You are on page 1of 7

CHEMCATCHEM

FULL PAPERS

DOI: 10.1002/cctc.201200721

Mechanistic Insights into the Reduction of Carbon Dioxide


with Silanes over N-Heterocyclic Carbene Catalysts
Siti Nurhanna Riduan, Jackie Y. Ying,* and Yugen Zhang*[a]

A metal-free reduction of carbon dioxide to methanol with N- be exothermic and were catalyzed by NHCs. The rate-deter-
heterocyclic carbenes (NHCs) as organocatalysts was achieved, mining step for the process was the first hydrosilylation step,
which was conducted under ambient conditions with hydrosi- which produced the formoxysilane intermediate, with an acti-
lane as the reducing reagent and the hydrogen source. Experi- vation energy barrier of 19.6 kcal mol 1. This observation ex-
mental results and density functional theory calculations indi- plained the high tendency of the system to form the methox-
cated that the overall process involved a three-step cascade re- ide end product, with a hydrogen-transfer yield in excess of
action. All three steps in the cascade reaction were found to 95 %.

Introduction

The rise in CO2 concentration in the atmosphere has prompted catalytic reduction of CO2 to methanol remains a research area
considerable research into CO2 fixation and utilization. Technol- of considerable interest, as it may be a viable alternative for
ogies involving the use of CO2 as a C1 feedstock have seen the industrial synthesis of methanol from CO2. Although such
major advancement in recent years.[1] Several notable recent reactions are still in the research stage, it is important to un-
examples include the conversion of CO2 to fuels by chemical[2] derstand their underlying mechanisms, so as to bring about
or photochemical[3] means, and CO2 as a comonomer in the major advances in this field.[12]
synthesis of polycarbonates[4] and as a starting material for the Recently, Wang’s group[13] reported a detailed computational
synthesis of motifs commonly found in pharmaceuticals.[5] mechanistic study on the N-heterocyclic carbene (NHC) cata-
Since our report of a successful reduction of CO2 to methanol lyzed CO2 reduction to methanol with hydrosilane that we re-
through an organocatalyzed hydrosilylation process,[2d] ported in 2009.[2d] However, we have found that their calculat-
a number of publications have ensued. The use of frustrated ed mechanism was partially inconsistent with our experimental
Lewis pairs, first mooted by the research groups of Stephan results. Herein, we report new insights into the hydrosilylation
and Erker, has been applied to activate H2 heterolytically,[6] and of CO2 with NHCs by combining experiments with DFT
more recently, frustrated Lewis pairs were shown to reversibly calculations.
uptake and release CO2.[7] A recent report on the room-temper-
ature reduction of CO2 to methanol with a frustrated Lewis
pair of aluminum and phosphines with ammonia borane as
a hydride donor yielded up to 51 % methanol, as determined
by NMR spectroscopy.[8] The same reduction with four equiva- Results and Discussion
lents of tetramethylpiperidine–boron-based frustrated Lewis
Elucidation of the reaction mechanism
pairs in an atmosphere of H2 resulted in quantitative conver-
sion with methanol isolated in 25 % yield.[9] The same tetrame- It was earlier communicated that the reduction of CO2 to
thylpiperidine–boron-based frustrated Lewis pair system was a silyl methoxide product proceeded under ambient condi-
used in combination with triethylsilane as a reductant to yield tions, without the need for external CO2 overpressure, or ele-
methane.[10] The reduction of CO2 catalyzed by transition-metal vated temperatures.[2d] This is advantageous as there is negligi-
complexes is still widely studied, with a recent report on an ef- ble energy input necessary for the process to occur. Our pro-
ficient nickel catalyst for the reduction of CO2, with boron as posed mechanism included the presence of two reaction inter-
a sacrificial reductant.[11] These recent reports illustrate that the mediates, which were observed with mass spectrometry (MS)
and NMR spectroscopy (Scheme 1). We performed NMR spec-
[a] S. N. Riduan, Prof. J. Y. Ying, Dr. Y. Zhang troscopic studies with equivalent amounts of imidazolium car-
Institute of Bioengineering and Nanotechnology boxylate and silane. The transformation from the imidazolium
31 Biopolis Way, The Nanos, 138669 (Singapore) carboxylate to the formoxysilane intermediate could be clearly
Fax: (+ 65) 6478 9080
observed. The addition of one equivalent of diphenylsilane to
E-mail: jyying@ibn.a-star.edu.sg
ygzhang@ibn.a-star.edu.sg the reaction ensured the complete conversion of the formoxy-
Supporting information for this article is available on the WWW under silane to a bis(silylacetal) intermediate (see the Supporting In-
http://dx.doi.org/10.1002/cctc.201200721. formation, Figure S1).

 2013 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim ChemCatChem 2013, 5, 1490 – 1496 1490
CHEMCATCHEM
FULL PAPERS www.chemcatchem.org

mol 1) with a four-membered-ring transition state.[13] We pro-


pose that the NHC first activates CO2 to form Imes–CO2. This
initial step was also observed experimentally. In addition, this
step also represents the rate-limiting step in the entire three-
step cascade reaction, as its energy barrier is the highest in
comparison with the rest of the transition states in the subse-
quent steps illustrated in Figure 2 B and C. In contrast, the
second step is energetically undemanding, with a small energy
difference of 9.1 kcal mol 1 between 1·Imes and TS-2. The
Scheme 1. Overall reaction scheme for the hydrosilylation of CO2. adduct 1·Imes includes the coordination of a formoxysilane in-
termediate to the NHC catalyst, whereas TS-2 involves the ap-
proach of a Ph2SiH2 molecule to the adduct, and the subse-
DFT calculations of each step
quent hydride transfer from the hydrosilane. The third reaction
step has a small energy barrier (9 kcal mol 1) in the elementary
With the experimental results at hand, we proceeded to per- step for the activation of 2 by NHC, forming transition state
form DFT calculations to further understand and elucidate the TS-3. The formation of a more stable Si O Si bond structure is
reaction mechanism. By using diphenylsilane as the silane brought about by rearrangement of the Si O bonds in the
starting material, it was found that the reaction proceeds in transition state. The ensuing intermolecular hydride transfer
a three-step cascade process, with the energy level of each in- then forms the silyl methoxide end product 3, closing the reac-
termediate step lower than that of the previous one (Figure 1). tion cycle and releasing the NHC catalyst. The calculated struc-
The complete conversion of a diphenylsilane (Ph2SiH2) mole- tures of the identified stationary points TS-1 to TS-3 are repre-
cule to a methoxide end product (Ph2HSiOSi(OCH3)Ph2), with sented in Figure S2 in the Supporting Information and were
formoxysilane (Ph2SiH(OCHO)) and bis(silylacetal) simplified to a reaction pathway involving only one Si H
(Ph2HSiOCH2OSiHPh2) as intermediates is a highly exothermic bond.
process with a DE value of 79.3 kcal mol 1. These calculations
are in agreement with our experimentally observed results.
Rate-determining step
Similarly, the three-step cascade reaction was also calculated in
a recent report by Wang et al.[13] The energy barriers of the three-step reaction reported by
The energy diagrams for each of the steps are shown in Wang’s group are 30.8, 40.7, and 29.8 kcal mol 1, respectively
Figure 2. The first step depicted in Figure 2 A is particularly in- (Figure 2).[13] In this case, the rate-limiting step would be the
teresting, as it involves the approach of the silane to the imida- second step in the three-step cascade reaction, contrary to our
zolium carboxylate (Imes–CO2), forming a five-membered-ring proposed mechanism and experimental results. Our experi-
transition state, and the subsequent hydride transfer to form mental observations strongly support the first hydrosilylation
the formoxysilane intermediate (1). It is noted that the energy step as the rate-limiting step, that is, with the largest energy
barrier for this step is relatively large, and the transition state barrier. In this step, the hydrosilane is converted to a formoxysi-
of the hydride transfer step (TS-1, see Figure S2 in the Support- lane in a CO2 atmosphere over the NHC catalyst. The reaction
ing Information for calculated structures) lies 19.6 kcal mol 1 tends to form the silyl methoxide end product or the bis(silyla-
above Imes–CO2+Ph2SiH2. However, this value is far lower than cetal) intermediate, even in the presence of excess CO2. Simi-
that obtained in the calculation by Wang’s group (+ 41.3 kcal larly, the rate-determining step of hydride transfer from

Figure 1. Energy diagram of the CO2 hydrosilylation reaction calculated by use of DFT. The values in black were calculated by us, whereas the values in gray
were calculated by Wang et al.[13]

 2013 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim ChemCatChem 2013, 5, 1490 – 1496 1491
CHEMCATCHEM
FULL PAPERS www.chemcatchem.org

Yield of methanol

The tendency of the system towards completion of


the reaction to the silyl methoxide end product was
also observed with regards to the yield of methanol.
The yield of methanol obtained after hydrolysis
would be dependent on the amount of the silyl
methoxide end product in the reaction mixture, and
it was found that a higher yield of methanol from
the reaction was obtained if a limited CO2 feedstock
was used. Yields of up to 90 % were achieved experi-
mentally in 6 h, if an excess of CO2 was introduced
into the reaction system together with a fixed
amount of Ph2SiH2. However, higher yields of more
than 95 % were achieved if a controlled amount of
CO2 (1 equiv.) was introduced (Scheme 2). These ob-
servations support our position that the formoxysi-
lane formation is the rate-limiting step. With a con-
trolled amount of CO2, the reaction would be atom-
efficient, whereas exactly three equivalents of hydride
would be required to completely convert CO2 into
the final methoxide end product, and this was re-
flected in the detected yield of methanol, > 95 %.
With an excess of CO2 introduced with a balloon,
there would be an increased probability of the hy-
drosilane reacting in the initial stages of the reaction,
resulting in a smaller amount of hydride available for
the complete reduction to silyl methoxide. However,
the yield of methanol was not significantly reduced
(from 95 % to 90 %), suggesting that the initial forma-
tion of formoxysilane has a significant energy barrier,
leaving behind considerable amounts of silane to
Figure 2. A–C) Calculated energy diagrams and related transition states of each reaction react with the formoxysilane and bis(silylacetyl) inter-
in the three-step cascade reaction, respectively. Transition states TS-1 a, TS-2 a and TS-3 a mediates to form the silylmethoxide end product.
were cited from the paper by Wang et al.[13] The high selectivity towards the methoxysilane end

a boron reductant was also reported for the tetramethylpiperi-


dine–boron-based frustrated Lewis pair system and other sys-
tems.[3, 9] The formoxysilane intermediate would react rapidly
with excess silane in the reaction mixture and was only ob-
served in the early stages of the reaction during our GC–MS
Scheme 2. Yield of methanol under different conditions.
monitoring studies.[2d] This result was further confirmed by our
NMR spectroscopic studies (Figure 3). Isotopically enriched
13
CO2 (99 atom % 13C) with diphenylsilane and triethylsilane
were employed to further investigate the intervening process- product and the corresponding high yield of methanol were
es of the reactions. The results of 13C NMR spectroscopy indi- a consequence of the three-step cascade reaction in which the
cated that the reaction with diphenylsilane was completed formoxysilane formation step was the rate-determining step.
within hours to form silylmethoxide as the major product (Fig-
ure 3 A and B), whereas the signals related to the formoxysi-
Effect of hydrosilanes
lane and bis(silylacetal) intermediates remained minor. We also
observed that the activation energy barrier to form the for- Our experimental studies have, thus far, utilized diphenylsilane
moxysilane was even more pronounced for trisubstituted si- as a starting material. It was demonstrated that the hydrosilyla-
lanes (Figure 3 C and D). Trace amounts of formoxysilane inter- tion of CO2 with diphenylsilane occurred smoothly, and com-
mediates of triethylsilane, Et3Si(OCHO) (168 ppm),[15, 16] were plete consumption of the silane was achieved in 6 h.[2d] Differ-
only observed after 12 days in a solution of the NHC catalyst ent silanes were also screened, and it was observed that trisub-
under an atmosphere of 13CO2. stituted silanes reacted sluggishly or were not at all reactive.

 2013 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim ChemCatChem 2013, 5, 1490 – 1496 1492
CHEMCATCHEM
FULL PAPERS www.chemcatchem.org

To further elucidate the reac-


tion mechanism and investigate
the effect of using sterically en-
cumbered silanes, a mixture of
diphenylsilane and phenyldime-
thylsilane was used. A typical
CO2 hydrosilylation reaction was
set up by using phenyldimethyl-
silane as the hydride source
(Scheme 3). Intermediates 5–7
were not observed in GC–MS
analysis; only the starting mate-
rials were found. Whereas the
first step of the hydrosilylation
reaction was energetically unfav-
orable for phenyldimethylsilane,
we believed that this compound
could participate in the subse-
quent steps of the reaction. If
Figure 3. 13C NMR spectra of the NMR tube reaction of 13CO2, silane (diphenylsilane in parts A and B, triethylsilane a mixture of diphenylsilane
in parts C and D), Imes–CO2 catalyst (5 mol %) in [D7]DMF. Spectra A and B show the conversion of 13CO2 (#) to (Ph2SiH2) and phenyldimethylsi-
13
CH2(OSiR3)2 (!) to 13CH3O SiR3 (heart) after 90 min and 24 h, respectively. Spectrum C proves that no reaction
lane (PhMe2SiH; 1:1 molar ratio)
occurred after 60 min. In spectrum D a very weak peak at 168 ppm is seen, indicating that a small amount of the
triethylformoxysilane (†) intermediate was generated after 12 d. was used as the silane feedstock,
diphenylsilane was consumed in

This is illustrated in Figure 3 D, as formoxysilane intermediates


for triethylsilane Et3Si(OCHO) were only observed with NMR
spectroscopy after 12 days. However, the calculated energy
profiles for all silanes used (Table 1) indicated that the reac-

Table 1. Hydrosilylation of CO2 with various silanes over Imes–CO2 1, and


their total energy differences.[a]

Entry Silane Si OCHO DE1 [kcal mol 1] Si OMe DE2 [kcal mol 1]
time [h][b] time [h][c]
1 PhSiH3 <1 12.06 <6 69.98
2 Ph2SiH2 1.5 21.13 6 79.34
3 Ph2MeSiH –[d] 33.58 – 79.18
4 PhMe2SiH – 33.64 – 103.8
5 Et2MeSiH – 29.35 – 83.10
6 Et3SiH – 17.43 – 76.30
7 Ph3SiH – 18.83 – 79.48

[a] Reaction conditions: silane (1 mmol), CO2 balloon, DMF (2 mL), room
temperature. [b] Time for the full consumption of silanes. [c] Time for re-
action completion. [d] Silane was not fully consumed after 3 d.

Figure 4. Calculated energy diagram and transition states of the reaction


tions were thermodynamically favored, which led us to con- with dimethylphenylsilane.
clude that the low activities of trisubstituted silanes were
mainly attributable to steric effects. The energy level of TS-
1 with bulky, trisubstituted silanes was significantly higher 1 h, whereas most of the phenyldimethylsilane remained in
than that with diphenylsilane. For example, the calculated acti- the reaction mixture. However, a significant amount of phenyl-
vation energy barrier for the first-step hydrosilylation with phe- dimethylsilane formed intermediate 5, and final products 6–9
nyldimethylsilane was 45.2 kcal mol 1 (Figure 4 A), which was were observed in the GC–MS analysis of the reaction mixture.
more than twice that of Ph2SiH2 (19.6 kcal mol 1; see Fig- Similar observations were made if Ph2SiH2 was added first and
ure 2 A). left to react for 30 min before the addition of PhMe2SiH (see

 2013 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim ChemCatChem 2013, 5, 1490 – 1496 1493
CHEMCATCHEM
FULL PAPERS www.chemcatchem.org

has also been proposed as the initial step for cyanosi-


lylation and siloxane polymerization reactions cata-
lyzed by NHCs.[18]
However, in our experiments, we did not observe
any interactions between the NHC and the silanes.
There were no changes in the chemical shift values
of the carbene center in the NMR spectra, even after
a solution of equimolar amounts of the Imes carbene
and diphenylsilane was allowed to stand for 2 weeks.
Interactions between silanes and the carbene center
were also not detected with a less sterically hindered
carbene, such as 1-butyl-3-methylimidazolylidene.
Such observations were also critically discussed by
Fuchter.[19]

Formaldehyde as an intermediate
Scheme 3. Products from the reactions with a mixed silane feedstock of Ph2SiH2 and
PhMe2SiH. It was proposed that the bis(silylacetal) intermediate
might undergo direct dissociation into formaldehyde
and disiloxanes as the third elementary step of the
Figures S4 and S5 in the Supporting Information). If PhMe2SiH
reaction.[13] The dissociation was mentioned to be akin to the
was added first, followed by the addition of Ph2SiH2 after
Brook rearrangement, in which an organosilyl group switches
30 min, only products 7, 8, and 9 were observed (i.e., 5 and 6
position with a hydroxyl proton to form a silyl ether in the
were not found). These experimental results further confirmed
presence of a base.[20] The resulting formaldehyde was then re-
the calculated reaction mechanism, in which the rate-deter-
ported to be able to undergo hydrosilylation with an NHC, cat-
mining step was the formation of the formoxysilane. Hydrosily-
alyzing the addition of the silane to the aldehyde. This pro-
lation of CO2 to formoxysilane with NHCs did not proceed with
posed mechanism trigged our interest in tracking the formal-
bulky trisubstituted silanes, as exemplified by the reaction with
dehyde intermediate during the reaction. However, in our NMR
phenyldimethylsilane (DE = + 45.2 kcal mol 1). However, after
monitoring experiments, no peaks associated with formalde-
formoxysilane 1 was formed in the reaction system with less
hyde were observed in either 1H or 13C NMR spectra. It is also
sterically hindered silanes, such as diphenylsilane, it could
known that the Si O bond is very strong and is unlikely to
react with the bulky phenyldimethylsilane in the second and
break under our reported reaction conditions, without the use
third step of the hydrosilylation reaction to yield silylacetal (5)
of heat or base.[2] Control experiments were performed by
and silylmethoxide (6) products. This reaction pathway was cal-
using both paraformaldehyde and formaldehyde (generated
culated to be favorable, and the energy difference between
from heating a solution of paraformaldehyde) in place of CO2,
the approach of phenyldimethylsilane to 1·Imes and the subse-
and no intermediates were observed;[21] diphenylsilane starting
quent hydride transfer (TS-2’) was minimal, yielding the silyla-
materials were recovered after 24 h. These results indicate that
cetal 5 (Figure 4 B). The exchange of silane moieties after for-
this proposed pathway,[13] which includes the hydrosilylation of
mation of the disiloxane was also considered in elucidating
formaldehyde, was not experimentally feasible.
our reaction pathway, and silane substitutions with a disiloxane
were not observed in our control reactions of mixing
(Ph2SiH)2O and PhMe2SiH. These experimental observations,
Conclusions
supplemented by DFT calculations, further reinforced our pro-
posed three-step cascade reaction mechanism. The first N-heterocyclic carbene (NHC) catalyzed CO2 reduction
reaction with hydrosilanes under ambient conditions was ex-
amined in detail. By combining DFT calculations with experi-
Formation of an NHC–Si adduct
mental investigations, several conclusions could be made. The
The formation of the NHC–CO2 adduct has been reported and proposed reaction pathway consists of a three-step cascade re-
widely accepted as an intermediate in many reactions.[14, 17] action. All three steps involve exothermic processes and are
However, there are still arguments for a favorable interaction catalyzed by the NHC catalyst. The primary activation model
between NHC and the silane, instead of CO2, as the primary ac- involves the reaction between the NHC and CO2, forming the
tivation mode in the first elementary step of the NHC-catalyzed NHC–CO2 adduct. The first step, which involves the hydrosilyla-
CO2 hydrosilylation reaction.[13, 18, 19] Although thermodynamical- tion of CO2 to form a formoxysilane intermediate, has the high-
ly unfavorable, the p–p stacking interactions between the NHC est activation energy barrier (DE = 19.6 kcal mol 1), and is es-
and the silane could activate the silane by stretching the Si H tablished as the rate-determining step. As a result, the system
bonds, elongating them, and pushing the electron density to- is highly selective towards a methoxide end product, with in
wards the hydrogen atoms. The formation of an NHC Si bond excess of a 95 % hydrogen transfer or carbon yield. It was also

 2013 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim ChemCatChem 2013, 5, 1490 – 1496 1494
CHEMCATCHEM
FULL PAPERS www.chemcatchem.org

shown that formaldehyde is unlikely to be an intermediate in er 24 h before an aliquot of isopropanol was added as an internal
this reaction. standard. An aliquot of 1 mL was removed from the sample and di-
luted with dichloromethane before the resulting mixture was sub-
jected to GC analysis with a Quadrex BTR-CW column (Carbowax
Experimental Section polyethylene glycol bonded phase).

General information
All solvents and chemicals were used as received from commercial NMR tube reaction and analysis
suppliers, unless otherwise noted. Dry solvents and an argon-gas-
1,3-Bis(2,4,6-trimethylphenyl)imidazolium carboxylate (Imes–CO2)
equipped glove box (Innovative Technologies, Inc.) were used for
was synthesized according the literature method,[14] and a stock so-
the reaction setup. Various imidazolium salts and silanes were pur-
lution of Imes–CO2 (0.05 mmol mL 1) was prepared in [D7]DMF. An
chased from Sigma–Aldrich. Imes–CO2 was synthesized according
aliquot corresponding to 0.01 equivalents of the catalyst was trans-
to the protocol detailed in the literature.[9] CO2, O2, and purified air
ferred into an NMR tube, and [D7]DMF (0.5 mL) was added. Silane
were supplied by SOXAL, and 13C-enriched CO2 was purchased
(0.1 equiv.) was subsequently added; the tube was sealed, and
from Sigma–Aldrich. GC–MS analyses were performed on a Shimad-
then evacuated and refilled with 13CO2 with two freeze–pump–
zu GCMS QP2010 system, and GC analyses were conducted on an
thaw cycles. The reaction was monitored by 13C decoupled and
Agilent GC6890N system. 1H and 13C NMR spectra were recorded
coupled NMR spectroscopy.
on a Bruker AV-400 instrument (400 MHz). Elemental analyses were
performed at the Chemical, Molecular, and Materials Analysis For reactions performed at stoichiometric ratios, one equivalent of
Centre, National University of Singapore. PA–FTIR spectra were re- silane was added to a solution of imidazolium carboxylate in an
corded on a Digilab FTS 7000 FTIR spectrometer equipped with an NMR tube with a resealable screw-top cap. Additional aliquots of
MTEC-300 photoacoustic detector. the silane were subsequently added in a glove box.

DFT calculations Reaction with a controlled amount of CO2


The DFT calculations were performed with the Gaussian 03 soft-
The reaction mixture was prepared according to the above proce-
ware.[22] The exchange–correlation function employed was B3LYP,[23]
dure, in which CO2 (15 mL) was introduced to the reaction mixture
which included a fraction of Hartree–Fock exchange to reduce the
with a gas-tight syringe. The reaction mixture was allowed to stir
self-interaction error. In this study, the 6-31(G) basis sets were used.
for 10 min before diphenylsilane was introduced. After all the CO2
After fully optimizing the structure of each compound, the total
was consumed, an argon-filled balloon was introduced to ensure
energy was obtained. For each reaction, the stationary structures
that the system was closed. Aliquots of the reaction mixture were
were obtained by optimizing the complex structures along the in-
withdrawn at hourly intervals and subjected to GC–MS analysis.
trinsic reaction coordinate of the reaction pathway. The calcula-
tions were performed with the reactants in a gaseous phase at RT
and ambient pressure, without any interactions with solvent
molecules. Acknowledgements

This work was supported by the Institute of Bioengineering and


Hydrosilylation of CO2 Nanotechnology (Biomedical Research Council, Agency for Sci-
The imidazolium salt (0.25 mmol) and sodium hydride (0.25 mmol) ence, Technology and Research, Singapore).
were dissolved in DMF (0.5 mL) in a crimp-top vial and stirred for
a minimum of 30 min for the carbene to be generated Keywords: carbenes · carbon dioxide fixation · density
(0.5 mmol mL 1). The solution was then centrifuged so that the in-
functional calculations · hydrosilylation · reduction
organic salts resulting from deprotonation would settle at the
bottom of the vial. An aliquot of 0.2 mL of the carbene solution
was transferred into a fresh vial, and the solvent (1.8 mL) was intro- [1] a) K. Huang, C.-L. Sun, Z.-J. Shi, Chem. Soc. Rev. 2011, 40, 2435 – 2452;
duced. The vial was sealed, and CO2 was introduced into the vial b) G. A. Olah, G. K. S. Prakash, A. Goeppert, J. Am. Chem. Soc. 2011, 133,
12881 – 12898; c) M. Cokoja, C. Bruckmeier, B. Reiger, W. A. Hermann,
through a balloon. The reaction was allowed to stir for 10 min,
F. E. Kuhn, Angew. Chem. 2011, 123, 8662 – 8690; Angew. Chem. Int. Ed.
after which the silane (1 mmol) was introduced. An internal stan-
2011, 50, 8510 – 8537; d) S. N. Riduan, Y. G. Zhang, Dalton Trans. 2010,
dard of mesitylene was added (0.5 mmol). Aliquots of the reaction 39, 3347 – 3357; e) T. Sakakura, J.-C. Choi, H. Yasuda, Chem. Rev. 2007,
mixture were withdrawn after specified reaction times and diluted 107, 2365 – 2387; f) Y. G. Zhang, S. N. Riduan, Angew. Chem. 2011, 123,
with methylene chloride before the GC–MS analysis. 6334 – 6336; Angew. Chem. Int. Ed. 2011, 50, 6210 – 6212.
[2] a) T. Matsuo, H. Kawaguchi, J. Am. Chem. Soc. 2006, 128, 12362 – 12363;
For conversion studies, a GC calibration curve was constructed
b) C. A. Huff, M. S. Sanford, J. Am. Chem. Soc. 2011, 133, 18122 – 18125;
with mesitylene and various concentrations of diphenylsilane. Ali- c) E. Balaraman, C. Gunanathan, J. Zhang, L. J. W. Shimon, D. Milstein,
quots were drawn from the reaction mixture at hourly intervals, Nat. Chem. 2011, 3, 609 – 614; d) S. N. Riduan, Y. G. Zhang, J. Y. Ying,
and diluted with methylene chloride before the GC analysis. Angew. Chem. 2009, 121, 3372 – 3375; Angew. Chem. Int. Ed. 2009, 48,
3322 – 3325; e) L. Gu, Y. G. Zhang, J. Am. Chem. Soc. 2010, 132, 914 –
915.
Hydrolysis reaction to release methanol [3] a) E. Barton Cole, P. S. Lakkaraju, D. M. Rampulla, A. J. Morris, E. Abelev,
A. B. Bocarsly, J. Am. Chem. Soc. 2010, 132, 11539 – 11551; b) A. J. Morris,
To produce methanol through hydrolysis of the reaction mixture, R. T. McGibbon, A. B. Bocarsly, ChemSusChem 2011, 4, 191 – 196.
the reaction was quenched after 18 h by adding 2 equivalents of [4] For reviews, see a) D. J. Darensbourg, Chem. Rev. 2007, 107, 2388 – 2410;
an NaOH/H2O solution. The reaction mixture was stirred for anoth- b) G. W. Coates, D. R. Moore, Angew. Chem. 2004, 116, 6784 – 6806;

 2013 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim ChemCatChem 2013, 5, 1490 – 1496 1495
CHEMCATCHEM
FULL PAPERS www.chemcatchem.org

Angew. Chem. Int. Ed. 2004, 43, 6618 – 6639; c) D. J. Darensbourg, Inorg. ers, P. Thuery, M. Ephritikhine, T. Cantat, Angew. Chem. 2012, 124, 191 –
Chem. 2010, 49, 10765 – 10780. 194; Angew. Chem. Int. Ed. 2012, 51, 187 – 190.
[5] a) J. Takaya, S. Tadami, K. Ukai, N. Iwasawa, Org. Lett. 2008, 10, 2697 – [13] F. Huang, G. Lu, L. Zhao, H. Li, Z.-X. Wang, J. Am. Chem. Soc. 2010, 132,
2700; b) T. Ohishi, M. Nishiura, Z. Hou, Angew. Chem. 2008, 120, 5876 – 12388 – 12396.
5879; Angew. Chem. Int. Ed. 2008, 47, 5792 – 5795; c) K. Kobayashi, Y. [14] H. A. Duong, T. N. Tekavec, A. M. Arif, J. Louie, Chem. Commun. 2004,
Kondo, Org. Lett. 2009, 11, 2035 – 2037; d) J. Takaya, K. Sasano, N. Iwasa- 112 – 113.
wa, Org. Lett. 2011, 13, 1698 – 1701; e) T. Fujihara, T. Xu, K. Semba, J. [15] A. Jansen, H. Gçrls, S. Pitter, Organometallics 2000, 19, 135 – 138.
Terao, Y. Tsuji, Angew. Chem. 2011, 123, 543 – 547; Angew. Chem. Int. Ed. [16] A. Jansen, S. Pitter, J. Mol. Catal. A: Chem. 2004, 217, 41 – 45.
2011, 50, 523 – 527; f) S. Li, W. Xuan, S. Ma, Angew. Chem. 2011, 123, [17] a) H. Zhou, W. Z. Zhang, C. H. Liu, J. P. Qu, X. B. Lu, J. Org. Chem. 2008,
2626 – 2630; Angew. Chem. Int. Ed. 2011, 50, 2578 – 2582; g) D. Yu, Y. G. 73, 8039 – 8044; b) Y. Kayaki, M. Yamamoto, T. Ikariya, Angew. Chem.
Zhang, Proc. Natl. Acad. Sci. USA 2010, 107, 20184 – 20189; h) I. I. F. Boo- 2009, 121, 4258 – 4261; Angew. Chem. Int. Ed. 2009, 48, 4194 – 4197; c) H.
gaerts, G. C. Fortman, M. R. L. Furst, C. S. J. Cazin, S. P. Nolan, Angew. Zhou, Y. M. Wang, W. Z. Zhang, J. P. Qu, X. B. Lu, Green Chem. 2011, 13,
Chem. 2010, 122, 8856 – 8859; Angew. Chem. Int. Ed. 2010, 49, 8674 – 644 – 650; d) Y. G. Zhang, J. Y. G. Chan, Energy Environ. Sci. 2010, 3, 408 –
8677; i) L. Zhang, J. Cheng, T. Oishi, Z. Hou, Angew. Chem. 2010, 122, 417.
8852 – 8855; Angew. Chem. Int. Ed. 2010, 49, 8670 – 8673; j) H. Mizuno, J. [18] a) J. J. Song, F. Gallow, J. T. Reeves, Z. Tan, N. K. Yee, C. H. Senanayake, J.
Takaya, N. Iwasawa, J. Am. Chem. Soc. 2011, 133, 1251 – 1253; k) C. Das Org. Chem. 2006, 71, 1273 – 1276; b) J. Raynaud, A. Ciolino, A. Baceiredo,
Neves Gomes, O. Jacquet, C. Villiers, P. Thuery, M. Ephritikhine, T. Cantat, M. Destarac, F. Bonette, T. Kato, Y. Gnanou, D. Taton, Angew. Chem.
Angew. Chem. 2012, 124, 191 – 194; Angew. Chem. Int. Ed. 2012, 51, 2008, 120, 5470 – 5473; Angew. Chem. Int. Ed. 2008, 47, 5390 – 5393; c) J.
187 – 190; l) D. Yu, Y. G. Zhang, Green Chem. 2011, 13, 1275 – 1279; m) D. Raynaud, Y. Gnanou, D. Taton, Macromolecules 2009, 42, 5996 – 6005.
Yu, M. X. Tan, Y. G. Zhang, Adv. Synth. Catal. 2012, 354, 969 – 974; n) S. [19] M. J. Fuchter, Chem. Eur. J. 2010, 16, 12286 – 12294.
Nurhanna Riduan, J. Y. Ying, Y. G. Zhang, Org. Lett. 2012, 14, 1780 – 1783. [20] A. G. Brook, Acc. Chem. Res. 1974, 7, 77 – 84.
[6] a) G. C. Welch, R. R. San Juan, J. D. Masuda, D. W. Stephan, Science 2006, [21] M. X. Tan, Y. G. Zhang, J. Y. Ying, Adv. Synth. Catal. 2009, 351, 1390 –
42, 4793 – 4795; b) P. A. Chase, G. C. Welch, T. Jurca, D. W. Stephan, 1394.
Angew. Chem. 2007, 119, 8196 – 8199; Angew. Chem. Int. Ed. 2007, 46, [22] Gaussian 03, Revision C.02, M. J. Frisch, G. W. Trucks, H. B. Schlegel, G. E.
8050 – 8053; c) P. Spies, S. Schwendemann, S. Lange, G. Kehr, R. Frçhlich, Scuseria, M. A. Robb, J. R. Cheeseman, J. A. Montgomery, Jr., T. Vreven,
G. Erker, Angew. Chem. 2008, 120, 7654 – 7657; Angew. Chem. Int. Ed. K. N. Kudin, J. C. Burant, J. M. Millam, S. S. Iyengar, J. Tomasi, V. Barone,
2008, 47, 7543 – 7546; d) P. A. Chase, D. W. Stephan, Angew. Chem. B. Mennucci, M. Cossi, G. Scalmani, N. Rega, G. A. Petersson, H. Nakatsu-
2008, 120, 7543 – 7547; Angew. Chem. Int. Ed. 2008, 47, 7433 – 7437; e) V. ji, M. Hada, M. Ehara, K. Toyota, R. Fukuda, J. Hasegawa, M. Ishida, T. Na-
Sumerin, F. Schulz, M. Nieger, M. Leskel, T. Repo, B. Rieger, Angew. kajima, Y. Honda, O. Kitao, H. Nakai, M. Klene, X. Li, J. E. Knox, H. P.
Chem. 2008, 120, 6090 – 6092; Angew. Chem. Int. Ed. 2008, 47, 6001 – Hratchian, J. B. Cross, V. Bakken, C. Adamo, J. Jaramillo, R. Gomperts,
6003. R. E. Stratmann, O. Yazyev, A. J. Austin, R. Cammi, C. Pomelli, J. W. Och-
[7] C. M. Mçmming, E. Otten, G. Kehr, R. Frçhlich, S. Grimme, D. W. Stephan, terski, P. Y. Ayala, K. Morokuma, G. A. Voth, P. Salvador, J. J. Dannenberg,
G. Erker, Angew. Chem. 2009, 121, 6770 – 6773; Angew. Chem. Int. Ed. V. G. Zakrzewski, S. Dapprich, A. D. Daniels, M. C. Strain, O. Farkas, D. K.
2009, 48, 6643 – 6646. Malick, A. D. Rabuck, K. Raghavachari, J. B. Foresman, J. V. Ortiz, Q. Cui,
[8] G. Mnard, D. W. Stephan, J. Am. Chem. Soc. 2010, 132, 1796 – 1797. A. G. Baboul, S. Clifford, J. Cioslowski, B. B. Stefanov, G. Liu, A. Liashenko,
[9] A. E. Ashley, A. L. Thompson, D. O’Hare, Angew. Chem. 2009, 121, P. Piskorz, I. Komaromi, R. L. Martin, D. J. Fox, T. Keith, M. A. Al-Laham,
10023 – 10027; Angew. Chem. Int. Ed. 2009, 48, 9839 – 9843. C. Y. Peng, A. Nanayakkara, M. Challacombe, P. M. W. Gill, B. Johnson, W.
[10] A. Berkefeld, W. E. Piers, M. Parvez, J. Am. Chem. Soc. 2010, 132, 10660 – Chen, M. W. Wong, C. Gonzalez, J. A. Pople, Gaussian, Inc., Wallingford
10661. CT, 2004.
[11] a) S. Charraborty, J. Zhang, J. A. Krause, H. Guan, J. Am. Chem. Soc. 2010, [23] a) A. D. Becke, J. Chem. Phys. 1993, 98, 5648; b) C. T. Lee, W. T. Yang,
132, 8872 – 8873; b) F. Huang, C. Zhang, J. Zhang, Z.-X. Wang, H. Guan, R. G. Parr, Phys. Rev. B 1988, 37, 785 – 789.
Inorg. Chem. 2011, 50, 3816 – 3825.
[12] a) O. Jacquet, C. D. N. Gomes, M. Ephritikhine, T. Cantat, J. Am. Chem. Received: October 15, 2012
Soc. 2012, 134, 2934 – 2937; b) C. Das Neves Gomes, O. Jacquet, C. Villi- Published online on March 6, 2013

 2013 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim ChemCatChem 2013, 5, 1490 – 1496 1496

You might also like