You are on page 1of 10

Research Article

pubs.acs.org/acscatalysis

Tantalum−Polyhedral Oligosilsesquioxane Complexes as Structural


Models and Functional Catalysts for Epoxidation
Pascal Guillo,†,‡,§ Michael I. Lipschutz,† Meg E. Fasulo,† and T. Don Tilley*,†,‡

Department of Chemistry, University of California, Berkeley, Berkeley, California 94720, United States

Chemical Sciences Division, Lawrence Berkeley National Laboratory, 1 Cyclotron Road, Berkeley, California 94720, United States
*
S Supporting Information

ABSTRACT: Tantalum-based supported catalysts have been shown to be very


See https://pubs.acs.org/sharingguidelines for options on how to legitimately share published articles.

selective for epoxidations with aqueous hydrogen peroxide. To gain information


relative to the active site on the surface, access to molecular complexes that
mimic the active site of the catalyst on the surface is of great interest. In this
contribution, several new Ta-polyhedral oligosilsesquioxane (POSS) complexes
Downloaded via MAHIDOL UNIV on August 23, 2018 at 15:21:25 (UTC).

with POSS ligands are used to model silica-bound Ta sites and are shown to be
active as epoxidation catalysts. Notably, a Ta-POSS complex modified by
germanoxy ligands and possessing a dinuclear Ta(μ-O)(μ-OH)Ta core is
observed to be very efficient for the epoxidation of cyclooctene using aqueous
hydrogen peroxide.
KEYWORDS: silsesquioxane, tantalum, catalysis, epoxidation, hydrogen peroxide

■ INTRODUCTION
Single-site, supported catalysts have attracted considerable
been developed for higher molecular weight olefins with
alkylhydroperoxides as the oxidant.29−31 However, although
attention in the context of probing catalytic mechanisms and aqueous hydrogen peroxide is a more attractive choice for the
optimizing efficiencies for a wide range of chemical trans- oxidant, due to its low cost and environmental appeal, silica-
formations.1−6 Despite tremendous progress in this area,7−9 supported titanium centers in mesoporous structures are
there remains a markedly incomplete understanding of the rapidly degraded under aqueous conditions.30,32,33 Among the
relationship between reaction mechanisms and structural different strategies to improve the catalyst stability, one of the
properties for the active sites of heterogeneous catalysts. One most promising is the incorporation of the catalytic center into
approach for identifying crucial structural features for supported a hydrophobic environment.32−37 Development of new
catalysts is based on the study of accurate molecular models, materials such as the mesoporous mesophase material Ti-
which may be interrogated with a wide array of spectroscopic MMM or the slow addition of H2O2 represent promising
methods to establish structure−activity relationships. For this additional approaches for overcoming the problem of catalyst
purpose, polyhedral oligosilsesquioxane (POSS) ligands10−13 degradation.38,39 The Ti-POSS homogeneous catalysts are
and their metal complexes have been extensively employed to efficient with alkylhydroperoxides as the oxidant, but only
mimic the structure and chemistry of silica-bound metal modest activity is observed in aqueous H2O2.21−25 Only two
species.14−19 Indeed, silsesquioxane compounds can be used reports describe homogeneous Ti-POSS complexes that
to mimic the different types of silanol groups found on a silica efficiently catalyze epoxidation with H2O2 (Scheme 1b).26,27
surface (isolated, vicinal, and geminal) and can provide In these examples involving a Cp*Ti center, it is proposed that
information on the role of nearest-neighbor atoms on structure the sterically demanding Cp* ligand stabilizes reactive species
and reactivity.14 Furthermore, silsesquioxanes may possess pKa through the catalytic cycle. Immobilization of Ti-POSS
values for Si−OH units that are close to those of silica.20 Thus, complexes in a hydrophobic environment such as a PDMS
the study of silsesquioxane-based homogeneous models for membrane has also been shown to be an effective way to utilize
single-site heterogeneous catalysts can contribute to a better Ti-POSS complexes for epoxidations with H2O2.40 Also,
understanding of heterogeneous catalysis at the molecular level. organo-bridged silsesquioxane titanates synthesized by a sol−
Moreover, they may function as related and efficient gel process are efficient catalysts for heterogeneous epoxidation
homogeneous catalysts.14 with hydrogen peroxide.41
Titanium-POSS (Ti-POSS) complexes, mainly based on the Research in this laboratory established the utility of site-
trisilanol 1 (Scheme 1a), have been extensively studied as isolated tantalum centers on silica as inherently more selective
molecular models for silica-supported titanium epoxidation than analogous titanium catalysts for epoxidations with aqueous
catalysts such as microporous TS-1 and Ti-MCM-41.21−27
Whereas TS-1 with its hydrophobic pore structure is useful for Received: January 4, 2017
the industrial-scale epoxidation of propene with H2O2,28 Revised: February 10, 2017
mesoporous catalysts such as Ti-MCM-41 or Ti-SBA15 have Published: February 14, 2017

© 2017 American Chemical Society 2303 DOI: 10.1021/acscatal.7b00020


ACS Catal. 2017, 7, 2303−2312
ACS Catalysis Research Article

Scheme 1. (a) Trisilanol 1, (b) the Cp’TiCl-POSS Series of Complexes (R = alkyl groups), and (c) Cp*Ta(X)-POSS Complexes

Scheme 2. Ta-SAB15 Catalyst and a Possible Molecular Model

Scheme 3. Syntheses of 6 and 7a

a
(a) Diethyl ether, 23 °C, 6 h. (b) 1, diethyl ether, 23 °C, 15 h.

hydrogen peroxide.33,42−46 This interesting difference in Cp*Ta(X)-POSS complexes (X = Cl (2), OTf (3), Me (4),
catalytic behavior points to different catalytic structures and BArf (5)) exhibit no activity for this reaction.
mechanisms and highlights the need to obtain useful structural The originally reported Ta-SBA15 epoxidation catalysts,
information and molecular models for the tantalum cata- obtained by anchoring the molecular precursor Ta(OiPr)2[OSi-
lysts.33,44 In addition to the Ta-based systems, Nb-based silica (OtBu)3]3 onto SBA15 silica, efficiently utilizes aqueous
catalysts also exhibit superior performance with hydrogen hydrogen peroxide as the oxidant.42 Importantly, the efficiency
peroxide (vs organic peroxides).47−52 In this context, it is of the H2O2 conversion in this catalysis is significantly
notable that Ta complexes of silsesquioxane ligands, i.e. Ta- enhanced by the incorporation of hydrophobic substituents
POSS complexes, have received little attention. To the best of onto the silica surface.43 Although the resulting organic−
inorganic hybrid tantalum catalysts are far superior to
our knowledge, only four reports have described Ta-POSS
analogous titanium catalysts in promoting selective epoxida-
complexes, only three of which involve Ta as part of the
tions, they are about an order of magnitude lower in catalytic
silsesquioxane framework.53−56 To gain information and to activity. Thus, an important goal in developing useful supported
improve Ta-based supported catalysts, access to Ta-POSS tantalum catalysts in selective oxidations is the realization of
complexes is important for the design and development of higher conversion rates. This might be achieved by chemical
more efficient supported catalysts. This laboratory reported the modifications of the catalytic center, and it has been established
synthesis and characterization of four Cp*Ta(X)POSS that Ti and Ta catalysts exhibit higher activities when M−O−Si
complexes (X = Cl (2), OTf (3), Me (4), and BArf (5); linkages are replaced by M−O−Ge moieties.57−61 Although the
Scheme 1c).55 Interestingly, while Notestein and co-workers precise origin of this “germanium effect” is yet to be thoroughly
reported that silica-supported Cp*Ta is an epoxidation catalyst understood, it is of interest to develop structure−activity
with aqueous hydrogen peroxide as oxidant,45 the molecular relationships for tantalum catalysts with well-defined structures.
2304 DOI: 10.1021/acscatal.7b00020
ACS Catal. 2017, 7, 2303−2312
ACS Catalysis Research Article

As structural and functional models for supported tantalum with a dative interaction.62,63 The Ta−OSi(OtBu)3 distances
catalysts, Ta-POSS derivatives should prove useful given (1.937 and 1.938 Å) are in the range of those observed for
identification of catalytically active examples. Currently, no previously reported Ta(OiPr)2[OSi(OtBu)3]3 (1.860(5) to
tantalum-based homogeneous epoxidation catalysts possessing 2.041(7) Å).42
Ta−O−Ge linkages have been described. In this contribution, An accepted mechanistic step for the activation of a
several new Ta-POSS complexes are presented along with their hydroperoxide in a metal-catalyzed epoxidation is protonolysis
activities as epoxidation catalysts. Notably, a Ta-POSS complex of an M−O−Si linkage to produce a (SiOH)M−OOR
modified by germanoxy ligands and possessing a dinuclear intermediate. Indeed, this type of process has been postulated
Ta(μ-O)(μ-OH)Ta core is observed to be very efficient for the for Ti-POSS complexes where cleavage of a Ti−O bond (in
epoxidation of cyclooctene using aqueous hydrogen peroxide. preference to Cp−Ti bonding) has been proposed.21,25,66 Thus,

■ RESULTS AND DISCUSSION


Synthesis and Characterization of Ta-POSS Com-
to stabilize the Ta−O bonds toward engagement in this type of
reversible process, a chelating bidentate bis(aryloxide) ligand
was employed. Synthesis of the bidentate ligand LH2 was
plexes. To mimic the silica-supported Ta catalysts previously adapted from the synthesis of a related compound,67 and
studied (Scheme 2),42 a Ta-POSS complex with −OSi(OtBu)3 reaction of Ta(NMe2)5 with 1 equiv of LH2 in diethyl ether
ligands was initially targeted. While Ta(OiPr)2[OSi(OtBu)3]3 afforded Ta(L)(NMe2)3 (8) in 73% yield as a pale yellow
was used as the Ta precursor in the synthesis of Ta-SBA15 powder (Scheme 4). Subsequent reaction of 8 with 1 equiv of
materials, Ta(NMe2)3[OSi(OtBu)3]2 was envisioned as a the trisilanol 1 in toluene led to formation of Ta(L)-POSS·
particularly useful starting material for incorporating a Ta[OSi- HNMe2 (9) in 60% isolated yield as a white solid.
(OtBu)3]2 group because the condensation of Ta(NMe2)3[OSi- The X-ray crystal structures of 8 and 9 (Figure 2) reveal the
(OtBu)3]2 with trisilanol 1 was expected to occur with efficient presence of a chelating binaphtholate ligand with dihedral
elimination of 3 equiv of Me2NH and formation of the fully (twist) angles of 50.3 and 52.0°, respectively (Figure 2).
condensed POSS with Ta in the POSS framework. Complex 8 possesses a distorted trigonal bipyramidal geometry,
The tris(amido) complex Ta(NMe2)3[OSi(OtBu)3]2 (6) was while in 9, the tantalum center is in an octahedral coordination
obtained in 86% yield as a white powder from the reaction environment with a dimethylamine molecule coordinated to
between Ta(NMe2)3Cl2 and 2 equiv of KOSi(OtBu)3 (Scheme tantalum. As for related Ta compounds with a binaphtholate
3). Reaction of 6 with trisilanol 1 led to formation of ligand, 8 features an oxygen atom from L in an equatorial
Ta[OSi(OtBu)3]2-POSS·HNMe2 (7) in 94% yield as a white position (with Ta−Oeq = 1.937(1) Å), while the other oxygen
powder after precipitation from a diethyl ether solution with is in an axial position with a Ta−Oax distance of 2.052(1) Å.62
acetonitrile. The X-ray crystal structure of 7 reveals an To access a molecular precursor with a Ta−O−Ge linkage,
octahedral coordination environment at Ta with a molecule we focused on introduction of −OGeiPr3 ligands, as previously
of dimethylamine (Me2NH) as a ligand (Figure 1). This reported for a Ti−Ge precursor.60 Initial attempts to achieve
this via reaction of Ta(NMe2)3Cl2 with 2 equiv of KOGeiPr3
did not provide the anticipated Ta(NMe2)3(OGeiPr3)2 complex
but instead gave a mixture of products that could not be
separated. The reaction of Ta(NMe2)5 with 5 equiv of
HOGeiPr3 did not give the expected products Ta(OGeiPr3)5
or Ta(NMe2)x(OGeiPr3)5−x but instead provided the ditanta-
lum complex 10 with μ-O and two μ-OH bridges (Scheme 5).
Monitoring the reaction by 1H NMR spectroscopy indicated
clean formation of the initial product Ta(NMe2)(OGeiPr3)4
with 1 equiv of HOGeiPr3 still present in solution. After
removal of solvent under vacuum at room temperature and
redissolution in benzene-d6, a mixture of Ta(NMe2)(OGeiPr3)4
and 10 was observed by 1H NMR spectroscopy. Subsequent
removal of solvent and heating of the resulting solid at 110 °C
under vacuum led to the exclusive and clean formation of
[(iPr3GeO)3Ta]2(μ-OH)2(μ-O) (10) as the sole product. This
transformation is associated with thermal decomposition of
HOGeiPr3 into unidentified volatile products. The synthesis of
10 was accomplished by reaction of Ta(NMe2)5 with 5 equiv of
HOGeiPr3 in pentane, followed by removal of solvent and
Figure 1. Molecular structure of 7 displaying thermal ellipsoids at the heating of the resulting solid at 110 °C under vacuum for 4 h.
50% probability level. H atoms and i-butyl groups were omitted for Compound 10 was then isolated as a white powder in 85%
clarity. yield after precipitation of a solution of 10 into acetonitrile.
Suitable crystals for X-ray crystallography were obtained by
retention of an equivalent of Me2NH as a ligand has been slow evaporation of a solution of 10 in diethyl ether at −30 °C.
observed for other Ta-NMe 2 precursors employed in The μ-oxo and μ-hydroxo ligands of 10 can be identified by
protonolysis reactions.62−65 The Ta center in 7 is in a distorted the X-ray crystallographic data (Figure 3). Thus, two distinctly
octahedral environment in which the O atom trans to the amine shorter Ta−O distances of 2.024(9) and 2.027(10) Å are
ligand is more tightly bound (the Ta−O4 distance is 1.909(6) attributed to the Ta−O−Ta group, and four longer distances
Å, whereas other Ta−O distances are 1.939(6)−1.947(6) Å). between 2.113(10) and 2.183(11) Å are assigned to the Ta−
The relatively long Ta−N distance of 2.337(6) Å is consistent OH−Ta linkages. The Ta−O−Ta bond angle is 6−8° greater
2305 DOI: 10.1021/acscatal.7b00020
ACS Catal. 2017, 7, 2303−2312
ACS Catalysis Research Article

Scheme 4. Syntheses of 8 and 9a

a
(a) Diethyl ether, 23 °C, 2 h. (b) 1, toluene, 23 °C, 15 h.

Figure 2. Molecular structures of 8 and 9 displaying thermal ellipsoids at the 50% probability level. H atoms and i-butyl groups were omitted for
clarity.

Scheme 5. Synthesis of 10 and 11a

a
(a) Pentane, 23 °C, 3 h, then vacuum, 110 °C, 4 h. (b) 1, diethyl ether, 23 °C, 15 h.

than the Ta−OH−Ta bond angles. This assignment of a Ta(μ- Complex 10 appears to represent an interesting precursor to
O)(μ-OH)2Ta core is also consistent with electroneutrality in olefin epoxidation catalysts as it contains Ta−O−Ge linkages as
the complex, and the presence of the hydroxo groups was well as a ditantalum core. The latter feature presents the
confirmed by infrared spectroscopy, which revealed the opportunity to probe possible cooperative effects between
presence of a sharp band at 3647 cm−1 for the O−H stretching metal centers in a bimetallic active site, and few detailed studies
mode. Moreover, on the basis of comparisons to another on such catalysts have been realized. Previous studies on
dinuclear Ta compound with the same core, [Cp*TaCl2]2(μ- supported dititanium catalysts suggest that catalysis by such
OH)2(μ-O), 2 bands at 564 and 644 cm−1 were attributed to species is possible.30,70,71 Additionally, it has been shown that
Ta−O−Ta vibrations for 10 (vs 577 and 629 cm−1 for the catalytically active surface-bound species from the reaction
[Cp*TaCl2]2(μ-OH)2(μ-O)).68,69 of Ti(OiPr)4 with silica is an oxo-bridged Ti−O−Ti dimer.70,71
2306 DOI: 10.1021/acscatal.7b00020
ACS Catal. 2017, 7, 2303−2312
ACS Catalysis Research Article

Figure 3. Molecular structures of 10 and 11 displaying thermal ellipsoids at the 50% probability level. H atoms, i-butyl groups on siloxide ligands,
and i-propyl groups on the germoxy ligands were omitted for clarity.

In related work, a monomeric POSS complex with two Ti Table 1. Cyclooctene Epoxidation Catalyzed by 2−5, 7, 9,
centers linked with an oxo bridge was reported.72 The latter and 11a
complex results from the condensation of CpTiCl3 with
t
Cy6Si6O7(OH)4 to give Cy6Si6O12Ti2Cp2. However, for the catalyst oxidant (h) yieldb(%)
oxidation of cyclohexene with tBuOOH and a catalyst loading TaCp*X-POSS X = Cl (2), Me (3), OTf CHP 24 0
of 2 mol %, the catalytic activity of Cy6Si6O12Ti2Cp2 was much (4), B[C6F5]− (5) TBHP 24 0
lower than that of its analogue with only one Ti center. H2O2, 30% 24 0
Reaction of 10 with 1 equiv of the trisilanol 1 in diethyl ether POSSTa[OSi(OtBu)3]2·HNMe2 (7) CHP 2 24
led to formation of 11, isolated in 95% yield by precipitation TBHP 2 30
from a diethyl ether solution with acetonitrile (Scheme 5). The H2O2, 30% 2 30
crystal structure of 11 revealed that the ditantalum unit is Ta(L)-POSS·HNMe2 (9) CHP 24 0
preserved in the product with μ-oxo and μ-hydroxy bridges, TBHP 24 0
while one of the hydroxyl bridges in 10 is replaced by a siloxide H2O2, 30% 24 0
bridge (Figure 3). The structure of 11 is very unusual for a 11 CHP 22 24
metal-POSS compound in that, to the best of our knowledge, it TBHP
is the first example of a bimetallic complex derived from 1 equiv H2O2, 30% 1 94
of trisilanol 1. As for 10, the Ta centers in 11 are in a highly a
All reactions were performed in toluene at 65 °C with [catalyst] =
distorted octahedral environment with roughly trans-O−Ta−O 0.78 mM and a catalyst:substrate:oxidant of 1:600:50. bYield
angles between 152.3(4) and 170.3(4)°. The presence of the μ- determined by gas chromatography. CHP: cumene hydroperoxide,
OH group is also confirmed by the FTIR spectrum, which TBHP: tert-butyl hydroperoxide.
possesses a characteristic band at 3577 cm−1.
Catalysis. The previously reported TaCp*-POSS complexes
2−555 and the new Ta-POSS compounds 7, 9, and 11 were POSS·HNMe2 (7), a yield of ca. 30% was observed with all
studied as catalysts for epoxidation reactions. The catalytic oxidants. Complex 11, with the germoxy ligands, presented the
efficiencies of these tantalum complexes (2 mol %) in most promising results; a yield of 94% was obtained after only 1
epoxidations of cyclooctene were determined for reactions at h with hydrogen peroxide. A lower conversion (22%) was
65 °C in toluene with various oxidants and an olefin:oxidant observed with CHP as the oxidant (Table 1). Interestingly,
ratio of 12:1 (Table 1). As mentioned in the Introduction, the supported Ta heterogeneous epoxidation catalysts have also
TaCp*X-POSS (X = Cl, Me, OTf, or Barf) complexes 2−5 are shown better results with hydrogen peroxide in comparison to
analogues of titanium-based TiCp*modifiedCl-POSS that have CHP or TBHP.43,44
been reported to be efficient epoxidation catalysts with H2O2 as To further investigate the catalytic behavior of 7 and 11,
the oxidant.26,27 However, attempts to observe catalytic activity additional conditions and substrates were utilized, and the
for 2−5 were unsuccessful, and in all cases, no conversion of results are summarized in Table 2. For 7, an excess of olefin
the cyclooctene was observed after 24 h at 65 °C in toluene. substrate is required for conversion of the alkene into the
Complex 9, with the bidentate bis(aryloxide) ligand, is also corresponding epoxide. When a 1:1 or 1:3 substrate:oxidant
inactive for the epoxidation of cyclooctene with tert-butyl ratio was used, no conversion of the cyclooctene was observed
hydroperoxide (TBHP), cumene hydroperoxide (CHP), and after 24 h at 65 °C. Moreover, even with an excess of alkene,
hydrogen peroxide (H2O2) (Table 1). For Ta[OSi(OtBu)3]2- when no further conversion of alkene was observed by GC, an
2307 DOI: 10.1021/acscatal.7b00020
ACS Catal. 2017, 7, 2303−2312
ACS Catalysis Research Article

Table 2. Olefin Epoxidation Catalyzed by 7 and 11a


cat/ yieldb (%)
T substrate/ (2nd run,
catalyst substrate (°C) H2O2 t (h) 3rd run) TONc
7 cyclooctene 65 1:600:50 2 30 15
cyclohexene 65 1:600:50 2 22 11
1-octene 65 1:600:50 24 4 2
cyclooctene 65 1:100:100 24 0 0
cyclooctene 65 1:100:300 24 0 0
11 cyclooctene 65 1:100:100 1 31 31
(21, nd)
cyclooctene 65 1:600:50 1 94 47
(79, 76)
cyclooctene 65 1:600:100 1.33 80 80
(63, 50)
cyclooctene 65 1:600:200 23 79 158
cyclooctene RT 1:50:50 24 94 43 Figure 4. Molecular structure of 12 displaying thermal ellipsoids at the
cyclooctene RT 1:100:100 7 67 67 50% probability level. H atoms, i-butyl groups, and [HNMe2]+ were
cyclooctene RT 1:600:50 7 94 46 omitted for clarity.
cyclooctene RT 1:600:100 11 80 80
cyclooctene RT 1:600:200 8 70 139 to have the same 1H NMR spectrum as the one obtained
cyclooctene RT 1:100:200 24 39 39 during the degradation reaction, confirming that 12 is the
cyclohexene RT 1:50:50 23 13 6 product of decomposition and not simply formed during the
cyclohexene RT 1:600:50 7 60 30 crystallization process. Attempts to observe epoxidation of
1-octene RT 1:50:50 23 8 4 cyclooctene with 12 as catalyst showed that there was no
1-octene RT 1:600:50 16 52 26 conversion of the substrate. For this system and in the
a
All reactions were performed in toluene with [catalyst] = 0.78 mM. conditions tested, a labile site on the Ta center seems to be
b
Yield was determined by gas chromatography. cTON: turnover necessary for binding and activation of the oxidant.
number, expressed as mol epoxide/mol catalyst. Of course, the decomposition of 7 to 12 is not possible for a
silica-supported Ta catalyst. However, during the decom-
additional aliquot of 50 equiv of oxidant led to no further position 7, half of the Ta is likely transformed into tantalum
reaction. This behavior strongly suggests that decomposition of oxide after decomplexation from the POSS core. A deactivation
the catalyst occurs during reaction. of the supported catalyst by reaction with water may hinder
To examine the behavior of 7 in the presence of a peroxide formation of the intermediate tantalum−hydroperoxide com-
reagent, this complex was allowed to react in the presence of 3 plex, as occurs with related Ti-based catalysts with aqueous
equiv of H2O2·urea in benzene-d6 at 65 °C (Scheme 6). Note hydrogen peroxide as oxidant.32,33
that this oxidant allows monitoring of the reaction by NMR For 11, with cyclooctene as substrate and a catalyst:alkene:
spectroscopy without the complications associated with a large H2O2 ratio of 1:600:50, the observed conversions were the
excess of water (as in 35% H2O2). This experiment revealed the same at 65 °C or at room temperature (94%), but the reaction
transformation of 7 into a new complex (12), identified by at 65 °C was much faster (1 h vs 7 h at room temperature).
crystallization of the product from the reaction mixture. The When a catalyst:alkene:H2O2 ratio of 1:100:100 was used, the
crystal structure of 12 shows that this complex contains two reaction was more efficient at room temperature (67%
POSS ligands bridged by an octahedral TaO6 center (Figure 4). conversion) than at 65 °C (31% conversion). With an excess
Complex 12 was independently synthesized via reaction of 2 of oxidant (catalyst:alkene:H2O2 ratio of 1:100:200), low
equiv of 1 with 1 equiv of Ta(NMe2)5 (Scheme 6) and shown conversion of cyclooctene was observed (39% at room

Scheme 6. Access to 12 by a Degradation Process of 7 or by Direct Synthesis

2308 DOI: 10.1021/acscatal.7b00020


ACS Catal. 2017, 7, 2303−2312
ACS Catalysis Research Article

temperature), suggesting a slow decomposition of the catalyst. absence of a catalyst possessing the same Ta2O2 core and no
This catalyst was efficient even at 0.5% loading, which gave a germanoxy ligands, the influence of the germanium is difficult
70% yield for the epoxidation of cyclooctene. It is also to evaluate for 11. However, as previously mentioned, only 2
interesting that at room temperature the catalyst is efficient equiv of germanol is released during the catalytic cycle,
with a cyclooctene:H2O2 ratio of 1:1 (94% conversion with a suggesting that 2 germanoxy ligands are still coordinated to the
catalyst:alkene:H2O2 ratio of 1:50:50). Notably, this is in Ta centers during the catalysis.
contrast to Ti-POSS complexes, for which an excess of the In conclusion, four new Ta-POSS complexes were
alkene was necessary when alkyl peroxides were used as synthesized, characterized, and examined as catalysts for
oxidants.25 Another interesting feature of 11 is that it is still epoxidation reactions. In the context of development of new
efficient after the first run. Indeed, addition of H2O2 after the homogeneous and supported oxidation catalysts based on Ta, 7
conversion of the substrate ceased led to a subsequent and 11 are most interesting. Indeed, 7, with one Ta center and
formation of epoxide. This operation was repeated twice. A two −OSi(OtBu)3 ligands, decomposes after a few catalytic
loss in conversion was observed after each run, indicating a cycles into inactive 12, possessing two POSS ligands. The
potential slow degradation of the catalyst. However, interest- decomposition of 7 into 12 suggests that the Ta center can be
ingly, with a catalyst:substrate:H2O2 ratio of 1:600:50, no extracted from the POSS ligand, and a similar process might be
further loss was observed during the second and third runs. All possible on the silica surface, leading to deactivation of a
of these results might indicate that 11 is not the actual catalyst supported catalyst. To the best of our knowledge, 11 is the first
but a precatalyst for the epoxidation reaction. Monitoring the Ta-POSS complex that possesses catalytic epoxidation activity
reaction by 1H NMR spectroscopy by addition of 10 equiv of and is an efficient catalyst for the epoxidation of cyclooctene
cyclooctene then 5 equiv of H2O2 (30%) to a solution of 11 in using aqueous hydrogen peroxide. Moreover, only a small loss
benzene-d6 and then heating to 65 °C revealed formation of 5 of activity is observed even after three runs of catalysis. The
equiv of cyclooctene oxide (100% yield based on H2O2) and presence of μ-oxo and μ-hydroxo bridges, and perhaps the
the release of 2 equiv of HOGeiPr3. Attempts to isolate the germanoxy ligands, are possible keys for stabilization of Ta
resulting compound, which might be the active catalyst, were during the catalysis.
unsuccessful.
The reactivity of 11 as a catalyst is strongly influenced by the
solvent. All experiments involved biphasic conditions with the
■ EXPERIMENTAL SECTION
General Considerations. All manipulations of air sensitive
catalyst and the alkene in toluene and the oxidant in water. compounds were conducted under a nitrogen atmosphere using
With THF as the solvent, in which both aqueous hydrogen standard Schlenk techniques or a nitrogen atmosphere glovebox.
peroxide and 11 are soluble, no conversion of cyclooctene was Solvents were stored in PTFE-valved flasks after drying using solvent
observed. This suggests that 11 is decomposed in the presence purification systems or by distillation under nitrogen from appropriate
of an excess of oxidant and that this decomposition is drying agents. Benzene-d6 was purchased from Cambridge Isotope
Laboratories, dried over Na/K alloy, and then degassed by several
diminished under biphasic conditions in which the catalyst is freeze−pump−thaw cycles. NMR spectra were recorded on Bruker
exposed to a low concentration of oxidant. However, as spectrometers at room temperature unless otherwise noted. Spectra
observed with the lower yield after the first run, a slow were referenced internally by solvent peaks for 1H and 13C{1H} NMR
decomposition process seems to occur. For cyclohexene and 1- and tetramethylsilane for 29Si−1H HMBC experiments. X-ray analyses
octene as substrate, lower conversions were observed even with were carried out at the UC Berkeley CHEXRAY crystallographic
an excess of the substrate. However, only the epoxide was facility. Measurements of 7, 8, 9, 10, 11, and 12 were made on an
detected, and no enone, enol, or diol products were detected by APEX CCD area detector with Mo Kα radiation (λ = 0.71069 Å)
GC. monochromated with QUAZAR multilayer mirrors. Elemental
The fact that 11 is an efficient catalyst for epoxidations while analyses were performed by the College of Chemistry Microanalytical
Laboratory at the University of California, Berkeley. Infrared spectra
7 rapidly decomposes under the same conditions provides were collected using a Thermo Nicolet 6700 FTIR spectrometer. GC
information that may be utilized for the design of more efficient analyses were performed on an HP 6890N system using a
Ta-based catalysts. First of all, Ta compound 12 is more stable phenylmethyl polysiloxane DB-5 capillary column (30.0 m × 320
toward oxidizing conditions than the monometallic compound μm × 1.00 μm), and integration was performed relative to a dodecane
7. Also, the Ta2O2 core of 11 with the μ-oxo/hydroxo bridges is internal standard. No diols were detected in the reaction mixtures by
reminiscent of a structural feature for the active site of GC analyses. HOSi(OtBu)3,78 KOSi(OtBu)3,78 Ta(NMe2)3Cl2,79
hydroxylase methane monooxygenase (MMO), which pos- HOGeiPr3,60 and 2−555 were prepared according to literature
sesses 2 Fe(III) centers with μ-hydroxo bridges in the resting methods. Incompletely condensed silsesquioxane 1 was purchased
from Hybrid Plastics Inc. and dried overnight under vacuum at 50 °C
state of the catalytic cycle and a di(μ-oxo)diiron(IV) Fe2O2
prior to use. Ta(NMe2)5 was purchased from Strem Chemicals, Inc.
“diamond core” structure responsible for oxidation of methane and freshly sublimed prior to use. Cyclohexene and 1-octene were
to methanol.73−77 However, in the case of 11, the bridging purchased from Aldrich and dried over CaH2 prior to use. tert-Butyl
oxygen atom is apparently not transferred to the substrate, as hydroperoxide (TBHP, 5.5 M in decane), cumene hydroperoxide
indicated by the absence of reaction between 11 and (CHP, 80%, technical grade), hydrogen peroxide (H2O2, 30 wt % in
cyclooctene in benzene-d6 at 65 °C for 4 h in the absence of H2O), urea hydrogen peroxide, 2-tert-butyl-4-methylphenol, and
oxidant. Moreover, 2 equiv of HOGeiPr3 is released when H2O2 cyclooctene were purchased from Aldrich and used as received. All
is added, suggesting that oxidizing species are formed on the Ta other chemicals were purchased from commercial sources and used
center(s) by exchange of two germanoxy ligands by (hydro)- without further purification.
Ta(NMe2)3[OSi(OtBu)3]2 (6). To 518 mg (1.35 mmol) of a
peroxo ligand(s). As in MMO enzymes, the presence of the oxo suspension of Ta(NMe2)3Cl2 in 5 mL of diethyl ether was added
bridge probably stabilizes the oxidizing species and prevents KOSi(OtBu)3 (816 mg, 2.7 mmol) in 5 mL of diethyl ether. The
rapid decomposition of the catalyst. As mentioned in the mixture was stirred at room temperature for 6 h. The solution turned
Introduction, the presence of germanium has been described in yellow, and a white precipitate of KCl appeared. The suspension was
the literature to enhance the catalytic activity of catalysts. In the filtered on a pad of Celite, and the pale yellow filtrate was concentrated

2309 DOI: 10.1021/acscatal.7b00020


ACS Catal. 2017, 7, 2303−2312
ACS Catalysis Research Article

to about 1 mL and stored at −30 °C for crystallization. After 24 h, 6H), 2.26 (s, 6H), 2.22−2.10 (m, 7H), 1.70 (s, 18H), 1.62 (sept, 1H,
crystals of 6 were isolated by filtration and dried under vacuum to give 6.6 Hz), 1.21 (d, 6H, 6.6 Hz), 1.18 (d, 6H, 6.6 Hz), 1.13 (d, 12H, 6.6
980 mg (86%) of the product as a white solid. This compound is not Hz), 1.11 (m, 12H), 1.06−1.02 (m, 2H), 0.93−0.84 (m, 16H), 0.55
stable at room temperature and had to be stored at −30 °C under (d, 2H, 7.2 Hz). 13C{1H} NMR (C6D6, 600 MHz, 25 °C) δ 156.30,
nitrogen atmosphere.1H NMR (C6D6, 600 MHz, 25 °C): δ 3.56 (s, 138.56, 133.43, 131.85, 129.87, 127.48, 42.01, 35.59, 26.91, 26.62,
18H), 1.46 (s, 54H). 13C{1H} NMR (C6D6, 600 MHz, 25 °C) δ 72.7, 26.43, 26.42, 26.28, 26.20, 26.13, 25.22, 25.04, 24.98, 24.90, 24.87,
47.5, 32.2. 29Si{1H} NMR (C6D6, 600 MHz, 25 °C) δ −98.10. Anal. 24.37, 23.72, 23.59, 23.48, 23.42, 21.26. 29Si{1H} NMR (C6D6, 600
Calcd for C30H72N3O8Si2Ta: C, 42.89; H, 8.64; N, 5.00. Found: C, MHz, 25 °C) δ −62.10, −65.38, −67.57, −68.73. Anal. Calcd for
42.58; H, 8.73; N, 4.73. C52H98NO14Si7Ta: C, 46.65; H, 7.38; N, 1.05. Found: C, 46.28; H,
Ta[OSi(OtBu)3]2-POSS·HNMe2 (7). To 500 mg (0.63 mmol) of 7.44; N, 0.91.
the trisilanol 1 in 5 mL of diethyl ether was added a solution of 6 (530 [(iPr3GeO)3Ta]2(μ-OH)2(μ-O) (10). To a solution of Ta(NMe2)5
mg, 0.63 mmol) in 5 mL of diethyl ether. The resulting solution was (0.500 g, 1.25 mmol) in 20 mL of pentane was added HOGeiPr3 (1.36
stirred at room temperature for 15 h. After filtration of a small g, 6.23 mmol) in 20 mL of pentane via canula transfer. After stirring at
precipitate on a short pad of Celite, the solution was evaporated under room temperature for 3 h, the resulting pale yellow solution was
reduced pressure. The residue was dissolved in 2 mL of diethyl ether, evaporated under vacuum at room temperature and then at 110 °C for
and acetonitrile was then added dropwise to the pale yellow solution 4 h. The resulting white solid was then dissolved in 2 mL of diethyl
to precipitate 7. This mixture was kept at −30 °C for 12 h to achieve ether, and acetonitrile was added dropwise to precipitate 10. After 24 h
complete precipitation. 7 was isolated by filtration and dried under at −30 °C to achieve precipitation, a white precipitate was isolated by
vacuum to give 820 mg (84%) of the product as a white powder. filtration and dried under vacuum to give 2.15 g (85%) of the product
Crystals suitable for X-ray diffraction were obtained by cooling a as a white powder. Crystals suitable for X-ray diffraction were obtained
saturated solution of 7 in pentane at −30 °C. 1H NMR (C6D6, 600 by cooling a saturated solution of 10 in diethyl ether at −30 °C. 1H
MHz, 25 °C) δ 4.37 (quin, 1H, 5.4 Hz), 2.75 (d, 6H, 5.4 Hz), 2.16 (m, NMR (C6D6, 600 MHz, 25 °C) δ 1.73 (sept, 18H, 7.8 Hz), 1.40 (d,
7H), 1.53 (s, 54H), 1.10−1.27 (m, 42H), 0.90−0.95 (m, 6H), 0.84 108H, 7.8 Hz). 13C{1H} NMR (C6D6, 600 MHz, 25 °C) δ 20.32,
(dd, 6H, 16.6 Hz, 6.6 Hz), 0.71 (dd, 2H, 15 Hz, 7.8 Hz). 13C{1H} 19.43. Anal. Calcd for C54H128Ge6O9Ta2: C, 37.72; H, 7.50. Found: C,
NMR (C6D6, 600 MHz, 25 °C) δ 73.35, 41.69, 32.60, 25.44, 25.20, 38.07; H, 7.28.
24.96, 24.93, 24.61, 24.32, 23.97, 23.90, 23.56. 29Si{1H} NMR (C6D6, [(iPr3GeO)2Ta]2(μ-OH)(μ-O)(μ-OSi)-POSS·HNMe2 (11). To 0.500
600 MHz, 25 °C) δ −98.4, −68.9, −67.8, −65.9, −64.8. Anal. Calcd g (0.63 mmol) of the trisilanol 1 in 5 mL of diethyl ether was added a
for C54H124NO20Si9Ta: C, 42.08; H, 8.11; N, 0.91. Found: C, 41.83; H, solution of 10 (1.09 g, 0.63 mmol) in 5 mL of diethyl ether. The
8.07; N, 1.15. resulting solution was stirred at room temperature for 15 h. After
LH2 (Adapted from Ref 67). 2-tert-Butyl-4-methylphenol (3.73 g, filtration of a small precipitate on a short pad of Celite, the solution
22.7 mmol) was dissolved in 23 mL of glacial acetic acid. Concentrated was evaporated under reduced pressure. The residue was dissolved in 1
sulfuric acid (4.5 mL) was added to a solution of K2Cr2O7 (2.27 g, mL of diethyl ether, and acetonitrile was then added dropwise to the
7.72 mmol) in 24 mL of water. This solution was added dropwise to pale yellow solution to precipitate 11. This mixture was kept at −30
the solution of the phenol over 30 min at 50 °C. The orange solution °C. After 12 h, a white precipitate was isolated by filtration and dried
turned green, and an orange oil appeared. The solution was stirred at under vacuum to give 1.23 g (95%) of the product as a white powder.
50 °C for 3 h. After cooling to room temperature, 100 mL of water Crystals suitable for X-ray diffraction were obtained by cooling a
was added, and the aqueous phase was extracted with 200 mL of saturated solution of 11 in diethyl ether at −30 °C. 1H NMR (C6D6,
dichloromethane. The operation was repeated two more times (until 600 MHz, 25 °C) δ 4.52 (s, 1H), 2.11−2.26 (m, 7H), 1.75−1.83 (m,
the organic phase was colorless). After evaporation of the organic 12H), 1.36−1.43 (m, 78H), 1.18−1.24 (m, 30H), 1.12 (d, 6H, 6.6
phase, the orange oil remaining was purified by column chromatog- Hz), 0.86−0.98 (m, 14H). 13C{1H} NMR (C6D6, 600 MHz, 25 °C) δ
raphy (eluant: hexane:diethyl ether = 99:1) and after concentration 26.90, 26.81, 26.76, 26.67, 26.61, 26.51, 26.30, 25.63, 25.34, 25.19,
gave 1.2 g (33%) of the product as a white foam. 1H NMR (C6D6, 600 25.02, 24.96, 24.64, 24.37, 24.13, 23.49, 20.48, 20.21, 20.15, 20.00,
MHz, 25 °C) δ 7.18 (s, 2H), 6.73 (s, 2H), 5.12 (s, 2H, OH), 2.11 (s, 19.98, 19.94, 19.63. 29Si{1H} NMR (C6D6, 600 MHz, 25 °C) δ
6H, Me), 1.52 (s, 18H, tBu). 13C{1H} NMR (C6D6, 600 MHz, 25 °C) −69.91, −67.59, −65.81, −62.14. Anal. Calcd for
δ 150.75, 130.19, 129.66, 129.10, 123.62, 35.50, 30.24, 21.17. Anal. C64H148Ge14O18Si7Ta2: C, 37.41; H, 7.26. Found: C, 37.44; H, 7.28.
Calcd for C22H30O2: C, 80.94; H, 9.26. Found: C, 80.65; H, 9.11. [Ta-(POSS)2][H2NMe2] (12). To 200 mg (0.25 mmol) of the
Ta(L)(NMe2)3 (8). To a solution of LH2 (200 mg, 0.61 mmol) in 5 trisilanol 1 in 5 mL of diethyl ether was added a solution of
mL of diethyl ether was added a solution of Ta(NMe2)5 (246 mg, 0.61 Ta(NMe2)5 (51 mg, 0.13 mmol) in 5 mL of diethyl ether. The
mmol) in 5 mL of diethyl ether via canula transfer. After stirring at resulting solution was stirred at room temperature for 12 h. After
room temperature for 2 h, the resulting yellow/orange solution was filtration of a small precipitate on a short pad of Celite, the solution
concentrated under vacuum. The light orange solid was washed with 5 was evaporated under reduced pressure. The residue was dissolved in 1
mL of cold diethyl ether and gave 284 mg (73%) of the product as a mL of diethyl ether, and acetonitrile was then added dropwise to the
pale yellow solid. Crystals of 8 suitable for X-ray analysis were pale yellow solution to precipitate 12. This mixture was kept at −30
obtained by slow diffusion of pentane into a solution of 8 in toluene at °C. After 12 h, a white precipitate was isolated by filtration and dried
−30 °C. 1H NMR (C6D6, 600 MHz, 25 °C) δ 7.25 (s, 2H), 7.11 (s, under vacuum to give 150 mg (65%) of the product as a white powder.
2H), 3.18 (s, 18H, NMe2), 2.30 (s, 6H, Me), 1.54 (s, 18H, tBu). Crystals suitable for X-ray diffraction were obtained by slow
13
C{1H} NMR (C6D6, 600 MHz, 25 °C) δ 159.07, 132.72, 131.97, evaporation of a saturated solution of 12 in benzene at room
128.85, 126.75, 45.81, 35.50, 30.94, 21.47. Anal. Calcd for temperature. 1H NMR (C6D6, 600 MHz, 25 °C) δ 8.49 (s, 2H), 2.11−
C28H46N3O2Ta: C, 52.74; H, 7.27; N, 6.59. Found: C, 52.98; H, 2.26 (m, 14H), 2.05 (s, 6H), 1.25 (d, 36H, 6.6 Hz), 1.16 (d, 36H, 6.6
7.14; N, 6.52. Hz), 1.12 (d, 12H, 6.6 Hz), 0.87 (m, 16H), 0.81 (d, 12H, 7.2 Hz).
Ta(L)-POSS·HNMe2 (9). Toluene (10 mL) was added to a solid 13
C{1H} NMR (C6D6, 600 MHz, 25 °C) δ 35.28, 26.94, 26.54, 26.35,
mixture of 8 (100 mg, 0.16 mmol) and 1 (118 mg, 0.15 mmol). After 25.13, 24.93, 24.90, 24.04, 23.63. 29Si{1H} NMR (C6D6, 600 MHz, 25
stirring at room temperature for 15 h, the resulting pale yellow °C) δ −67.87, −64.46, −65.00. Anal. Calcd for C58H134NO24Si14Ta: C,
solution was evaporated under vacuum. The resulting foam was 38.62; H, 7.48; N, 0.78. Found: C, 38.38; H, 7.38; N, 0.75.
dissolved in 1 mL of diethyl ether, and 5 mL of hexamethyldisiloxane Catalytic Alkene Epoxidation. Under a flow of nitrogen, to a
was added. After 24 h at −30 °C, a white precipitate was isolated by 0.78 mM solution of the complex in 5 mL of toluene in a 25 mL three-
filtration and dried under vacuum to give 129 mg (60%) of the necked round-bottom flask fitted with a reflux condenser and two
product as a white foam. Crystals suitable for X-ray diffraction were septums were added the alkene and dodecane (68 μL) as an internal
obtained by cooling a saturated solution of 9 in pentane at −30 °C. 1H standard. The vessel was immersed in an oil bath where it was allowed
NMR (C6D6, 600 MHz, 25 °C) δ 7.28 (s, 2H), 6.99 (s, 2H), 2.27 (s, to equilibrate for 15 min at the desired temperature. The oxidant was

2310 DOI: 10.1021/acscatal.7b00020


ACS Catal. 2017, 7, 2303−2312
ACS Catalysis Research Article

added via syringe to the rapidly stirred solution to access the targeted (14) Quadrelli, E. A.; Basset, J.-M. Coord. Chem. Rev. 2010, 254,
catalyst:alkene:oxidant ratio. The course of the reaction was monitored 707−728.
by GC by taking aliquots in the organic phase, and products were (15) Lorenz, V.; Edelmann, F. T. Adv. Organomet. Chem. 2005, 53,
assigned based on known samples analyzed under the same conditions. 101−153.


*
ASSOCIATED CONTENT
S Supporting Information
(16) Levitsky, M. M.; Bilyachenko, A. N. Coord. Chem. Rev. 2016,
306, 235−269.
(17) Abbenhuis, H. C. L. Chem. - Eur. J. 2000, 6, 25−32.
(18) Ward, A. J.; Lesic, R.; Masters, A. F.; Maschmeyer, T. Proc. R.
The Supporting Information is available free of charge on the Soc. London, Ser. A 2012, 468, 1968−1984.
ACS Publications website at DOI: 10.1021/acscatal.7b00020. (19) Ward, A. J.; Masters, A. F.; Thomas, M. In Advances in Silicon
Crystal data for compounds 7, 8, 9, 10, 11, and 12; CIF Science 3; Hartmann-Thompson, C., Ed.; Springer: Berlin, 2011.
files can also be obtained free of charge from the (20) Duchateau, R.; Cremer, U.; Harmsen, R. J.; Mohamud, S. I.;
Cambridge Crystallographic Data Centre under reference Abbenhuis, H. C. L.; van Santen, R. A.; Meetsma, A.; Thiele, S. K. H.;
numbers 1525464, 1525465, 1525466, 1525467, van Tol, M. F. H.; Kranenburg, M. Organometallics 1999, 18, 5447−
1525468, and 1525469 (PDF) 5459.


(21) Abbenhuis, H. C. L.; Krijnen, S.; vanSanten, R. A. Chem.
Commun. 1997, 331−332.
AUTHOR INFORMATION (22) Maschmeyer, T.; Klunduk, M. C.; Martin, C. M.; Shephard, D.
Corresponding Author S.; Thomas, J. M.; Johnson, B. F. G. Chem. Commun. 1997, 1847−
*E-mail: tdtilley@berkeley.edu. 1848.
(23) Thomas, J. M.; Sankar, G.; Klunduk, M. C.; Attfield, M. P.;
ORCID Maschmeyer, T.; Johnson, B. F. G.; Bell, R. G. J. Phys. Chem. B 1999,
Pascal Guillo: 0000-0002-9818-7824 103, 8809−8813.
Present Address (24) Crocker, M.; Herold, R. H. M.; Orpen, A. G. Chem. Commun.
§ 1997, 2411−2412.
P.G.: Univ Toulouse, UPS, INPT, CNRS, LCC, 205 Route
Narbonne, F-31077 Toulouse, France. Univ Toulouse, Inst (25) Crocker, M.; Herold, R. H. M.; Orpen, A. G.; Overgaag, M. T.
A. J. Chem. Soc., Dalton Trans. 1999, 3791−3804.
Univ Technol Paul Sabatier, Dept Chim, Ave Georges
(26) Ventura, M.; Mosquera, M. E. G.; Cuenca, T.; Royo, B.;
Pompidou, BP 20258, F-81104 Castres, France. Jimenez, G. Inorg. Chem. 2012, 51, 6345−6349.
Notes (27) Ventura, M.; Tabernero, V.; Cuenca, T.; Royo, B.; Jiménez, G.
The authors declare no competing financial interest. Eur. J. Inorg. Chem. 2016, 2016, 2843−2849.

■ ACKNOWLEDGMENTS
This work was primarily funded by the U.S. Department of
(28) Sheldon, R. A. J. Mol. Catal. 1980, 7, 107−126.
(29) Maschmeyer, T.; Rey, F.; Sankar, G.; Thomas, J. M. Nature
1995, 378, 159−162.
(30) Brutchey, R. L.; Mork, B. V.; Sirbuly, D. J.; Yang, P. D.; Tilley,
Energy, Office of Science, Office of Basic Energy Sciences, T. D. J. Mol. Catal. A: Chem. 2005, 238, 1−12.
Chemical Sciences, Geosciences, and Biosciences Division (31) Coles, M. P.; Lugmair, C. G.; Terry, K. W.; Tilley, T. D. Chem.
under Contract DE-AC02-05CH11231.


Mater. 2000, 12, 122−131.
(32) Brutchey, R. L.; Ruddy, D. A.; Andersen, L. K.; Tilley, T. D.
REFERENCES Langmuir 2005, 21, 9576−9583.
(1) Coperet, C.; Chabanas, M.; Saint-Arroman, R. P.; Basset, J. M. (33) Ruddy, D. A.; Brutchey, R. L.; Tilley, T. D. Top. Catal. 2008, 48,
Angew. Chem., Int. Ed. 2003, 42, 156−181. 99−106.
(2) Thomas, J. M. Proc. R. Soc. London, Ser. A 2012, 468, 1884−1903. (34) Melero, J. A.; Iglesias, J.; Arsuaga, J. M.; Sainz-Pardo, J.; de
(3) Thomas, J. M.; Raja, R. Top. Catal. 2006, 40, 3−17. Frutos, P.; Blazquez, S. J. Mater. Chem. 2007, 17, 377−385.
(4) Thomas, J. M.; Raja, R.; Lewis, D. W. Angew. Chem., Int. Ed. 2005, (35) Nur, H.; Hau, N. Y.; Misnon, I. I.; Hamdan, H.; Muhid, M. N.
44, 6456−6482. M. Mater. Lett. 2006, 60, 2274−2277.
(5) Dal Santo, V.; Liguori, F.; Pirovano, C.; Guidotti, M. Molecules (36) Cagnoli, M. V.; Casuscelli, S. G.; Alvarez, A. M.; Bengoa, J. F.;
2010, 15, 3829−3856. Gallegos, N. G.; Crivello, M. E.; Herrero, E. R.; Marchetti, S. G. Catal.
(6) Coperet, C.; Comas-Vives, A.; Conley, M. P.; Estes, D. P.; Today 2005, 107−108, 397−403.
Fedorov, A.; Mougel, V.; Nagae, H.; Nunez-Zarur, F.; Zhizhko, P. A. (37) Kapoor, M. P.; Bhaumik, A.; Inagaki, S.; Kuraoka, K.; Yazawa, T.
Chem. Rev. 2016, 116, 323−421. J. Mater. Chem. 2002, 12, 3078−3083.
(7) Lelli, M.; Gajan, D.; Lesage, A.; Caporini, M. A.; Vitzthum, V.; (38) Kholdeeva, O. A.; Mel’gunov, M. S.; Shmakov, A. N.; Trukhan,
Mieville, P.; Heroguel, F.; Rascon, F.; Roussey, A.; Thieuleux, C.; N. N.; Kriventsov, V. V.; Zaikovskii, V. I.; Malyshev, M. E.;
Boualleg, M.; Veyre, L.; Bodenhausen, G.; Coperet, C.; Emsley, L. J. Romannikov, V. N. Catal. Today 2004, 91−92, 205−209.
Am. Chem. Soc. 2011, 133, 2104−2107. (39) Fraile, J. M.; García, J. I.; Mayoral, J. A.; Vispe, E. Appl. Catal., A
(8) Rossini, A. J.; Zagdoun, A.; Lelli, M.; Gajan, D.; Rascon, F.; 2003, 245, 363−376.
Rosay, M.; Maas, W. E.; Coperet, C.; Lesage, A.; Emsley, L. Chem. Sci. (40) Aish, E. H.; Crocker, M.; Ladipo, F. T. J. Catal. 2010, 273, 66−
2012, 3, 108−115. 72.
(9) Kubacka, A.; Iglesias-Juez, A.; Martinez-Arias, A.; Di Michiel, M.; (41) Wang, Y.; Magusin, P.; Vansanten, R.; Abbenhuis, H. J. Catal.
Newton, M. A.; Fernandez-Garcia, M. ChemCatChem 2012, 4, 725− 2007, 251, 453−458.
737. (42) Brutchey, R. L.; Lugmair, C. G.; Schebaum, L. O.; Tilley, T. D. J.
(10) Feher, F. J.; Newman, D. A.; Walzer, J. F. J. Am. Chem. Soc. Catal. 2005, 229, 72−81.
1989, 111, 1741−1748. (43) Ruddy, D. A.; Tilley, T. D. Chem. Commun. 2007, 3350−3352.
(11) Feher, F. J.; Budzichowski, T. A. Polyhedron 1995, 14, 3239− (44) Ruddy, D. A.; Tilley, T. D. J. Am. Chem. Soc. 2008, 130, 11088−
3253. 11096.
(12) Cordes, D. B.; Lickiss, P. D.; Rataboul, F. Chem. Rev. 2010, 110, (45) Morlanes, N.; Notestein, J. M. J. Catal. 2010, 275, 191−201.
2081−2173. (46) Morlanes, N.; Notestein, J. M. Appl. Catal., A 2010, 387, 45−54.
(13) Chandrasekhar, V.; Boomishankar, R.; Nagendran, S. Chem. Rev. (47) Tiozzo, C.; Bisio, C.; Carniato, F.; Gallo, A.; Scott, S. L.; Psaro,
2004, 104, 5847−5910. R.; Guidotti, M. Phys. Chem. Chem. Phys. 2013, 15, 13354−13362.

2311 DOI: 10.1021/acscatal.7b00020


ACS Catal. 2017, 7, 2303−2312
ACS Catalysis Research Article

(48) Gallo, A.; Tiozzo, C.; Psaro, R.; Carniato, F.; Guidotti, M. J.
Catal. 2013, 298, 77−83.
(49) Thornburg, N. E.; Nauert, S. L.; Thompson, A. B.; Notestein, J.
M. ACS Catal. 2016, 6, 6124−6134.
(50) Thornburg, N. E.; Thompson, A. B.; Notestein, J. M. ACS Catal.
2015, 5, 5077−5088.
(51) Feliczak-Guzik, A.; Wawrzyńczak, A.; Nowak, I. Microporous
Mesoporous Mater. 2015, 202, 80−89.
(52) Ivanchikova, I. D.; Maksimchuk, N. V.; Skobelev, I. Y.; Kaichev,
V. V.; Kholdeeva, O. A. J. Catal. 2015, 332, 138−148.
(53) Fei, Z. F.; Busse, S.; Edelmann, F. T. J. Chem. Soc., Dalton Trans.
2002, 2587−2589.
(54) Chabanas, M.; Quadrelli, E. A.; Fenet, B.; Coperet, C.; Thivolle-
Cazat, J.; Basset, J. M.; Lesage, A.; Emsley, L. Angew. Chem., Int. Ed.
2001, 40, 4493−4496.
(55) Guillo, P.; Fasulo, M. E.; Lipschutz, M. I.; Tilley, T. D. Dalton
Trans. 2013, 42, 1991−1995.
(56) Ehle, S.; Lorenz, V.; Liebing, P.; Hilfert, L.; Edelmann, F. T.
Inorg. Chem. Commun. 2016, 74, 82−85.
(57) Oldroyd, R. D.; Sankar, G.; Thomas, J. M.; Ozkaya, D. J. Phys.
Chem. B 1998, 102, 1849−1855.
(58) Oldroyd, R. D.; Thomas, J. M.; Sankar, G. Chem. Commun.
1997, 2025−2026.
(59) Liu, T.; Hacarlioglu, P.; Oyama, S. T.; Luo, M.-F.; Pan, X.-R.;
Lu, J.-Q. J. Catal. 2009, 267, 202−206.
(60) Cordeiro, P. J.; Guillo, P.; Spanjers, C. S.; Chang, J. W.;
Lipschutz, M. I.; Fasulo, M. E.; Rioux, R. M.; Tilley, T. D. ACS Catal.
2013, 3, 2269−2279.
(61) Cordeiro, P. J.; Tilley, T. D. Langmuir 2011, 27, 6295−6304.
(62) Schweiger, S. W.; Tillison, D. L.; Thorn, M. G.; Fanwick, P. E.;
Rothwell, I. P. J. Chem. Soc., Dalton Trans. 2001, 2401−2408.
(63) Son, A. J. R.; Schweiger, S. W.; Thorn, M. G.; Moses, J. E.;
Fanwick, P. E.; Rothwell, I. P. Dalton Trans. 2003, 1620−1627.
(64) Thorn, M. G.; Moses, J. E.; Fanwick, P. E.; Rothwell, I. P. J.
Chem. Soc., Dalton Trans. 2000, 2659−2660.
(65) Reznichenko, A. L.; Emge, T. J.; Audoersch, S.; Klauber, E. G.;
Hultzsch, K. C.; Schmidt, B. Organometallics 2011, 30, 921−924.
(66) Bordiga, S.; Bonino, F.; Damin, A.; Lamberti, C. Phys. Chem.
Chem. Phys. 2007, 9, 4854−4878.
(67) Totland, K. M.; Boyd, T. J.; Lavoie, G. G.; Davis, W. M.;
Schrock, R. R. Macromolecules 1996, 29, 6114−6125.
(68) Jernakoff, P.; Debellefon, C. D.; Geoffroy, G. L.; Rheingold, A.
L.; Geib, S. J. Organometallics 1987, 6, 1362−1364.
(69) Kwon, D.; Curtis, M. D.; Rheingold, A. L.; Haggerty, B. S. Inorg.
Chem. 1992, 31, 3489−3490.
(70) Bouh, A. O.; Rice, G. L.; Scott, S. L. J. Am. Chem. Soc. 1999, 121,
7201−7210.
(71) Sensarma, S.; Bouh, A. O.; Scott, S. L.; Alper, H. J. Mol. Catal. A:
Chem. 2003, 203, 145−152.
(72) Giovenzana, T.; Guidotti, M.; Lucenti, E.; Biroli, A. O.; Sordelli,
L.; Sironi, A.; Ugo, R. Organometallics 2010, 29, 6687−6694.
(73) Hanson, R. S.; Hanson, T. E. Microbiol. Mol. Biol. Rev. 1996, 60,
439−471.
(74) Merkx, M.; Kopp, D. A.; Sazinsky, M. H.; Blazyk, J. L.; Muller,
J.; Lippard, S. J. Angew. Chem., Int. Ed. 2001, 40, 2782−2807.
(75) Kovaleva, E. G.; Neibergall, M. B.; Chakrabarty, S.; Lipscomb, J.
D. Acc. Chem. Res. 2007, 40, 475−483.
(76) Wallar, B. J.; Lipscomb, J. D. Chem. Rev. 1996, 96, 2625−2657.
(77) Tinberg, C. E.; Lippard, S. J. Acc. Chem. Res. 2011, 44, 280−288.
(78) Choi, Y. S.; Moschetta, E. G.; Miller, J. T.; Fasulo, M.;
McMurdo, M. J.; Rioux, R. M.; Tilley, T. D. ACS Catal. 2011, 1,
1166−1177.
(79) Chisholm, M. H.; Huffman, J. C.; Tan, L. S. Inorg. Chem. 1981,
20, 1859−1866.

2312 DOI: 10.1021/acscatal.7b00020


ACS Catal. 2017, 7, 2303−2312

You might also like