.
nism for epoxidation of alkenes by Mo(VI) peroxido complexes, less [48] D. Prat, R. Lett, Tetrahedron Lett. 27 (1986) 707.
is known about the mechanism for epoxidation by related W(VI) [49] D. Prat, B. Delpech, R. Lett, Tetrahedron Lett. 27 (1986) 711.
[50] K.A. Saegebarth, J. Org. Chem. 24 (1959) 1212.
compounds. The mechanism for Mo(VI) catalysis of epoxidation is
[51] M. Mugdan, D.P. Young, J. Chem. Soc. (1949) 2988.
usually extended to W(VI). However, reasons for differences in acti- [52] G.B. Payne, C.W. Smith, J. Org. Chem. 22 (1957) 1682.
vation of TBHP and H2 O2 between Mo(VI) and W(VI) catalysts are [53] C. Venturello, E. Alneri, M. Ricci, J. Org. Chem. 48 (1983) 3831.
unexplained. We believe that these and related questions will be [54] P.U. Maheswari, P. de Hoog, R. Hage, P. Gamez, J. Reedijka, Adv. Synth. Catal.
347 (2005) 1759.
answered by theoretical studies in the near future. [55] K. Vassilev, R. Stamenova, C. Tsvetanov, React. Funct. Polym. 46 (2000)
165.
[56] K. Sato, M. Aoki, R. Noyori, Science 281 (1998) 1646.
Acknowledgements
[57] W. Zhu, H. Li, X. He, Q. Zhang, H. Shu, Y. Yan, Catal. Commun. 9 (2008) 551.
[58] G. Gelbard, F. Raison, E. Roditi-Lachter, R. Thouvenot, L. Ouahab, D. Grandjean,
M. Amini thanks the Research Council of the University of J. Mol. Catal. A: Chem. 114 (1996) 77.
Maragheh for financial support of this work. [59] M. Quenard, V. Bonmarin, G. Gelbard, Tetrahedron Lett. 28 (1987) 2237.
[60] N.M. Gresley, W.P. Griffith, B.C. Parkin, A.J.P. White, D.J. Williams, J. Chem. Soc.
Dalton Trans. (1996) 2039.
References [61] J.-Y. Piquemal, C. Bois, J.-M. Brégeault, Chem. Commun. (1997) 473.
[62] J.-Y. Piquemal, S. Halut, J.-M. Brégeault, Angew. Chem. Int. Ed. 37 (1998) 1146.
[1] T. Katsuki, Coord. Chem. Rev. 140 (1995) 189. [63] R. Noyori, M. Aokib, K. Satoc, Chem. Commun. (2003) 1977.
[2] G. Licini, V. Conte, A. Coletti, M. Mba, C. Zonta, Coord. Chem. Rev. 255 (2011) [64] C. Denis, K. Misbahi, A. Kerbal, V. Ferrières, D. Plusquellec, Chem. Commun.
2345. (2001) 2460.
[3] C. Bolm, Coord. Chem. Rev. 237 (2003) 245. [65] J. Prandi, H.B. Kagan, H. Mimoun, Tetrahedron Lett. 27 (1986) 2617.
[4] A.G.J. Ligtenbarg, R. Hage, B.L. Feringa, Coord. Chem. Rev. 237 (2003) 89. [66] K. Kamata, K. Yamaguchi, S. Hikichi, N. Mizuno, Adv. Synth. Catal. 345 (2003)
[5] N. Mizunoa, K. Yamaguchi, K. Kamata, Coord. Chem. Rev. 249 (2005) 1944. 1193.
[6] M.R. Maurya, A. Kumar, J.C. Pessoa, Coord. Chem. Rev. 255 (2011) 2315. [67] N. Mizuno, S. Hikichi, K. Yamaguchi, S. Uchida, Y. Nakagawa, K. Uehara, K.
[7] M. Amini, M.M. Haghdoost, M. Bagherzadeh, A. Ellern, L.K. Woo, Polyhedron Kamata, Catal. Today 117 (2006) 32.
61 (2013) 94. [68] K. Kamata, K. Yamaguchi, N. Mizuno, Chem. Eur. J. 10 (2004) 4728.
[8] J. Pisk, D. Agustin, V. Vrdoljak, R. Poli, Adv. Synth. Catal. 353 (2011) 2910. [69] K. Kamata, T. Hirano, S. Kuzuya, N. Mizuno, J. Am. Chem. Soc. 131 (2009) 6997.
[9] J. Morlot, N. Uyttebroeck, D. Agustin, R. Poli, ChemCatChem 5 (2013) 601. [70] H.-C. Shi, X.-Y. Wang, R. Hua, Z.-G. Zhang, J. Tang, Tetrahedron 61 (2005) 1297.
[10] B. Machura, J. Palion, J. Mroziński, B. Kalińska, M. Amini, M.M. Najafpour, R. [71] S.P. Das, J.J. Boruah, H. Chetry, N.S. Islam, Tetrahedron Lett. 53 (2012) 1163.
Kruszynski, Polyhedron 53 (2013) 132. [72] K. Kamata, T. Hirano, R. Ishimotoa, N. Mizuno, Dalton Trans. 39 (2010) 5509.
[11] S. Patai, The Chemistry of Peroxides, Wiley, Chichester, 1983. [73] P. Hazarika, D. Kalita, S. Sarmah, R. Borah, N.S. Islam, Polyhedron 25 (2006)
[12] M. Bagherzadeh, M. Amini, H. Parastar, M. Jalali-Heravi, A. Ellern, L.K. Woo, 3501.
Inorg. Chem. Commun. 20 (2012) 86. [74] S.-W. Park, K.-J. Kim, S.S. Yoon, Bull. Korean Chem. Soc. 21 (2000) 446.
[13] W. Ando, Organic Peroxides, Wiley, Chichester, 1992. [75] A. Jimtaisong, R.L. Luck, Inorg. Chem. 45 (2006) 10391.
[14] M. Bagherzadeh, M.M. Haghdoost, M. Amini, P.G. Derakhshandeh, Catal. Com- [76] D.V. Deubel, J. Sundermeyer, G. Frenking, Eur. J. Inorg. Chem. (2001) 1819.
mun. 23 (2012) 14. [77] D.V. Deubel, J. Sundermeyer, G. Frenking, J. Am. Chem. Soc. 122 (2000) 10101.
[15] M. Bagherzadeh, M. Amini, A. Ellern, L.K. Woo, Inorg. Chem. Commun. 15 [78] C. Di Valentin, P. Gisdakis, I.V. Yudanov, N. Rösch, J. Org. Chem. 65 (2000) 2996.
(2012) 52. [79] M.S. Reynolds, S.J. Morandi, J.W. Raebiger, S.P. Melican, S.P.E. Smith, Inorg.
[16] M. Bagherzadeh, M. Zare, V. Amani, A. Ellern, L.K. Woo, Polyhedron 53 (2013) Chem. 33 (1994) 4977.
223. [80] S.K. Maiti, S. Dinda, N. Gharah, R. Bhattacharyya, New J. Chem. 30 (2006) 479.
[17] S.K. Maiti, K.M.A. Malik, S. Gupta, S. Chakraborty, A.K. Ganguli, A.K. Mukherjee, [81] S.K. Maiti, S. Banerjee, A.K. Mukherjee, K.M. Abdul Malikc, R. Bhattacharyya,
R. Bhattacharyya, Inorg. Chem. 45 (2006) 9843. New J. Chem. 29 (2005) 554.
[18] R. Curci, J.O. Edwards, in: G. Strukul (Ed.), Catalytic Oxidations with Hydrogen [82] D.E. Richardson, H. Yao, K.M. Frank, D.A. Bannett, J. Am. Chem. Soc. 122 (2000)
Peroxide as Oxidant, Kluwer, Dordrecht, 1992, pp. 97–151. 1729.
[19] J.A.L. da Silva, J.J.R.F. da Silva, A.J.L. Pombeiro, Coord. Chem. Rev. 255 (2011) [83] B.S. Lane, M. Vogt, V.J. DeRose, K. Burgess, J. Am. Chem. Soc. 124 (2002) 11946.
2232. [84] K. Kamata, S. Kuzuya, K. Uehara, S. Yamaguchi, N. Mizuno, Inorg. Chem. 46
[20] M. Jia, W.R. Thiel, Chem. Commun. (2002) 2392. (2007) 3768.
100 M. Amini et al. / Coordination Chemistry Reviews 268 (2014) 83–100

[85] I.C.M.S. Santos, F.A. Almeida Paz, M.M.Q. Simões, M.G.P.M.S. Neves, J.A.S. Cav- [105] D. Hoegaerts, B.F. Sels, D.E. de Vos, F. Verpoort, P.A. Jacobs, Catal. Today 60
aleiro, J. Klinowski, A.M.V. Cavaleiro, Appl. Catal. A: Gen. 351 (2008) 166. (2000) 209.
[86] R. Ishimoto, K. Kamata, N. Mizuno, Angew. Chem. Int. Ed. 51 (2012) 4662. [106] B. Tamami, H. Yeganeh, React. Funct. Polym. 50 (2002) 101.
[87] R. Ishimoto, K. Kamata, N. Mizuno, Eur. J. Inorg. Chem. (2013) 194. [107] M.R. Maurya, M. Kumar, S. Sikarwar, React. Funct. Polym. 66 (2006)
[88] M. Bagherzadeh, S. Ghazali-Esfahani, New J. Chem. 36 (2012) 971. 808.
[89] M. Bagherzadeh, S. Ghazali-Esfahani, Tetrahedron Lett. 44 (2013) 8943. [108] J.-M. Brégeault, J.-Y. Piquemal, E. Briot, E. Duprey, F. Launay, L. Salles, M.
[90] Y. Qiao, Z. Hou, H. Li, Y. Hu, B. Feng, X. Wang, L. Hua, Q. Huang, Green Chem. Vennat, A.-P. Legrand, Microporous Mesoporous Mater. 44/45 (2001) 409.
11 (2009) 1955. [109] E. Duprey, J. Maquet, P.P. Man, J.-M. Manoli, M. Delamar, J.-M. Brégeault, Appl.
[91] S.K. Maiti, S. Dinda, S. Banerjee, A.K. Mukherjee, R. Bhattacharyya, Eur. J. Inorg. Catal. A: Gen. 128 (1995) 89.
Chem. (2008) 2038. [110] G. Gelbard, T. Gauducheau, E. Vidal, V.I. Parvulescu, A. Crosman, V.M. Pop, J.
[92] N. Gharah, S. Chakraborty, A.K. Mukherjee, R. Bhattacharyya, Inorg. Chim. Acta Mol. Catal. A: Chem. 182/183 (2002) 257.
362 (2009) 1089. [111] Y. Liu, K. Murata, M. Inaba, Green Chem. 6 (2004) 510.
[93] K. Kamata, R. Ishimoto, T. Hirano, S. Kuzuya, K. Uehara, N. Mizuno, Inorg. [112] Y.-H. Yeom, N. Ulagappan, H. Frei, J. Phys. Chem. A 106 (2002) 3350.
Chem. 49 (2010) 2471. [113] A. Crosman, G. Gelbard, G. Poncelet, V.I. Parvulescu, Appl. Catal. A: Gen. 264
[94] L. Salles, J.-Y. Piquemal, R. Thouvenot, C. Minot, J.-M. Brégeault, J. Mol. Catal. (2004) 23.
117 (1997) 375. [114] T. Kovalchuk, H. Sfihi, V. Zaitsev, J. Fraissard, J. Catal. 249 (2007) 1.
[95] P. Fantucci, S. Lolli, C. Venturello, J. Catal. 169 (1997) 228. [115] J. Tang, L. Wang, G. Liu, Y. Liu, Y. Hou, W. Zhang, M. Jia, W.R. Thiel, J. Mol. Catal.
[96] S.E. Jacobson, D.A. Muccigrosso, F. Mares, J. Org. Chem. 44 (1979) 921. A: Chem. 313 (2009) 31.
[97] P. Jin, D. Wei, Y. Wen, M. Luo, X. Wang, M. Tang, J. Mol. Struct. 992 (2011) 19. [116] M. Vafaeezadeh, M.M. Hashemi, M. Shakourian-Fard, Catal. Commun. 26
[98] L. Salles, C. Aubry, R. Thouvenot, F. Robert, C. Dorémieux-Morin, G. Chottard, (2012) 54.
H. Ledon, Y. Jeannin, J.-M. Brégeault, Inorg. Chem. 33 (1994) 871. [117] Z. Bohström, I. Rico-Lattesb, K. Holmberg, Green Chem. 12 (2010) 1861.
[99] L.P. Panicheva, G.P. Meteleva, O.V. Berlina, S.A. Panichev, Petrol. Chem. 46 [118] C.-Y. Cheng, K.-J. Lin, M.R. Prasad, S.-J. Fu, S.-Y. Chang, S.-G. Shyu, H.-S. Sheu,
(2006) 422. C.-H. Chen, C.-H. Chuang, M.-T. Lin, Green Chem. Catal. Commun. 8 (2007)
[100] M. Amini, M. Bagherzadeh, B. Eftekhari-Sis, A. Ellern, L.K. Woo, J. Coord. Chem. 1060.
66 (2013) 1897. [119] K. Yamaguchi, C. Yoshida, S. Uchida, N. Mizuno, J. Am. Chem. Soc. 127 (2005)
[101] G. Gelbard, F. Breton, M.T. Charreyre, D. Dong, Makromol. Chem. Macromol. 530.
Symp. 59 (1992) 353. [120] N. Mizuno, K. Yamaguchi, Chem. Rec. 6 (2006) 12.
[102] G. Gelbard, F. Breton, M. Benelmoudeni, M. Quenard, React. Funct. Polym. 33 [121] Y. Qiao, H. Li, L. Hua, L. Orzechowski, K. Yan, B. Feng, Z. Pan, N. Theyssen, W.
(1997) 117. Leitner, Z. Hou, ChemPlusChem 77 (2012) 1128.
[103] G. Gelbard, F. Breton, M. Quenard, D.C. Sherrington, J. Mol. Catal. A: Chem. 153 [122] P. Tundo, A. Perosa, Chem. Soc. Rev. 36 (2007) 532.
(2000) 7. [123] J. Zhu, P.C. Wang, M. Lu, New J. Chem. 36 (2012) 2587.
[104] G. Gelbard, C. R. Acad. Sci. Chem. 3 (2000) 757. [124] K. Chen, A.T. Bell, E. Iglesia, J. Phys. Chem. B 104 (2000) 1292.

Você também pode gostar