Silylative Coupling of Terminal Alkynes with Iodosilanes: New

Apr 13, 2011 - Table 1. [{Ir(μ-Cl)(CO)2}2]-Catalyzed Silylative Coupling of ... 1:10−2, toluene, 110, NEt(i-Pr)2, 85, 76(9) ... 1:10−2, toluene, ...
0 downloads 0 Views 1MB Size
ARTICLE pubs.acs.org/Organometallics

Silylative Coupling of Terminal Alkynes with Iodosilanes: New Catalytic Activation of sp-Hybridized CarbonHydrogen Bonds Ireneusz Kownacki, Bogdan Marciniec,* Beata Dudziec, and Maciej Kubicki Department of Organometallic Chemistry, Faculty of Chemistry, Adam Mickiewicz University, Grunwaldzka 6, 60-780 Poznan, Poland

bS Supporting Information ABSTRACT: The [{Ir(μ-Cl)(CO)2}2] (I)-catalyzed reaction of terminal alkynes and diynes with Me3SiI with the aid of NEt(i-Pr)2 occurs smoothly, leading to the formation of monoand bis-silyl-functionalized alkynyl derivatives. This new reaction, which occurs via direct activation of the CspH bond in the starting alkyne, is a very efficient and easy tool for the synthesis of unique silylated alkynes. Separate experiments of the equimolar reactions of the precursor (I) with selected reaction substrates provided the evidence for the sequence reactions as well as for the real catalyst to be [IrI(CO)(NEt(i-Pr)2)2].

’ INTRODUCTION Organofunctionalized alkynylsilanes are important synthons of alkynyl carboanions for the synthesis of organic and natural products1 as well as precursors of optoelectronic materials.2 They are usually prepared by classical methods of synthesis with employment of organometallic reagents3 and more recently by zinc saltmediated silylation of terminal alkynes with aminosilanes,4a chlorosilanes,4b and Me3SiOTf4c (OTf = O3SCF3) as well as by TM-catalyzed dehydrogenative silylation of terminal alkynes with hydrosilanes4d (by [Ir4(CO)12]/PPh3 system) and crossmetathesis of functionalized alkynes.4e Recently, we have reported a method for the preparation of alkynylsilanes via silylative crosscoupling of selected terminal alkynes with various vinylsilanes catalyzed by complexes containing [Ru]H and/or [Ru]Si bonds and occurring according to the following equation (Scheme 1).5 Unfortunately, the silylation of phenylethyne with vinylsilane does not take place because of the concurrent dimerization of the initial alkyne preferred under the conditions of catalysis. Vinylsilicon compounds in this reaction act as silylating agents and hydrogen acceptors. For general mechanistic implications see a recent review.6 Most of the above-mentioned methods have many drawbacks and limitations, in particular intolerance to functional groups such as OH, NH2, C(O), NdCR, or (C(O))2NR present in the substituents bonded to the Csp moiety. Therefore, in this paper we present a new catalytic reaction that is a coupling of various terminal alkynes with iodotrimethylsilane catalyzed by iridium(I) complexes in the presence of tertiary organic amine as a hydrogen halogen acceptor.7 ’ RESULTS AND DISCUSSION A new horizon of this strategy is to extend the silylative agents (over vinylsilanes) to SiX derivatives with elimination of a [base 3 H]X adduct as a byproduct according to Scheme 2. Various iridium(I) complexes were examined as catalysts of the silylative coupling reaction of phenylacetylene with Me3SiI as r 2011 American Chemical Society

Scheme 1. Ruthenium-Catalyzed Silylative Coupling of Terminal Alkynes with Vinyl-Substituted Organosilicon Compounds

a representative reagent system. The reaction was found to proceed efficiently in the presence of a catalytic system consisting of iridium(I) chlorocarbonyl complex [{Ir(μ-Cl)(CO)2}2] (I) and NEt(i-Pr)2 as a hydrogen iodide acceptor, leading to the formation of appropriate silyl-substituted alkynyl derivatives and evolution of ammonium salt according to Scheme 2 (Table 1). Surprisingly, other very popular iridium(I) complexes appeared to be unselective ([{Ir(μ-Cl)(cod)}2 ]) or even inactive ([Ir(Cl)(cod)(PCy3)], trans-[Ir(Cl)(CO)(PPh3)2], trans-[Ir(Cl)(CO)(PCy 3 )2 ]) in this reaction. The process of optimization of the reaction conditions showed that other parameters, not only the type of catalyst but also the types of amine and solvent, had a strong influence on the selectivity and yield of the products formed. Catalytic tests performed for the model substrate system, i.e., phenylacetylene and Me3SiI, proved that a too low reaction temperature and too low loading of the iridium catalyst do not give acceptable yields of the desired product. As follows from the results obtained for this reaction at different temperatures, i.e., 80 and 110 °C, a definitely higher conversion as well as yield of the target compound was observed for the process led at 80 °C. We suppose that side processes occurring in the catalytic system at higher temperature, e.g., dimerization and trimerization of the initial alkyne (see Table 1), effectively deactivate catalytically active iridium species. Moreover, lower efficiency of the process studied at a higher temperature can also be caused by thermal Received: January 17, 2011 Published: April 13, 2011 2539

dx.doi.org/10.1021/om200038r | Organometallics 2011, 30, 2539–2545

Organometallics

ARTICLE

Scheme 2. Coupling of Terminal Alkynes with Iodotrimethylsilane (Me3SiI)

Table 1. [{Ir(μ-Cl)(CO)2}2]-Catalyzed Silylative Coupling of Phenylacetylene with Me3SiI: Optimization of the Reaction Conditionsa solvent

temp [°C]

base

conversion [%]

1:102

toluene

60

NEt(i-Pr)2

36

1:102b

toluene

80

NEt(i-Pr)2

97

97

1:102

toluene

110

NEt(i-Pr)2

85

76(9)c

1:2  103

toluene

80

NEt(i-Pr)2

21

21

1:102

toluene

80

NEt3

68

68

1:102 1:102

toluene THF

80 80

C5H5N NEt(i-Pr)2

0 68

0 68

1:102

DMF

80

NEt(i-Pr)2

83

83

[PhCtCH]:[Ir]

yield [%] 36

a

Reaction conditions: [Me3SiI]:[NEt(i-Pr)2] = 1.6:1.8, 24 h. b 12 h, closed system. Conversion and yield were determined by GC analysis. Decane was used as internal standard. c The target compound formation was accompanied with dimerization and trimerization products of the initial alkyne.

instability of species IV (see Scheme 4). Further tests performed for other exemplary amines proved that less hindered amines efficiently block catalytic activity of the initial iridium complex (see Table 1). In addition, in this reaction, more polar solvents containing in their structure donating heteroatoms suppress effectively the activity of the iridium(I) catalytic system used. On the basis of the data collected in Table 1, the iridium(I) binuclear complex [{Ir(μ-Cl)(CO)2}2] with NEt(i-Pr)2 was chosen for further study of this reaction. Silylative coupling of selected alkynes, i.e., alkyl, cycloalkyl, silyl, aryl as well as H2N- or HO-functionalized alkynes with iodotrimethylsilane catalyzed by the precursor [{Ir(μ-Cl)(CO)2}2] (I) under the optimum conditions gives the respective alkynylsilane exclusively (113) (see the data presented in Table 2). Most of the new silyl-alkynyl derivatives were isolated and fully characterized by spectroscopic methods (see the Supporting Information). The method developed seems to be universal, because it can be utilized efficiently for silylation of not only nonfunctionalized alkynes. This iridium(I) catalytic system appeared to be resistant to the initial ethyne derivatives containing reactive functionalities, such us OH and NH2, which in the first step are O- or N-silylated by the appropriate equivalent of Me3SiCl or Me3SiI and further coupled with Me3SiI (see the Supporting Information). It is worth emphasizing that the silylation of terminal diynes by iodotrimethylsilane (see Table 3) gives exclusively and mostly quantitatively bissilyl derivatives (1418), which confirms the advantages of this method over the coupling of terminal alkynes with vinylsilanes catalyzed by ruthenium complexes5 as well as other methods.1,4 All new bissilyl-functionalized diynes were isolated and fully characterized by spectroscopic methods (see the Supporting Information), and the structures of compounds 17 and 18 have been solved by an X-ray method (Figures 1 and 2).

According to the available crystallographic data for substituted alkynes, the C1C2 distance of 1.212(17) Å supports its triplebond character. Similar bond lengths were found in our bissilylfunctionalized diynes, 1.202(2) and 1.203(2) Å in 18, and in 17 these distances are in the range 1.197(3)1.201(3) Å. In the latter compound, there are two symmetry-independent molecules in the asymmetric part of the unit cell; the geometrical features of both molecules are very similar. Formation of the silyl-functionalized alkyne derivative as well as the hydroamonium iodide, according to Scheme 2, suggests that the reaction of terminal alkynes with Me3SiI occurs through the iridium-mediated CspH bond activation in the initial terminal alkyne/diyne molecule, and it seems to be a key step of this new process. The CH bond activation of 1-alkynes by a transition-metal catalyst constitutes one of the most important methods in the preparation of enynes. Since the Glaser8a coupling of terminal acetylenes under copper catalysis with oxidation to give butadiynes and the Strauss8b nonoxidative variant of coupling were discovered, many papers have been published dealing with the transition-metalcatalyzed transformation of terminal alkynes via CspH bond activation, particularly with employment of Ru,9 Rh,10 and Pd11 complexes. Recently also complexes of Ni,10mo,12 Cu,13 Au,14,13l Co,15 and Fe16 have been reported to catalyze the reaction of terminal alkynes via alkynyl CH bond cleavage. In the field of iridium-catalyzed transformations of terminal alkynes, Miyaura et al. have reported that the iridium-catalyzed dimerization of terminal alkynes involves the oxidative addition of alkyne to a low-valent Ir complex, followed by insertion of an alternative alkyne to given dimers.17 Also Carreira et al. reported the addition of TMSacetylene to various aldimines catalyzed by [{Ir(μ-Cl)(cod)}2], which occurs by CspH bond activation, leading to the corresponding amine-alkynyl derivatives.18 The same conclusion was drawn by Ishii19 and Toyota20 studying the activity of iridium complexes in dimerization of terminal alkynes and co-coupling of the terminus with internal alkynes. On the basis of the study of iridium-catalyzed cross-coupling of the terminus with internal alkynes, Ishii19 proposed the reaction mechanism based on the spectroscopic determination of the structures of the catalytically obtained products using Csp deuteriumlabeled phenylacetylene, whereas the mechanism we propose in this paper is based on a series of experiments involving the equimolar reactions of the precursor l with initial substrates. Direct reaction of precursor l with model silylethyne ((i-Pr)3SiCtCH) in the presence of pyridine enabled us to isolate six-coordinate iridium(III) complex IIIa. Formation of such alkynyl-hydrydo iridium(III) complexes as products of various terminal alkynes' oxidative addition to [Ir(η2-acac)(PPh3)2] has been previously reported by Oro.21 The iridium(III) organometallic species IIIa was fully characterized by 1H and 13C NMR spectroscopy. In the 1H NMR spectrum, the lines at 9.42 (dm), 6.68 (dtt), and 6.37 (dtm) ppm were assigned to two pyridine ligands bonded to the iridium center. The resonance lines at 1.28 (s) and 1.26 (s) assigned to i-Pr groups at silicon, the singlet in the hydride region at 17.54 ppm, and the lines at 100.01 and 97.24 ppm in the 13C NMR spectrum, assigned to an ethynyl moiety, unambiguously confirmed the formation of the silylethynyl-hydride complex IIIa via CspH bond activation in the initial silylethyne, according to Scheme 3. The structure of complex IIIa was also solved by X-ray analysis, giving direct evidence of activation of the CspH bond by an iridium(I) carbonyl complex. 2540

dx.doi.org/10.1021/om200038r |Organometallics 2011, 30, 2539–2545

Organometallics

ARTICLE

Table 2. Iridium(I)-Catalyzed Silylation of Terminal Alkynesa

Reaction conditions: [alkyne]:[Me3SiI]:[NEt(i-Pr)2]:[Ir] = 1:1.6:1.8:102, 80 °C. b [alkyne]:[Me3SiI]:[NEt(i-Pr)2]:[Ir] = 1:1.6:1.8:(2  102), 80 °C. c [alkyne]:[Me3SiI]:[NEt(i-Pr)2]:[Ir] = 1:3.8:4.2:(2  102), 80 °C. The reactions were performed in a closed system. [Ir] = [{Ir(µ-Cl)(CO)2}2]. Toluene was used as a solvent: 8 mL for runs 111; 18 mL for runs 12 and 13. Conversion and yield were determined by GC analysis and calculated using solvent as a standard. a

The coordination number of the iridium atom is 6, and the geometry of the coordination is close to that of a distorted octahedron. The position of the hydrogen atom connected with iridium was found in the difference Fourier map and is consistent with both chemical arguments and the data from the Cambridge Structural Database. In version 5.31 (November 2009 updated August 2010) there are 676 examples of structures with an IrH bond, and more than a half of them (387) have the coordination number 6. The IrH distances have a mean value of 1.59 Å, consistent with the length of 1.58 Å found in the present structure. The C1C2 distance of 1.212(17) Å supports its triple-bond character.

As shown in Figure 3, hydride and (i-Pr)3SiCtC ligands occupy the cis position, while the chloride ligand occupies the trans position in relation to the ethynyl ligand and the cis position to the hydride ligand. Such a spatial configuration of the ligands attached to the iridium center allows further steps, i.e., dehydrohalogenation of complex III by tertiary amine and formation of iridium ethynyl species IV in the catalytic cycle (Scheme 4). A separate experiment of the equimolar reaction of precursor I with HCtCSi(i-Pr)3 but in the presence of not only amine (NEt(i-Pr)2) but also iodosilane (monitored by GC-MS and NMR techniques) provided evidence of the sequence reactions 2541

dx.doi.org/10.1021/om200038r |Organometallics 2011, 30, 2539–2545

Organometallics

ARTICLE

Table 3. Iridium(I)-Catalyzed Silylation of Diynesa

a Reaction conditions: [diyne]:[Me3SiI]:[NEt(i-Pr)2]:[Ir] = 1:3.4:3.6:4  102, 80 °C. The reactions were performed in a closed system. [Ir] = [{Ir(μ-Cl)(CO)2}2]. Toluene was used as a solvent: 6 mL for runs 418. Conversion and yield were determined by GC analysis and calculated using solvent as a standard.

Figure 1. Perspective view of one of the molecules of 17 with atom labeling. Ellipsoids are drawn at the 33% probability level; hydrogen atoms are shown as spheres of arbitrary radii. Selected geometrical parameters (distances, Å; angles, deg; the values for the second symmetry-independent molecules are given in square brackets): Si1C11 1.850(3) [1.850(3)], Si1C21 1.862(2) [1.862(3)], Si1C31 1.836(2) [1.831(2)], C31C32 1.197(3) [1.201(3)], C32Si3 1.841(2) [1.841(2)], Si1C41 1.830(2) [1.834(2)], C41C42 1.196(3) [1.200(3)], C42Si4 1.845(3) [1.842(2)], Si1 C31C32 178.7(2) [174.1(2)], C31C32Si3 176.6(3) [177.7(3)], Si1C41C42 177.5(3) [176.3(2)], C41C42Si4 178.2(3) [178.1(3)].

of oxidative addition of alkynes, followed by the elimination of ammonium chloride to form complex IV and subsequent oxidative addition of iodosilane to get complex V. Its reductive

Figure 2. Perspective view of one of the molecules of 18 with atom labeling. Ellipsoids are drawn at the 33% probability level; hydrogen atoms are shown as spheres of arbitrary radii. Selected geometrical parameters (distances, Å; angles, deg): Si1C11 1.831(2), Si1C21 1.828(2), Si1C31 1.866(2), Si1C41 1.864(2), C11C12 1.202(2), C21C22 1.203(2), Si1C11C12 178.3(2), C11C12Si13 178.9(2), Si1 C21C22 176.7(2), C21C22Si23 177.3(2).

elimination leads to formation of appropriate silylethyne and regeneration of the tetracoordinated IrI complex II (Scheme 5). The latter II appeared to be the catalytically active species of the cycle illustrated in Scheme 4. The singlet at 19.66 ppm assigned to the [Ir]-H moiety in the 1H NMR spectrum recorded to monitor the reaction of precursor I with HCtCSi(i-Pr)3 and NEt(i-Pr)2 at 2542

dx.doi.org/10.1021/om200038r |Organometallics 2011, 30, 2539–2545

Organometallics

ARTICLE

Scheme 3. Activation of a HCsp Bond: A Key Step of the Mechanism

Scheme 4. Mechanism Proposed for the Silylative Coupling of Terminal Alkynes with Me3SiI

Figure 3. X-ray structure of the species IIIa. Selected geometrical parameters (distances, Å; angles, deg): Ir1Cl1 2.412(3), Ir1N21 2.119(8), Ir1N15 2.201(11), Ir1C13 1.827(17), Ir1C1 1.969(15), Ir1H1 1.58, C1C2 1.212(17), C13O14 1.148(17), C13Ir1 N21 171.2(5), C1Ir1Cl1 171.9(3), N15Ir1H1 158, Ir1C1 C2 176.9(11), Ir1C13C14 178.2(16).

50 °C as well as the lines at 98.99 and 95.23 ppm in 13C NMR spectrum assigned to the (i-Pr)3SiCtC ligand confirm the formation of complex III, which at higher temperatures and in an excess of organic base (tertiary amine) gives complex IV. Formation of the latter species is confirmed by the appearance of new lines at 94.21 and 90.83 in the 13C NMR spectrum, assigned to (i-Pr)3SiCtC-[Ir] species, and by the disappearance of the singlet at 19.66 ppm in 1H NMR spectrum after the reaction at 80 °C. Addition of ISiMe3 leads to complex V, which in the next step is transformed to complex II with evolution of 1-triisopropylsilyl-2trimethylsilylethyne (confirmed by GC-MS analysis).

’ CONCLUSIONS In conclusion, we have described a novel catalytic route for the activation of CspH bonds via the silylative coupling of terminal (also carbon functionalized) alkynes and diynes by iodotrimethylsilane in the presence of [{Ir(μ-Cl)(CO)2}2] as a catalyst precursor. This new reaction under optimum conditions appears to be a very efficient and easy method for the synthesis of novel and unique mono- or multisilylated alkynes, which might be used as building blocks for construction of molecular and macromolecular organic compounds with precisely dedicated chemical, electronic, and optoelectronic properties. ’ EXPERIMENTAL SECTION General Methods and Chemicals. All synthesis and manipulations were carried out under argon using standard Schlenk-line and vacuum techniques. 1H, 13C, and 29Si NMR spectra were recorded on

Varian Gemini 300 VT and Varian Mercury 300 VT spectrometers in benzene-d6 and CDCl3. The chemicals were obtained from the following sources: IrCl3 3 3H2O from Pressure Chemicals; cod, PPh3, PCy3, C6D6, CDCl3, DMF, THF, decane, C5H5N, NEt3, NEt(i-Pr)2, Me3SiI, toluene from Aldrich; CO from Linde Gas. The complexes [{Ir(μ-Cl)(cod)}2],22 [Ir(Cl)(cod)(PCy3)],23 trans-[Ir(Cl)(CO)(PPh3)2],24 [Ir(Cl)(CO)(PCy3)2],25 and [{Ir(μ-Cl)(CO)2}2]26 were synthesized according to the published methods. All solvents and liquid reagents were dried and distilled under argon prior to use. General Procedure for Catalytic Tests. The glass Schlenk reactor (10 mL) equipped with a magnetic stirring bar was evacuated and flushed with argon. The calculated amount of the complex [{Ir(μCl)(CO)2}2] (0.005 or 0.0025 mmol) was placed in the reactor in the flow of argon; then 3 mL of solvent and a calculated amount of amine (1.8 mmol) were added. The mixture obtained was stirred for about 10 min. In the next step, appropriate amounts of phenylacetylene (0.102 g, 1 mmol), Me3SiI (0.32 g, 1.6 mmol), and decane (5% of initial mixture volume) were added, and the reaction was conduced at the given temperature. The reaction mixture was analyzed by GC and GC-MS at the beginning and after 24 h. Conversion and yield were calculated using the internal standard calculation method. General Procedure for Silylation of Terminal Alkynes. The glass Schlenk reactor equipped with a magnetic stirring bar was evacuated and flushed with argon. The calculated amount of the complex [{Ir(μ-Cl)(CO)2}2] was placed in the reactor in the flow of argon; then toluene and a calculated amount of NEt(i-Pr)2 were added. The mixture obtained was stirred for 10 min. In the next step, appropriate amounts of terminal alkyne or diyne and Me3SiI were added. The reaction was conducted in a closed system at 80 °C upon stirring, until the full conversion of initial alkyne/diyne. After the reaction completion, toluene and the excess of other reagents were evaporated under reduced pressure. The crude product was isolated by purification on the SiO2 column with hexane as eluent. 2543

dx.doi.org/10.1021/om200038r |Organometallics 2011, 30, 2539–2545

Organometallics

ARTICLE

Scheme 5. Transformation of the Initial Iridium(I) Precursor

Synthesis of [Ir(Cl)(CO)(H)(CtCSi(i-Pr3))(py)2] (IIIa). Pyridine (0.316 g, 4.00 mmol) was introduced into a vigorously stirred suspension of [{Ir(μ-Cl)(CO)2}2] (0.284 g, 0.5 mmol) in 5 mL of toluene at room temperature under a dry argon atmosphere over 10 min; then to the yellow solution obtained was added HCtCSi(i-Pr)3 (0.218 g, 1.2 mmol). The reaction was conducted for 24 h at 80 °C. After this time the content was cooled to room temperature. Subsequently, the reaction mixture was filtered off by a cannula system, then the solvent was removed under reduced pressure, and pentane was added. The white suspension obtained was washed with pentane by decantation. Yield: 0.536 g (90%). 1 H NMR (300 MHz, C6D6, 300 K): δ (ppm) 9.42 (dm), 6.68 (dtt), 6.37 (dtm) (10 H, C5H5N); 1.28 (s), 1.26 (s) (21H, i-Pr); 17.54 (s, 1H, Ir-H). 13C NMR (75.42 MHz, C6D6, 300 K): δ (ppm) 163.42 (CO); 153.58, 151.66, 138.05, 137.94, 125.08, 124.47 (C5H5N), 100.01, 97.24 (CtCSi(i-Pr)3); 19.36 (Me); 12.62 (CHMe2). 29Si NMR (59.59 MHz, C6D6, 300 K): δ (ppm) 6.48 (Si(i-Pr)3). Sequential Stoichiometric Reaction of Precursor I with (i-Pr)3SiCtCH, NEt(i-Pr)2, and ISiMe3. A portion of NEt(i-Pr)2 (0.168 g, 1.3 mmol) was introduced into a vigorously stirred suspension of [{Ir(μ-Cl)(CO)2}2] (0.0738 g, 0.13 mmol) in 0.6 mL of benzene-d6 in a Young NMR tube at room temperature under a dry argon atmosphere. After 30 min to the yellow solution obtained was added HCtCSi(i-Pr)3 (0.0766 g, 0.387 mmol). The reaction was carried out for 3 h at 50 °C. After NMR analysis the reaction was continued for 18 h at 80 °C. Afterward, to the dark red solution was added 0.078 g (0.39 mmol) of ISiMe3. The reaction was monitored by NMR and GC-MS techniques. Crystallography. For crystal data, data collection, and structure refinement see the Supporting Information. Crystallographic data (excluding structure factors) for the structural analysis have been deposited with the Cambridge Crystallographic Data Centre, Nos. CCDC-762977 (IIIa), CCDC-806590 (17), and CCDC806591 (18). Copies of this information may be obtained free of charge from The Director, CCDC, 12 Union Road, Cambridge, CB2 1EZ, UK. Fax: þ44(1223)336-033, e-mail:[email protected], or www: www.ccdc.cam.ac.uk.

’ ASSOCIATED CONTENT

bS

Supporting Information. This material is available free of charge via the Internet at http://pubs.acs.org.

’ AUTHOR INFORMATION Corresponding Author

*Tel: þ48 061 8291366. Fax: þ48 061 8291508. E-mail: bogdan. [email protected].

’ ACKNOWLEDGMENT This work was made possible by grant no. N 204 443840 from the Ministry of Science and Higher Education.

’ REFERENCES (1) (a) Greene, T. W.; Wuts, P. G. M. Protective Groups in Organic Synthesis; Wiley: New York, 1999. (b) Casida, J. E. J. Agric. Food Chem. 1991, 39, 1335. (c) Anderson, C. J.; Munday, R. H. J. Org. Chem. 2004, 69, 8971. (d) Takahashi, S.; Kuroyama, Y.; Sonogashira, K.; Hagihara, N. Synthesis 1980, 627. (e) Uenishi, J.; Matsui, K. Tetrahedron Lett. 2001, 42, 4353. (f) Tykwinski., R. R. Angew. Chem., Int. Ed. 2003, 42, 1566. (g) Negishi, E.; Anastasia, L. Chem. Rev. 2003, 103, 1979. (2) (a) Brizius, G.; Bunz, U. H. F. Org. Lett. 2002, 4, 2829. (b) Yamamoto, Y. Chem. Rev. 2008, 108, 3199. (c) Tour, J. Acc. Chem. Res. 2000, 33, 791. (d) Carrolland, R. L.; Gorman, Ch. B. Angew. Chem., Int. Ed. 2002, 41, 4378. (e) Nishihara, Y.; Ando, J.; Kato, T.; Mori, A. Macromolecules 2000, 33, 2779. (f) Martinand, R. E.; Diederich, F. Angew. Chem., Int. Ed. 1999, 38, 1350. (g) Castanet, A.-S.; Colobert, F.; Schlama, T. Org. Lett. 2000, 2, 3559. (h) Blumrnkopf, T. A.; Overman, L. E. Chem. Rev. 1986, 86, 857. (3) (a) Davidsohn, W. E.; Henry, M. C. Chem. Rev. 1967, 67, 73. (b) Colvin, E. Silicon in Organic Synthesis; Butterworths: London, 1981; p 165. (c) Brandsma, L.; Verkruijsse, H. D. Synthesis 1999, 1727. (d) Weber, W. P. Silicon Reagents for Organic Synthesis; Springer-Verlag: Berlin, 1983; p 129. (4) (a) Anreev, A. A.; Konshin, V. V.; Komarov, N. V.; Rubin, M.; Brouwer, C.; Gevorgyan, V. Org. Lett. 2004, 6, 421. (b) Jiang, H.; Zhu, S. Tetrahedron Lett. 2005, 46, 517. (c) Rahaim, R. J., Jr.; Shaw, J. T. J. Org. Chem. 2008, 73, 2912. (d) Shimizu, R.; Fuchikami, T. Tetrahedron Lett. 2000, 41, 907. (e) F€urstner, A.; Mathes, Ch. Org. Lett. 2001, 3, 221. (5) (a) Marciniec, B.; Dudziec, B.; Kownacki, I. Angew. Chem., Int. Ed. 2006, 45, 8180.(b) Marciniec, B.; Dudziec, B.; Szubert, K. Polish Patent Application P 380 422, 2006 (c) Marciniec, B.; Dudziec, B.; Szubert, K.; Walkowiak, J. WO 2008020774, 2008. (d) Dudziec, B.; Marciniec, B. Organometallics 2008, 27, 5598. (6) Marciniec, B. Acc. Chem. Res. 2007, 40, 943. (7) (a) Kownacki, I.; Marciniec, B.; Dudziec, B.; Kownacka, A. Polish Patent Application P 393475, 2010. (b) Kownacki, I.; Marciniec, B.; Dudziec, B.; Majchrzak, M.; Kownacka, A. Polish Patent Application P 393476, 2010. (c) Kownacki, I.; Marciniec, B. Polish Patent Application P 394454, 2011. (8) (a) Glaser, C. Ber. Dtsch. Chem. Ges. 1869, 2, 422. (b) Strauss, F. Liebigs Ann. Chem. 1905, 342, 190. (9) (a) Bianchini, C.; Peruzzini, M.; Zanobini, F.; Frediani, P.; Albinati, A. J. Am. Chem. Soc. 1991, 113, 5453. (b) Dahlenburg, L.; Frosin, K.-M.; Kerstan, S.; Werner, D. J. Organomet. Chem. 1991, 407, 115. (c) Echavarren, A. M.; Lopez, J.; Santos, A.; Montoya, J. J. Organomet. Chem. 1991, 414, 393. (d) Bruce, M. I.; Swincer, A. G. Adv. Organomet. Chem. 1983, 22, 59. (e) Bruce, M. I. Chem. Rev. 1991, 91, 197. (f) Schafer, M.; Mahr, N.; Wolf, J.; Werner, H. Angew. Chem., Int. Ed. Engl. 1993, 32, 1315. (g) Wakatsuki, Y.; Yamazaki, H.; Kumegawa, N.; Satoh, T.; Satoh, J. Y. J. Am. Chem. Soc. 1991, 113, 9604. (h) Naota, T.; Takaya, H.; Murahashi, S. I. Chem. Rew. 1998, 98, 2599. (i) Dyker, G. Angew. Chem. 1999, 38, 1698. (j) Li, C. J.; Wei, C. M. Chem. Commun. 2002, 268. (k) Wei, C. M.; Li, C. J. Green Chem. 2002, 39. (l) Yi, Ch. S.; Yun, S. Y.; Guzei, L. A. J. Am. Chem. Soc. 2005, 127, 5782. (10) (a) Singer, H.; Wilkinson, G. J. Chem. Soc. (A) 1968, 849. (b) Schmitt, H. J.; Singer, H. J. Organomet. Chem. 1978, 153, 165. (c) Schafer, H.; Marcy, R.; Ruping, T.; Singer, H. J. Organomet. Chem. 2544

dx.doi.org/10.1021/om200038r |Organometallics 2011, 30, 2539–2545

Organometallics 1982, 240, 17. (d) Carlton, L.; Read, G. J. Chem. Soc., Perkin Trans. 1978, 1, 1631. (e) Yoshikawa, S.; Kiji, J.; Furukawa, J. Makromol. Chem. 1977, 178, 1007. (f) Aresta, M.; DeFazio, M. J. Organomet. Chem. 1980, 186, 109. (g) Albano, P.; Aresta, M. J. Organomet. Chem. 1980, 190, 243. (h) Kovalev, I. P.; Yevdakov, K. V.; Strelenko, Y. A.; Vinogradov, M. G.; Nikishin, G. I. J. Organomet. Chem. 1990, 386, 139. (i) Yamaguchi, M.; Omata, K.; Hirama, M. Tetrahedron Lett. 1994, 35, 5689. (j) Nishimura, T.; Guo, X.-X.; Ohnishi, K.; Hayashi, T. Adv. Synth. Catal. 2007, 349, 2669. (k) Katagiri, T.; Tsurugi, H.; Funayama, A.; Satoh, T.; Miura, M. Chem. Lett. 2007, 36, 830. (l) Katagiri, T.; Tsurugi, H.; Satoh, T.; Miura, M. Chem. Commun. 2008, 3405. (m) Matsuyama, N.; Tsurugi, H.; Satoh, T.; Miura, M. Adv. Synth. Catal 2008, 350, 2274. (n) Shirakura, M.; Suginome, M. Org. Lett. 2009, 11, 523. (o) Shirakura, M.; Suginome, M. J. Am. Chem. Soc. 2009, 131, 5060. (11) (a) Modern Acetylene Chemistry; Stang, P. J., Diederich, F., Eds.; VCH: Weinheim, 1995. (b) Brandsma, L.; Vasilevsky, S. F.; Verkruijsse, H. D. Application of Transition Metal Catalysts in Organis Synthesis; Springer: Berlin, 1998; p 198. (c) Sabourin, E. T. J. Mol. Catal. 1984, 26, 363. (d) Trost, B. M.; Chan, C.; Ruhter, G. J. Am. Chem. Soc. 1987, 109, 3486. (e) Trost, B. M.; Matsubara, S.; Caringi, J. J. J. Am. Chem. Soc. 1989, 111, 8745. (f) Trost, B. M.; Li, C.-J. Synthesis 1994, 1267. (g) Trost, B. M.; McIntosh, M. C. J. Am. Chem. Soc. 1995, 117, 7255. (h) Trost, B. M.; Frontier, A. J. J. Am. Chem. Soc. 2000, 122, 11727. (i) Trost, B. M.; Kottirsch, G. J. Am. Chem. Soc. 1990, 112, 2816. (j) Trost, B. M.; Sorum, M. T.; Chan, C.; Harms, A. E.; Ruhter, G. J. Am. Chem. Soc. 1997, 119, 698. (k) Ishikawa, M.; Ohshita, J.; Ito, Y.; Minato, A. J. Organomet. Chem. 1988, 346, C58. (l) Ohshita, J.; Furumori, K.; Matsuguchi, A.; Ishikawa, M. J. Org. Chem. 1990, 55, 3277. (m) Bruyere, D.; Grigg, R.; Hinsley, J.; Hussain, R. K.; Korn, S.; De La Cierva, C. O.; Sridharan, V.; Wang, J. Tetrahedron Lett. 2003, 44, 8669. (n) Rubin, M.; Markov, J.; Chuprakov, S.; Wink, D. J.; Gevorgyan, V. J. Org. Chem. 2003, 68, 6251. (o) Kim, S. H.; Yoon, J.; Chang, S. Org. Lett. 2011, 13, 1474. (p) Tenaglia, A.; Giordano, L.; Buono, G. Org. Lett. 2006, 8, 4315. (q) Yin, J.; Chisholm, J. D. Chem. Commun. 2006, 632. (r) Tsukada, N.; Ninomiya, S.; Aoyama, Y.; Inoue, Y. Org. Lett. 2007, 9, 2919. (12) (a) Ogata, K.; Sugasawa, J.; Atsuumi, Y.; Fukuzawa, S. Org. Lett. 2010, 12, 148. (b) Matsuyama, N.; Tsurugi, H.; Satoh, T.; Miura, M. Adv. Synth. Catal 2008, 350, 2274. (c) Ogata, K.; Murayama, H.; Sugasawa, J.; Suzuki, N.; Fukuzawa, S.-I. J. Am. Chem. Soc. 2009, 131, 3176. (d) Ogata, K.; Sugasawa, J.; Fukuzawa, S.-I. Angew. Chem., Int. Ed. 2009, 48, 6078. (e) Shirakura, M.; Suginome, M. J. Am. Chem. Soc. 2008, 130, 5410. (13) (a) Siemsen, P.; Livingston, R. C.; Diederich, F. Angew. Chem., Int. Ed. 2000, 39, 2632. (b) Ju, Y.; Li, C.-J.; Varma, R. S. QSAR Comb. Sci. 2004, 23, 891. (c) Gommermann, N.; Koradin, C.; Polborn, K.; Knochel, P. Angew. Chem., Int. Ed. 2003, 42, 5763. (d) Koradin, C.; Gommermann, N.; Polborn, K.; Knochel, P. Chem.—Eur. J. 2003, 9, 2797. (e) Gommermann, N.; Gehrig, A.; Knochel, P. Synlett 2005, 2796. (f) Gommermann, N.; Knochel, P. Synlett 2005, 2799. (g) Yasuike, S.; Kofink, C. C.; Kloetzing, R. J.; Gommermann, N.; Tappe., K.; Gavryushin, A.; Knochel, P. Tetrahedron: Asymmetry 2005, 16, 3385. (h) Gommermann., N.; Knochel, P. Tetrahedron 2005, 61, 11418. (i) Gommermann, N.; Knochel, P. Chem.— Eur. J. 2006, 12, 4380. (j) Knoepfel, T. F.; Aschwanden, P.; Ichikawa, T.; Watanabe, T.; Carreira, E. M. Angew. Chem., Int. Ed. 2004, 43, 5971. (k) Aschwanden., P.; Stephenson, C. R .J.; Carreira, E. M. Org. Lett. 2006, 8, 2437. (l) Huang, B. S.; Yao, X. Q.; Li, C.-J. Adv. Synth. Catal. 2006, 348, 1528. (m) Koradin., C.; Polborn, K.; Knochel, P. Angew. Chem., Int. Ed. 2002, 41, 2535. (n) Wang, M.; Li, P. H.; Wang, L. Eur. J. Org. Chem. 2008, 2255. (o) Groossen, L. J.; Rodriguez, N.; Manjolinho, F.; Lange, P. P. Adv. Synth. Catal. 2010, 352, 2913. (14) (a) Lo, V. K. Y.; Liu, Y.; Wong, M. K.; Che, C. M. Org. Lett. 2006, 8, 1529. (b) de Haro, T.; Nevado, C. J. Am. Chem. Soc. 2010, 132, 1512. (15) Chen, W.-W.; Bi, H.-P.; Li, Ch.-J. Synlett 2010, 3, 475. (16) Kohno, K.; Nakagawa, K.; Yahagi, T.; Choi, J.-C.; Yasuda, H.; Sakakura, T. J. Am. Chem. Soc. 2009, 131, 2784. (17) Ohmura, T.; Yarozuya, S.; Yamamoto, Y.; Miyaura, N. Organometallics 2000, 19, 365.

ARTICLE

(18) Fischer, C.; Carreira, E. M. Org. Lett. 2001, 3, 4319. (19) (a) Satoshi, S.; Takashi, K.; Ishii, Y. Angew. Chem., Int. Ed. 2001, 40, 2534. (b) Hirabayashi, T.; Sakaguchi, S.; Ishii, Y. Adv. Synth. Catal. 2005, 347, 872. (20) Ogata, K.; Toyota, A. J. Organomet. Chem. 2007, 692, 4139. (21) Esteruelas, M. A.; Lahoz, F. J.; Onate, E.; Oro, L. A.; Rodrıguez, L. Organometallics 1996, 15, 823. (22) Herde, J. L.; Lambert, J. V.; Senoff, C. V. Inorg. Synth. 1974, 15, 18. (23) Kownacki, I.; Kubicki, M.; Marciniec, B. Inorg. Chim. Acta 2002, 334, 301. (24) Vrieze, K.; Collman, J. P.; Sears, C. P., Jr.; Kubota, M. Inorg. Synth. 1968, 11, 101. (25) Kownacki, I.; Marciniec, B.; Macina, A.; Rubinsztajn, S.; Lamb, D. Appl. Catal. A: Gen. 2007, 317, 53. (26) Roberto, D.; Cariati, E.; Psaro, R.; Ugo, R. Organometallics 1994, 13, 4227.

2545

dx.doi.org/10.1021/om200038r |Organometallics 2011, 30, 2539–2545