Two-Silicon Cycle for Carbonyl Hydrosilylation with Nikonov's Cationic

Oct 26, 2017 - Citation data is made available by participants in Crossref's Cited-by Linking service. For a more comprehensive list of citations to t...
0 downloads 0 Views 384KB Size
Subscriber access provided by University of Missouri-Columbia

Article

Two-Silicon Cycle for Carbonyl Hydrosilylation with Nikonov's Cationic Ruthenium(II) Catalyst Julien Fuchs, Hendrik F. T. Klare, and Martin Oestreich ACS Catal., Just Accepted Manuscript • DOI: 10.1021/acscatal.7b03336 • Publication Date (Web): 26 Oct 2017 Downloaded from http://pubs.acs.org on October 27, 2017

Just Accepted “Just Accepted” manuscripts have been peer-reviewed and accepted for publication. They are posted online prior to technical editing, formatting for publication and author proofing. The American Chemical Society provides “Just Accepted” as a free service to the research community to expedite the dissemination of scientific material as soon as possible after acceptance. “Just Accepted” manuscripts appear in full in PDF format accompanied by an HTML abstract. “Just Accepted” manuscripts have been fully peer reviewed, but should not be considered the official version of record. They are accessible to all readers and citable by the Digital Object Identifier (DOI®). “Just Accepted” is an optional service offered to authors. Therefore, the “Just Accepted” Web site may not include all articles that will be published in the journal. After a manuscript is technically edited and formatted, it will be removed from the “Just Accepted” Web site and published as an ASAP article. Note that technical editing may introduce minor changes to the manuscript text and/or graphics which could affect content, and all legal disclaimers and ethical guidelines that apply to the journal pertain. ACS cannot be held responsible for errors or consequences arising from the use of information contained in these “Just Accepted” manuscripts.

ACS Catalysis is published by the American Chemical Society. 1155 Sixteenth Street N.W., Washington, DC 20036 Published by American Chemical Society. Copyright © American Chemical Society. However, no copyright claim is made to original U.S. Government works, or works produced by employees of any Commonwealth realm Crown government in the course of their duties.

Page 1 of 6

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Catalysis

Two-Silicon Cycle for Carbonyl Hydrosilylation with Nikonov’s Cationic Ruthenium(II) Catalyst Julien Fuchs, Hendrik F. T. Klare, and Martin Oestreich* Institut für Chemie, Technische Universität Berlin, Strasse des 17. Juni 115, 10623 Berlin, Germany Supporting Information Placeholder ABSTRACT: An experimental analysis proves that Nikonov’s carbonyl hydrosilylation proceeds through a two-silicon cycle rather than the originally proposed one-silicon cycle. The intermediate ruthenium(II) monohydride is not sufficiently hydridic to transfer its hydride onto the silylcarboxonium ion. However, that hydricity is enhanced by oxidative addition of another hydrosilane molecule to afford the corresponding ruthenium(IV) silyl dihydride as the actual hydride donor. The present study also demonstrates that the acetonitrile ligands in Nikonov’s ruthenium(II) catalyst are not innocent. That complex is able to hydrosilylate its own ligand(s), and the resulting N,N-disilylated amine base accounts for competing dehydrogenative silylation of enolizable carbonyl compounds, explaining the formation of a silyl enol ether in substantial quantities next to the expected silyl ether. Both findings lead to a revised mechanistic picture that provides the basis for the development of more efficient and chemoselective catalysts.

KEYWORDS: homogeneous catalysis, hydrosilanes, ionic hydrosilylation, Lewis-acid catalysis, reaction mechanism, ruthenium, Si–H bond activation

INTRODUCTION

F

Ionic outer-sphere mechanisms have been documented for both main-group- and transition-metal-catalyzed carbonyl hydrosilylation with hydrosilanes.1 Prominent examples include contributions by Piers2,3 and Brookhart4,5 employing either B(C6F5)3 (1) or [(POCOP)Ir(H)(acetone)]+[B(C6F5)4]– (2+[B(C6F5)4]–) as catalysts (Figure 1) [POCOP = 2,6-bis(ditert-butylphosphinito)phenyl].6 These processes share electrophilic Si–H bond activation7 through η1- or η2-σ coordination of the Si–H bond by the neutral or cationic Lewis-acidic catalyst (LA) followed by silyl transfer from the activated hydrosilane to the carbonyl oxygen atom (Scheme 1). Their difference lies in the actual reduction step, i.e., delivery of the hydride to the silylcarboxonium ion (gray box). In the case of neutral 1, the silylcarboxonium ion arises as an ion pair together with the borohydride and is eventually reduced by [HB(C6F5)3]–.3 In the other case with cationic pincer complex 2+, the resulting neutral iridium(III) dihydride does not react with the silylcarboxonium ion. Another molecule of the hydrosilane is essential to turn on its ability to release one of its hydrides either by formation of the corresponding hydrosilane–dihydride Lewis adduct or by oxidative addition to the iridium(V) silyl trihydride.5b This conforms with a twosilicon cycle while the main-group catalysis is a one-silicon cycle.

F

+

F

+

O F F

B

F F

F

F F

F

F

(tBu)2P F

F

O Ir O

Me

F

H

P(tBu)2 Ph P Ru 3

NCCH3 NCCH3 [B(C6F5)4]–

[B(C6F5)4]–

Me

1

2+[B(C6F5)4]–

3+[B(C6F5)4]–

one-silicon cycle clarified

two-silicon cycle clarified

this work

Figure 1. Lewis-acid catalysts 1 (Piers),2,3 2+[B(C6F5)4]– (Brookhart),4,5 and 3+[B(C6F5)4]– (Nikonov)8–10 for ionic outersphere carbonyl hydrosilylation. Si

1st hydrosilane molecule Si H

1st hydrosilane molecule Si H

O R1 LA

R2 H

[LA]+

Si H 2nd hydrosilane molecule Si

H

LA

O

LA]– +

good hydride donor

O R1

Si [H

R1

Si +

R2

H

LA

H

poor hydride donor

R2

O R1

LA = B(C6F5)3 (1) one-silicon cycle

ACS Paragon Plus Environment

[LA]+ = 2+[B(C6F5)4]– two-silicon cycle

[LA]+

R2

ACS Catalysis

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 2 of 6

Scheme 1. Ionic Outer-Sphere Carbonyl Hydrosilylation Catalyzed by Main-Group and Transition-Metal Lewis Acids: One-Silicon versus Two-Silicon Cycle (Si = R3Si) Armed with this insight, we asked ourselves whether neutral transition-metal hydrides generally require the assistance of a hydrosilane to close the catalytic cycle. We now report here the reinvestigation of Nikonov’s famous carbonyl hydrosilylation promoted by [Cp(Ph3P)Ru(acetonitrile)2]+[B(C6F5)4]– (3+[B(C6F5)4]–, Figure 1).8–10 Our experimental analysis again supports a two-silicon cycle instead of the postulated onesilicon cycle8 (Scheme 2). The catalytically active 16-electron complex 4+ is formed by reversible dissociation of acetonitrile from 3+. The Si–H bond of the hydrosilane 5 then coordinates to 4+ in η2-σ fashion to form 6+.11 The subsequent silyl transfer from 6+ to the carbonyl oxygen atom of 7 was calculated to be rate determining. The step in question, that is the hydride transfer from the neutral ruthenium(II) monohydride 8 to the silylcarboxonium ion 9+ (gray box), is believed to proceed over a low barrier to regenerate 4+ along with the silyl ether 10. 3+

O Ph

+ CH3CN

SiR3

– CH3CN +

Me H 10

Ph3P

Ru

R3Si 5

NCCH3

H

4+ hydride transfer

silane activation

+

O Ph

SiR3 + Me

9+

Ph3P

Ru

Ru NCCH 3 Ph3P H R3Si 6+

NCCH3 H

8 silyl transfer

O Ph

Me 7

Scheme 2. Proposed One-Silicon Cycle of Nikonov’s Catalytic Carbonyl Hydrosilylation ([B(C6F5)4]− as Counteranion Omitted for Clarity)

ppm -9.3 -9.2 -9.1 -9.0 -8.9 -8.8 -8.7

5.0

4.9

4.8

4.7

4.6

4.5

4.4

4.3

4.2

-8.6 ppm

Figure 2. Selected segment of the 1H/1H NOESY NMR spectrum (500/500 MHz, CD2Cl2, 300 K, Tm = 600 ms) of hydrosilane– ruthenium(II) adduct 6a+[B(C6F5)4]– (Si = Me2PhSi).

Our earlier mechanistic analyses of the carbonyl hydrosilylations catalyzed by B(C6F5)3 (1)3b and Brookhart’s complex 2+[B(C6F5)4]–5b had showcased the use of silicon-stereogenic hydrosilanes as a tool to probe reaction mechanisms by tracking the stereochemical course at the silicon atom.13 Inversion is seen with 13b,13c and stoichiometry-dependent racemization (equimolar hydrosilane) or partial inversion (excess hydrosilane) are obtained with 2+[B(C6F5)4]–.5b Racemization at low hydrosilane concentration relative to acetophenone is explained by the intermediacy of configurationally labile pentacoordinate silicon compounds competing with the hydride transfer at the stage of the silylcarboxonium ion.5b For the stereochemical analysis of Nikonov’s carbonyl hydrosilylation catalyzed by 3+[B(C6F5)4]– we employed enantiomerically enriched acyclic hydrosilane (SiS)-5b13b (e.r. = 95:5, Scheme 3, top). To our surprise, we obtained the expected silyl ether 10b as the minor (!) component of a mixture with the silyl enol ether 11b (see below). Racemization at the silicon atom was found for that silyl enol ether 11b regardless of the amount of (SiS)-5b. To check whether catalyst 3+ is able to erode the enantioenrichment of (SiS)-5b, we treated (SiS)-5b with 2.7 mol % of 3+; no loss of enantiomeric purity was detected after 24 h (Scheme 3, bottom). These results suggest the same scenario for Nikonov’s ruthenium(II) complex 3+[B(C6F5)4]– as for Brookhart’s iridium(III) complex 2+[B(C6F5)4]–.5b

RESULTS AND DISCUSSION Our mechanistic investigation began with an NMR spectroscopic analysis of the Si–H bond activation by complex 3+. In a stoichiometric reaction using hydrosilane Me2PhSiH (5a), the short-lived adduct 6a+ previously described by Nikonov and co-workers8 was studied in detail. Reversibility of the Si– H bond coordination was verified by chemical exchange between the hydride of the coordinated hydrosilane at δ –9.0 ppm and the hydride of the free hydrosilane at δ 4.4 ppm in the 1 H NMR spectrum (Figure 2). The 1JH,Si coupling constant of 56.3 Hz of the coordinated hydrosilane in 6a+ suggests a nonclassical σ interaction between the hydrosilane and the ruthenium center.7b,8 We note here that the generation of 6a+ is accompanied by the formation of doubly hydrosilylated acetonitrile, thereby showing that 3+ is able to hydrosilylate its own ligand(s).12

ACS Paragon Plus Environment

Page 3 of 6

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Catalysis

a Diastereomeric ratios and ratios of silyl ether to silyl enol ether determined by GLC analysis. bEnantiomeric ratios determined by HPLC analysis using chiral stationary phases.

Scheme 3. Stereochemical Stereogenic Silane

Analysis

with

Silicon-

The fact that the silyl enol ether 11b is formed as the major product (with liberation of dihydrogen) raises the question of whether the desired silyl ether 10b traces back to hydrogenation of 11b catalyzed by 3+. To verify this, we used deuteriumlabeled acetophenone-d3 (7-d3) in the hydrosilylation with Me2PhSiH (5a) (Scheme 4, top). The NMR spectroscopic analysis showed no deuterium incorporation at the carbinol carbon atom in silyl ether 10a-d3, hence excluding hydrogenation of 11a-d2 with generated D–H. The attempted reduction of silyl enol ether 11b with dihydrogen under the catalytic setup did not afford the silyl ether 10b (Scheme 4, bottom).

a Ratio of silyl ether to silyl enol ether determined by 1H NMR analysis.

Scheme 5. Ratio of Silyl Ether and Silyl Enol Ether Depending on the Hydrosilane-to-Acetophenone Ratio To identify the actual hydride donor in Nikonov’s carbonyl hydrosilylation, we initially targeted the preparation of the postulated neutral ruthenium(II) monohydride 8 (Scheme 6, top) but its independent synthesis failed. However, we were able to access the ruthenium(IV) silyl dihydride 13a (Scheme 6, bottom)14 which can be regarded as the oxidative-addition complex of hydrosilane 5a added to the coordinatively unsaturated ruthenium(II) monohydride 12; 12, in turn, could be available from 8 by acetonitrile dissociation (Scheme 6, top).15 That silyl dihydride 13a is exactly what, in analogy to Brookhart’s case,5b we imagined to be the hydride source in a two-silicon cycle. Neutral 13a was prepared from cationic complex 3+[PF6]– (Scheme 6, bottom).

Scheme 6. Synthesis of Ruthenium(IV) Silyl Dihydride 13a Scheme 4. Control Experiments Excluding Hydrogenation of the Silyl Enol Ether Nikonov and co-workers had not included the formation of silyl enol ethers in their communication.8 We noticed that the hydrosilane-to-acetophenone ratio influences the reaction outcome (Scheme 5). The ratio of 30:70 gradually changed to 60:40 when increasing the amount of Me2PhSiH (5a) from 1.0 to 5.0 equivalents. This trend supports the notion that another hydrosilane molecule enhances the hydride transfer to silylcarboxonium ion 9a+ at the cost of its α deprotonation. A distinct isotope effect is seen with acetophenone-d3 (cf. Scheme 4, top); the deuterium-labeling slows down the α deprotonation of 9a+-d3 and shifts the ratio from 30:70 to 60:40 in favor of the silyl ether. With sterically more demanding (SiS)-5b, there was no dependence on the hydrosilane-toacetophenone ratio. We interpret this as additional evidence of the involvement of two silicon compounds in the reduction step, either preventing the anticipated activation of the neutral ruthenium(II) monohydride 8 or the hydride delivery to the silylcarboxonium ion for steric reasons.

With silyl dihydride 13a in hands, we performd a series of control experiments to prove or disprove whether 13a is a catalytically competent intermediate in the catalytic cycle (Scheme 7). To substantiate its hydride donor ability, 13a was reacted with freshly prepared silylcarboxonium ion 9c+ in a stoichiometric experiment (Scheme 7A; note that 13a and 9c+ contain different silyl groups). This reaction afforded the silyl ether 10c and ethylbenzene 14 in a complex mixture. This corroborates our hypothesis that the ruthenium(IV) silyl dihydride 13a does indeed release hydride. Conversely, the same complex 13a cannot promote the carbonyl hydrosilylation (Scheme 7B). Hydride abstraction from 13a by a silylcarboxonium ion is a prerequisite to (re)generate Nikonov’s catalytically active intermediates 4+ and 6a+. A clearer picture emerges when a hydride is abstracted from 13a by silylcarboxonium ion 9c+ in the presence of acetophenone (7) prior to the addition of hydrosilane 5a (Scheme 7C). 7 then fully converted into silyl ether 10a (cf. Scheme 7A) along with silyl ether 10a without the formation of either silyl enol ether 11a or 11c. This result leads to the following conclusions: It demonstrates that (1) acetonitrile-free 13a is part of the catalytic cycle and (2) the acetonitrile ligand introduced with catalyst 3+ is involved in the dehydrogenation pathway, i.e., the formation of the silyl enol ether by α deprotonation of the silylcarboxonium ion. To further verify the latter, we repeated control experiment 7C with addition of doubly hydrosilylated acetonitrile 15a (formed in the catalysis with 3+, see above). Indeed, formation of silyl ethers 10a and 10c was now accompanied by the formation of silyl enol ether 11a (Scheme 7D) in the expected ratio (cf. Scheme 5). Notably, the catalysis can also be initiated in the same way from silyl dihydride 13a by using the

ACS Paragon Plus Environment

ACS Catalysis

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

trityl cation as the hydride acceptor instead of silylcarboxonium ion 9c+ (not shown; see the Supporting Information for details). Finally, we were able to show that 13a does not undergo reductive elimination of hydrosilane 5a. A mixture of 13a and deuterium-labeled 5a-d1 did not equilibrate to 13a and 13a-d1 (Scheme 7E). A O

SiEt3 C6D5Cl r.t.

Ph Me [B(C6F5)4]– 9c+[B(C6F5)4]– B

Me 7

C

Ph

Me 7

D

Me 7

then Me 2PhSiH (5a, 1.0 equiv) 1,2-Cl 2C6D4 r.t., 40 min

then Me 2PhSiH (5a, 1.0 equiv) 1,2-Cl 2C6D4 r.t., 40 min

E

O

Ru Ph3P SiMe 2Ph H H 13a

Me

decomposition

H 10c

SiMe2Ph

O Ph

11a

Si

Me H 10a (Si = Me 2PhSi) 10c (Si = Et3Si)

SiMe2Ph

+

O

Si

+

Ph

Ph 11a (Si = Me 2PhSi) 11c (Si = Et3Si)

10a:10c:11a:11c = 95:5:0:0 a

O

Si

Me H 10a (Si = Me 2PhSi) 10c (Si = Et3Si) Ph

O

Si

+ Ph 11a (Si = Me 2PhSi) 11c (Si = Et3Si)

10a:10c:11a:11c = 17:3:80:0 a

– Me2PhSiH 5a

a

Ph

Me H 10a

Ph

13a (30 mol %) 9c+[B(C6F5)4]– (20 mol %) 15a (60 mol %)

O Ph

CD2Cl2 r.t., 24 h no reaction

O

13a (30 mol %) 9c+[B(C6F5)4]– (20 mol %)

O

Me

+

14 13a (2.7 mol %) Me 2PhSiH (5a, 1.0 equiv)

O Ph

+ Ph

um(II) monohydride 17. The hydride transfer from 13a to silylcarboxonium ion 9a+ closes the catalytic cycle, releasing silyl ether 10a and η2-σ hydrosilane adduct 16a+ with stabilization of L.

SiEt3

O

H H

13a (1.0 equiv)

Page 4 of 6

12

+ Me 2PhSiD 5a-d1

Ru Ph3P SiMe2Ph H D 13a-d1

Ratios determined by GLC analysis.

Scheme 7. Control Experiments with Ruthenium(IV) Silyl Dihydride 13a Combining the above experimental observations leads to a revised catalytic cycle for Nikonov’s carbonyl hydrosilylation (Scheme 8, counterclockwise right). As catalyst 3+ is able to hydrosilylate its own ligands12 and since the presence of acetonitrile is not essential, η2-σ hydrosilane adduct 16a+ is a reasonable starting point of the catalytic cycle (L = acetonitrile or donor). An SN2-type nucleophilic attack of the carbonyl oxygen atom of acetophenone (7) at the activated hydrosilane in 16a+ yields the silylcarboxonium ion 9a+ and the neutral ruthenium(II) monohydride 17. This step proceeds with inversion of the configuration at the silicon atom but subsequent racemization by pseudorotational processes is facile due formation of the pentacoordinate intermediate 18a+.5b The actual hydride donor, silyl dihydride 13a, is generated by oxidative addition of another molecule of hydrosilane 5a to rutheni-

Scheme 8. Revised Catalytic Cycle for the Hydrosilylation of Acetophenone with Nikonov’s Catalyst (Si = Me2PhSi, L = Acetonitrile or Donor) The dehydrogenation pathway leading to silyl enol ether 11a is also mechanistically understood (Scheme 8, clockwise left). Doubly hydrosilylated acetonitrile 15a, existing in twice the amount of catalyst 3+, acts as the base, converting the silylcarboxonium ion 9a+ into 11a by α deprotonation. The resulting ammonium ion 19a+ then protonates the dormant ruthenium(II) monohydride 17 in the presence of hydrosilane 5a with evolution of dihydrogen to arrive back at the η2-σ hydrosilane adduct 16a+.

CONCLUSIONS Nikonov’s carbonyl hydrosilylation is now another example of a one-silicon cycle revised to a two-silicon cycle.5b We think that there are several other hydrosilylation reactions where the initially formed (neutral) transition-metal hydride is not sufficiently hydridic to transfer its hydride but where that hydricity is boosted by the aid of another hydrosilane molecule. We are currently in quest of similar cases. Moreover, we have also revealed the interfering role of the acetonitrile ligand. Its doubly hydrosilylated derivative is an amine base that is responsible for competing dehydrogenative

ACS Paragon Plus Environment

Page 5 of 6

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Catalysis

silylation of acetophenone. Nikonov and co-workers had reported 30% yield for the hydrosilylation of acetophenone while not mentioning the formation of the corresponding silyl enol ether (see Table 1 in Reference 8). We were able to show that the dehydrogenation pathway can be completely silenced when using an acetonitrile-free catalyst. Hence, this mechanistic insight opens the door to the development of chemoselective catalysts in a targeted way.

ASSOCIATED CONTENT Supporting Information The Supporting Information is available free of charge on the ACS Publications website at http://pubs.acs.org. Experimental details, characterization and spectroscopic data (PDF)

AUTHOR INFORMATION Corresponding Author

(11) That η2-σ coordination is supposed not to involve interligand interactions between the hydrosilane and the acetonitrile ligand based on quantum-chemical calculations.8 (12) Unlike [Cp(iPr3P)Ru(acetonitrile)2]+[B(C6F5)4]–, 3+[B(C6F5)4]– employed by us was shown by Gutsulyak and Nikonov not to promote catalytic hydrosilylation of nitriles with Me2PhSiH (see the Supporting Information for Gutsulyak, D. V.; Nikonov, G. I. Angew. Chem., Int. Ed. 2010, 49, 7553–7556). However, [Cp(Ph3P)Ru(acetonitrile)2]+[B(C6F5)4]– did react in our hands in a stoichiometric reaction. (13) For additional examples, see: (a) Rendler, S.; Oestreich, M.; Butts, C. P.; Lloyd-Jones, G. C. J. Am. Chem. Soc. 2007, 129, 502– 503. (b) Klare, H. F. T.; Oestreich, M.; Ito, J.-i.; Nishiyama, H.; Ohki, Y.; Tatsumi, K. J. Am. Chem. Soc. 2011, 133, 3312–3315. (c) Fallon, T.; Oestreich, M. Angew. Chem., Int. Ed. 2015, 54, 12488–12491. (14) For the synthesis and crystallographic characterization of related ruthenium(IV) silyl dihydrides, see: (a) Osipov, A. L.; Gerdov, S. M.; Kuzmina, L. G.; Howard, J. A. K.; Nikonov, G. I. Organometallics 2005, 24, 587–602. (b) Gutsulyak, D. V.; Osipov, A. L.; Kuzmina, L. G.; Howard, J. A. K.; Nikonov, G. I. Dalton Trans. 2008, 6843–6850. (15) Campion, B. K.; Heyn, R. H.; Tilley, T. D. J. Chem. Soc., Chem. Commun. 1992, 1201–1203.

[email protected]

Notes The authors declare no competing financial interest.

ACKNOWLEDGMENT J.F. thanks the Fonds der Chemischen Industrie for a predoctoral fellowship (2016−2018), and M.O. is indebted to the Einstein Foundation (Berlin) for an endowed professorship. Additional support by the Cluster of Excellence Unifying Concepts in Catalysis of the Deutsche Forschungsgemeinschaft (EXC 314/2) is also acknowledged. We thank Dr. Sebastian Kemper (TU Berlin) for expert advice with the NMR measurements.

REFERENCES (1) Iglesias, M.; Fernández-Alvarez, F. J.; Oro, L. A. ChemCatChem 2014, 6, 2486–2489. (2) Parks, D. J.; Piers, W. E. J. Am. Chem. Soc. 1996, 118, 9440– 9441. (3) (a) Parks, D. J.; Blackwell, J. M.; Piers, W. E. J. Org. Chem. 2000, 65, 3090–3098. (b) Rendler, S.; Oestreich, M. Angew. Chem., Int. Ed. 2008, 47, 5997–6000. (c) Sakata, K.; Fujimoto, H. J. Org. Chem. 2013, 78, 12505–12512. (d) Houghton, A. Y.; Hurmalainen, J.; Mansikkamäki, A.; Piers, W. E.; Tuononen, H. M. Nat. Chem. 2014, 6, 983–988. (4) Park, S.; Brookhart, M. Organometallics 2010, 29, 6057–6064. (5) (a) Yang, J.; White, P. S.; Schauer, C. K.; Brookhart, M. Angew. Chem., Int. Ed. 2008, 47, 4141–4143. (b) Metsänen, T. T.; Hrobárik, P.; Klare, H. F. T.; Kaupp, M.; Oestreich, M. J. Am. Chem. Soc. 2014, 136, 6912–6915. (6) Robert, T.; Oestreich, M. Angew. Chem., Int. Ed. 2013, 52, 5216–5218. (7) (a) For an authoritative review of electrophilic Si–H bond activation, see: Lipke, M. C.; Liberman-Martin, A. L.; Tilley, T. D. Angew. Chem., Int. Ed. 2017, 56, 2260–2294. (b) For a comprehensive overview of Si–H bond activation by transition-metal complexes, see: Corey, J. Y. Chem. Rev. 2011, 111, 863–1071. (8) Gutsulyak, D. V.; Vyboishchikov, S. F.; Nikonov, G. I. J. Am. Chem. Soc. 2010, 132, 5950–5951. (9) Yang, Y.-F.; Chung, L. W.; Zhang, X.; Houk, K. N.; Wu, Y.-D. J. Org. Chem. 2014, 79, 8856–8864. (10) [Cp(iPr3P)Ru(acetonitrile)2]+[B(C6F5)4]– was originally used as the hydrosilylation catalyst.8 However, we found that 3+[B(C6F5)4]– with a triphenylphosphine ligand performs equally well in the hydrosilylation of acetophenone. We decided to continue working with 3+ because 1H NMR spectra are better resolved.

ACS Paragon Plus Environment

ACS Catalysis

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 6 of 6

6

ACS Paragon Plus Environment