Manganese-Catalyzed Asymmetric Hydrosilylation ... - ACS Publications

4 downloads 0 Views 668KB Size Report
Aug 18, 2017 - N. A.; Nerush, A.; Shimon, L. J. W.; Leitus, G.; Avram, L.; Ben-David,. Y.; Milstein, D. ... (11) (a) Cavanaugh, M. D.; Gregg, B. T.; Cutler, A. R..
This is an open access article published under an ACS AuthorChoice License, which permits copying and redistribution of the article or any adaptations for non-commercial purposes.

Article http://pubs.acs.org/journal/acsodf

Manganese-Catalyzed Asymmetric Hydrosilylation of Aryl Ketones Xiaochen Ma, Ziqing Zuo, Guixia Liu, and Zheng Huang* State Key Laboratory of Organometallic Chemistry, Shanghai Institute of Organic Chemistry, 345 Lingling Road, Shanghai 200032, China S Supporting Information *

ABSTRACT: We disclose the synthesis of a series of manganese complexes of chiral iminopyridine oxazoline ligands and their application in the first manganese-catalyzed asymmetric ketone hydrosilylations. The most sterically hindered manganese catalyst bearing two CH(Ph)2 groups at the 2,6-ortho positions of the imino aryl ring and a tBu group on the oxazoline ring furnishes the secondary alcohols in high enantioselectivities and yields.



INTRODUCTION The asymmetric hydrosilylation of ketones provides a convenient and efficient way to optically active secondary alcohols, which are valuable building blocks in organic synthesis.1 As a consequence, tremendous interest has been driven to the development of efficient catalysts for this transformation.2 Over the past several decades, a variety of noble-metal-based catalysts, such as Rh, Ir, and Pt, have been applied in the asymmetric hydrosilylation of ketones.3 Considering the high cost and intrinsic toxicity associated with the noble metals, it is desirable to search for cheap and base-metal-based alternatives. Recently, progress has been made in the development of asymmetric ketone hydrosilylations using nonprecious metals, such as Fe,4 Co,5 Cu,6 and Zn.7 Manganese is the third earth-abundant transition metal8 and indispensable in many biological systems.9 Because of the low cost and environmentally benign nature of the metal, manganese catalysis has attracted much attention during the past 5 years.10 Among them, manganese(I) carbonyl complexes11 and a tris(dipivaloylmethanato)Mn(III)12 could be utilized as the catalysts for the hydrosilylation of ketones. More recently, manganese complexes with salen-,13 carbene-,14 or bis(imino)pyridine-based ligands15 have proven to be highly effective for this transformation. Despite these advances, enantiomeric hydrosilylation of ketones catalyzed by manganese complex has remained unknown. Recently, this laboratory16 and the Lu group17 independently designed the so-called “iminopyridine oxazoline” (IPO) ligands and applied the corresponding cobalt and iron complexes for catalytic, enantioselective reduction events. In line with our interest in developing base-metal-catalyzed organic transformations, we describe herein the synthesis of manganese complexes ligated by chiral IPO ligands and their catalytic performance in asymmetric ketone hydrosilylations.

Scheme 1. Synthesis of (IPO)Mn Complexes (S)-3a−d

the imino aryl ring with anhydrous Mn(THF)2Cl2 formed the Mn(II) dichloride complexes [(S)-iPr2IPOR1]MnCl2 (R1 = Bn, (S)-3a; R1 = iPr, (S)-3b) in good yields. Following the same procedure, the sterically more demanding complexes (S)-3c and (S)-3d [(S)-(Ph2CH)2IPOR2]MnCl2 (R2 = iPr, (S)-3c; R2 = tBu, (S)-3d) bearing the CH(Ph)2 substituents at the 2,6-aryl positions were also produced in high yields. The single-crystal X-ray diffraction analysis of (S)-3d revealed a distorted squarepyramidal geometry around the Mn center (Figure 1). The catalytic performance of the manganese complex 3a was initially investigated in the asymmetric hydrosilylation of 4chloroacetophenone (4a) using several commercially available hydrosilanes (Table 1, entries 1−4). Upon activation with NaBHEt3 (2 mol %),18 3a (1 mol %) exhibits high activity with phenylsilane (PhSiH3) as the reductant at room temperature, furnishing the desired alcohol in high yield and moderate enantioselectivity after hydrolysis (entry 1). A secondary silane, Ph2SiH2, gave a slightly lower selectivity than PhSiH3 (entry 2).



RESULTS AND DISCUSSION This study started with the preparation of the chiral IPO manganese complexes (Scheme 1). Treatment of the IPO ligands containing two iPr groups at the 2,6-ortho positions of © 2017 American Chemical Society

Received: June 1, 2017 Accepted: July 31, 2017 Published: August 18, 2017 4688

DOI: 10.1021/acsomega.7b00713 ACS Omega 2017, 2, 4688−4692

ACS Omega

Article

solvents resulted in either decreased reaction rate or lower enantioselectivity. In addition, the concentration of the substrates had an important impact on the enantioselectivity, and the best result was obtained when the concentration of ketone was 0.1 M (for detailed optimization studies, see Table S1 in the Supporting Information). Using NaOtBu instead of NaBHEt3 could afford the desired product in a relatively lower yield and enantioselectivity (Table 1, entry 10). Finally, under optimized conditions, hydrosilylation using the sterically most demanding complex (S)-3d afforded high enantioselectivity (99.5:4.5 e.r.) and excellent isolated yield (98%) (Table 1, entry 9). Having the optimized conditions established, we evaluated the scope of the manganese-catalyzed asymmetric ketone hydrosilylation (Table 2). With the most sterically hindered IPO manganese complexes (S)-3d as the precatalyst, hydrosilylations of a variety of aryl alkyl ketones occurred smoothly at the ambient temperature with low catalyst loadings (1 mol %). In most cases, high isolated yields and good enantioselectivities were achieved. This method is effective with acetophenones bearing both electron-donating and electron-withdrawing groups. The substituent pattern on the aryl rings of acetophenones can have a significant influence on the activity as well as the enantiomeric selectivity. Substrates with chloro group and methoxy group at the ortho position barely formed the reduction products at 25 °C, but they could be transformed to the corresponding alcohols smoothly at 60 °C with good enantioselectivity (5i, 92:8 e.r.; 5l, 91.5:8.5 e.r.). In addition, although high yield and e.r. value were obtained for the metamethoxy-substituted substrate (5k), the para-methoxy-substituted substrate afforded the product in relatively low enantioselectivity (5j). However, for the chloro-substituted acetophenones, the position of the chloro group did not affect the enantioselectivity significantly. The substrates with chloro groups at the ortho, meta, or para position were reduced in comparable enantioselectivities (5a, 5h, and 5i). Remarkably, the bromo (5m), iodo (5n), and cyano (5r) group could also be tolerated, affording the desired products in moderate to high yields and enantioselectivities. The sterically demanding 1mesitylethanone underwent hydrosilylation successfully to form the product in 96% yield and 95:5 e.r. (5g). The reaction using 2-acetylnaphthalene as the substrate afforded the desired product in a high yield and moderate enantioselectivity (5s). Aryl alkyl ketones in which the alkyl group is larger than the methyl group were also explored in the asymmetric hydrosilylation (Table 2, 5t−y). Good enantioselectivity could be obtained when propyl phenyl ketone was employed as the substrate (5t, 94:6 e.r.). However, a further increase of the steric hindrance of the substituent at the benzylic position resulted in reduced enantioselectivity (5u, 69.5:30.5 e.r.). The reaction of the substrate bearing a cyclopropyl group was sluggish under standard conditions. Thus, a less sterically hindered IPO manganese complex (S)-3c was used as the precatalyst, which gave a satisfied yield and moderate enantioselectivity (5v, 89% yield, 71.5:28.5 e.r.). Cyclic aryl ketones resulted in moderate stereoselectivity (5w and 5x). The hydrosilylation of 1,1-diaryl ketones gave a low e.r. value (5y).

Figure 1. Solid-state structure of (S)-3d. Selected bond distances (Å) and angles (deg): Mn1−N1, 2.304(4); Mn1−N2, 2.175(3); Mn1−N3, 2.384(4); Mn1−Cl1, 2.3148(14); Mn1−Cl2, 2.3578(15); N1−Mn1− N2, 74.07(14); N2−Mn1−N3, 69.49(13); N1−Mn1−N3, 141.16(13); N2−Mn1−Cl1, 136.66(12); N2−Mn1−Cl2, 100.90(11); and Cl1−Mn1−Cl2, 121.94(6).

Table 1. Optimization of Reaction Conditionsa

entry

cat.

silane

yield (%)

e.r.

1 2 3 4 5b 6b 7b 8b 9b,c 10b,c,d

(S)-3a (S)-3a (S)-3a (S)-3a (S)-3a (S)-3b (S)-3c (S)-3d (S)-3d (S)-3d

PhSiH3 Ph2SiH2 (EtO)3SiH MD’M PhSiH3 PhSiH3 PhSiH3 PhSiH3 PhSiH3 PhSiH3

94 91 82 N/A 89 89 92 98 98 86

76.5:23.5 71.5:28.5 59.5:40.5 N/A 75:25 77.5:22.5 80.5:19.5 88:12 95.5:4.5 90:10

a Reaction conditions: 4a (0.5 mmol), silane (1 equiv), IPO manganese complex (1 mol %), and NaBEt3H (2 mol %) in toluene (1 mL) at 25 °C. Yields were determined by 1H NMR spectroscopy using mesitylene as an internal standard. The e.r. values were determined by HPLC analysis. b4b was used as the substrate. cToluene (5 mL) was used. dNaOtBu (2 mol %) was used instead of NaBHEt3.

The tertiary silane (EtO)3SiH is active for hydrosilylation but gives low enantioselectivity, whereas methylmethoxysilane (MD’M) did not deliver the hydrosilylation product (entries 3 and 4). The enantioselectivity is affected by using different silane sources, which indicates that silane might be involved in the enantio-determining transition state.4k Hence, PhSiH3 was chosen as the silane source for further studies. Next, manganese complexes (S)-3a−3d with different IPO ligands were evaluated in the hydrosilylation of ketone 4b (Table 1, entries 5−8). The reaction with (S)-3a or (S)-3b bearing iPr substituents at the 2,6-ortho position of the imino aryl ring provided high reactivity but moderate enantioselectivity (entries 5 and 6). Increasing the steric hindrance of the ligand by replacing the iPr groups in (S)-3b with two CH(Ph)2 units, [(S)-3c], led to a significantly improved enantioselectivity (entry 7, 80.5:19.5 e.r.). The substitution of tBu for iPr on the oxazoline moiety resulted in a further enhancement of the enantioselectivity (entry 8, 88:12 e.r.). Screening of the solvents indicated that toluene was the optimal medium, while the reactions with other common



CONCLUSIONS In summary, we have developed the first manganese-catalyzed asymmetric hydrosilylation of ketones using a chiral IPO manganese pincer complex as the precatalyst. This protocol provides an efficient way for the synthesis of optically active 4689

DOI: 10.1021/acsomega.7b00713 ACS Omega 2017, 2, 4688−4692

ACS Omega

Article

Spectra were recorded on an Agilent 400 MHz or a Varian Mercury 400 MHz instrument. 1H NMR chemical shifts were referenced to residual protio solvent peaks or tetramethylsilane signal (0 ppm), and 13C NMR chemical shifts were referenced to the solvent resonance. Data for 1H NMR were recorded as follows: chemical shift (δ, ppm), multiplicity (s = singlet, d = doublet, t = triplet, quint = quintuplet, sext = sextuplet, m = multiplet or unresolved, coupling constant(s) in Hz, integration). Data for 13C NMR were reported in terms of chemical shift (δ, ppm). Enantiomeric excesses were determined by highperformance liquid chromatography (HPLC) with a Dionex or Agilent chromatography [Phenomenex Lux 5u Cellulose-1 (0.46 × 25 cm), Phenomenex Lux 5u Cellulose-3 (0.46 × 25 cm), Phenomenex Lux 5u Cellulose-4 (0.46 × 25 cm), Daicel CHIRALPAK AD-H (0.46 × 25 cm), Daicel CHIRALPAK OD-H (0.46 × 25 cm), CHIRALPAK IC (0.46 × 25 cm), and Lux 5u Amylose-2 (0.46 × 25 cm)] in comparison with authentic racemic materials. Optical rotations were measured on a Rudolph Research Analytical Autopol I Polarimeter. Elemental analyses and high-resolution mass spectrometry were carried out by the Analytical Laboratory of Shanghai Institute of Organic Chemistry (CAS). Synthesis of Complex [(S)-Bn-IPO]MnCl2 [(S)-3a]. To a yellow solution of (S)-1a (260 mg, 0.66 mmol) in approximately 30 mL of THF, 179 mg (0.66 mmol) of MnCl2(THF)2 was added. The resulting mixture was stirred at 60 °C for 24 h. After that, it was filtered through a plug of Celite, and the volatiles were removed in vacuo. The residue was washed with toluene (3 mL*5) to afford product (S)-3a as an orange solid (390 mg, 96%). Anal. Calcd for C29H33Cl2MnN3O: C, 61.60; H, 5.88; N, 7.43. Found: C, 61.72; H, 5.75; N, 7.33. Synthesis of Complex [(S)-iPr-IPO]MnCl2 [(S)-3b]. This compound was prepared in a manner similar to (S)-3a with 178 mg (0.66 mmol) of (S)-1b, 179 mg of MnCl2(THF)2 (0.66 mmol), and approximately 30 mL of THF. This procedure yielded 330 mg (96%) of an orange solid identified as (S)-3b. Anal. Calcd for C46H43Cl2MnN3O: C, 70.86; H, 5.56; N, 5.39. Found: C, 70.82; H, 5.55; N, 5.64. Synthesis of Complex [Dibenzhydryl-(S)-iPr-IPO]MnCl2 [(S)-3c]. This compound was prepared in a manner similar to (S)-3a with 430 mg (0.66 mmol) of (S)-2a, 178 mg of MnCl2(THF)2 (0.66 mmol), and approximately 20 mL of THF. This procedure yielded 460 mg (90%) of an orange solid identified as (S)-3c. Anal. Calcd for C47H45Cl2MnN3O: C, 71.12; H, 5.71; N, 5.29. Found: C, 70.71; H, 5.56; N, 5.19. Synthesis of Complex [Dibenzhydryl-(S)-tBu-IPO]MnCl2 [(S)-3d]. This compound was prepared in a manner similar to (S)-3a with 162 mg (0.24 mmol) of (S)-2b, 65 mg of MnCl2(THF)2 (0.24 mmol), and approximately 20 mL of THF. This procedure yielded 178 mg (90%) of an orange solid identified as (S)-3d. Anal. Calcd for C47H45Cl2MnN3O: C, 71.12; H, 5.71; N, 5.29. Found: C, 70.71; H, 5.56; N, 5.19. General Procedure for the Hydrosilylation of Ketones with the Manganese Complex. In a nitrogen-filled glovebox, to a solution of (S)-3d (0.005 mmol, 4.0 mg) in 2 mL of toluene, a solution (1.0 M in THF) of NaBHEt3 (10 μL, 0.01 mmol) was slowly added at 25 °C. After stirring for 1 min, PhSiH3 (54.1 mg, 0.5 mmol, 1 equiv) and the ketone substrate 4 (0.5 mmol) were sequentially added. The reaction mixture was stirred at 25 °C for 3 h and then was quenched by exposing the solution to air. Then, MeOH (1.5 mL) and 10% NaOH (2 mL) were added with vigorous stirring for 10 h. The resulting

Table 2. Mn-Catalyzed Asymmetric Hydrosilylation of Various Aryl Ketonesa

a

Reaction conditions: 4 (0.5 mmol), PhSiH3 (0.5 mmol), (S)-3d (1 mol %), NaBEt3H (2 mol %), and toluene (2.0 mL) at 25 °C. Isolated yields. The e.r. values were determined by HPLC analysis. The absolute configurations were assigned by comparison with the reported optical rotations. bReaction was conducted at 60 °C. c(S)3c was used as the precatalyst.

secondary alcohols under mild conditions with low catalyst loadings. Further studies on the development of new asymmetric transformations using the IPO manganese catalysts are currently underway in our laboratory.



EXPERIMENTAL SECTION General Description. Unless otherwise noted, all reagents were purchased from commercial suppliers and used without further purification. All manipulations were carried out using standard Schlenk, high-vacuum, and glovebox techniques. Tetrahydrofuran (THF) and toluene were distilled from sodium benzophenone ketyl prior to use. 4690

DOI: 10.1021/acsomega.7b00713 ACS Omega 2017, 2, 4688−4692

ACS Omega

Article

Am. Chem. Soc. 2015, 137, 2456. (l) Bleith, T.; Gade, L. H. J. Am. Chem. Soc. 2016, 138, 4972. (5) For representative examples, see: (a) Chen, X.; Lu, Z. Org. Lett. 2016, 18, 4658. (b) Sauer, D. C.; Wadepohl, H.; Gade, L. H. Inorg. Chem. 2012, 51, 12948. (c) Yu, F.; Zhang, X.-C.; Wu, F.-F.; Zhou, J.N.; Fang, W.; Wu, J.; Chan, A. S. C. Org. Biomol. Chem. 2011, 9, 5652. (d) Inagaki, T.; Phong, L. T.; Furuta, A.; Ito, J.-i.; Nishiyama, H. Chem.Eur. J. 2010, 16, 3090. (6) For representative examples, see: (a) Sirol, S.; Courmarcel, J.; Mostefai, N.; Riant, O. Org. Lett. 2001, 3, 4111. (b) Lipshutz, B. H.; Noson, K.; Chrisman, W.; Lower, A. J. Am. Chem. Soc. 2003, 125, 8779. (c) Lipshutz, B. H.; Lower, A.; Noson, K. Org. Lett. 2002, 4, 4045. (d) Lee, D.-w.; Yun, J. Tetrahedron Lett. 2004, 45, 5415. (e) Zhang, X.-C.; Wu, Y.; Yu, F.-F.; Wu, F.; Wu, J.; Chan, A. S. C. Chem.Eur. J. 2009, 15, 5888. (f) Junge, K.; Wendt, B.; Addis, D.; Zhou, S.; Das, S.; Beller, M. Chem.Eur. J. 2010, 16, 68. (g) Issenhuth, J.-T.; Dagorne, S.; Bellemin-Laponnaz, S. C. R. Chim. 2010, 13, 353. (h) Yu, F.; Zhou, J.-N.; Zhang, X.-C.; Sui, Y.-Z.; Wu, F.-F.; Xie, L.-J.; Chan, A. S. C.; Wu, J. Chem.Eur. J. 2011, 17, 14234. (i) Sui, Y.-Z.; Zhang, X.-C.; Wu, J.-W.; Li, S.; Zhou, J.-N.; Li, M.; Fang, W.; Chan, A. S. C.; Wu, J. Chem.Eur. J. 2012, 18, 7486. (j) Qi, S.-B.; Li, M.; Li, S.; Zhou, J.-N.; Wu, J.-W.; Yu, F.; Zhang, X.-C.; Chan, A. S. C.; Wu, J. Org. Biomol. Chem. 2013, 11, 929. (7) For representative examples, see: (a) Mimoun, H. J. Org. Chem. 1999, 64, 2582. (b) Bette, V.; Mortreux, A.; Lehmann, C. W.; Carpentier, J.-F. Chem. Commun. 2003, 332. (c) Mastranzo, V. M.; Quintero, L.; Anaya de Parrodi, C.; Juaristi, E.; Walsh, P. J. Tetrahedron 2004, 60, 1781. (d) Bette, V.; Mortreux, A.; Savoia, D.; Carpentier, J.F. Tetrahedron 2004, 60, 2837. (e) Gérard, S.; Pressel, Y.; Riant, O. Tetrahedron: Asymmetry 2005, 16, 1889. (f) Bandini, M.; Melucci, M.; Piccinelli, F.; Sinisi, R.; Tommasi, S.; Umani-Ronchi, A. Chem. Commun. 2007, 4519. (g) Gajewy, J.; Kwit, M.; Gawroǹski, J. Adv. Synth. Catal. 2009, 351, 1055. (h) Nishiyama, H.; Inagaki, T.; Yamada, Y.; Phong, L. T.; Furuta, A.; Ito, J.-i. Synlett 2009, 253. (i) Liu, S.; Peng, J.; Yang, H.; Bai, Y.; Li, J.; Lai, G. Tetrahedron 2012, 68, 1371. (j) Łowicki, D.; Bezłada, A.; Mlynarski, J. Adv. Synth. Catal. 2014, 356, 591. (k) Szewczyk, M.; Stanek, F.; Bezłada, A.; Mlynarski, J. Adv. Synth. Catal. 2015, 357, 3727. Szewczyk, M.; Bezłada, A.; Mlynarski, J. ChemCatChem 2016, 8, 3575. (8) Ritter, S. K. Chem. Eng. News 2012, 90, 12. (9) (a) Ferreira, K. N.; Iverson, T. M.; Maghlaoui, K.; Barber, J.; Iwata, S. Science 2004, 303, 1831. Wu, A. J.; Penner-Hahn, J. E.; Pecoraro, V. L. Chem. Rev. 2004, 104, 903. (10) For recent reviews, see: (a) Shi, Z.-H.; Li, N.-G.; Shi, Q.-P.; Tang, Y.-P.; Tang, H.; Shen, M.-Z.; Duan, J.-A. Curr. Org. Synth. 2014, 11, 204. (b) Liu, W.; Ackermann, L. ACS Catal. 2016, 6, 3743. For selected recent examples, see: (c) Zhou, B.; Chen, H.; Wang, C. J. Am. Chem. Soc. 2013, 135, 1264. (d) He, R.; Huang, Z.-T.; Zheng, Q.-Y.; Wang, C. Angew. Chem., Int. Ed. 2014, 53, 4950. (e) Elangovan, S.; Topf, C.; Fischer, S.; Jiao, H.; Spannenberg, A.; Baumann, W.; Ludwig, R.; Junge, K.; Beller, M. J. Am. Chem. Soc. 2016, 138, 8809. (f) Mukherjee, A.; Nerush, A.; Leitus, G.; Shimon, L. J. W.; David, Y. B.; Jalapa, N. A. E.; Milstein, D. J. Am. Chem. Soc. 2016, 138, 4298. (g) Zhang, G.; Zeng, H.; Wu, J.; Yin, Z.; Zheng, S.; Fettinger, J. C. Angew. Chem., Int. Ed. 2016, 55, 14369. (h) Elangovan, S.; Garbe, M.; Jiao, H.; Spannenberg, A.; Junge, K.; Beller, M. Angew. Chem., Int. Ed. 2016, 55, 15364. (i) Ruan, Z.; Sauermann, N.; Manoni, E.; Ackermann, L. Angew. Chem., Int. Ed. 2016, 56, 3172. (j) Atack, T. C.; Cook, S. P. J. Am. Chem. Soc. 2016, 138, 6139. (k) Espinosa-Jalapa, N. A.; Nerush, A.; Shimon, L. J. W.; Leitus, G.; Avram, L.; Ben-David, Y.; Milstein, D. Chem.Eur. J. 2017, 23, 5934. (l) Papa, V.; CabreroAntonino, J. R.; Alberico, E.; Spanneberg, A.; Junge, K.; Junge, H.; Beller, M. Chem. Sci. 2017, 8, 3576. (11) (a) Cavanaugh, M. D.; Gregg, B. T.; Cutler, A. R. Organometallics 1996, 15, 2764. (b) Son, S. U.; Paik, S.-J.; Lee, I. S.; Lee, Y.-A.; Chung, Y. K.; Seok, W. K.; Lee, H. N. Organometallics 1999, 18, 4114. (c) Son, S. U.; Paik, S.-J.; Chung, Y. K. J. Mol. Catal. A: Chem. 2000, 151, 87. (d) Hilal, H. S.; Suleiman, M. A.; Jondi, W. J.; Khalaf, S.; Masoud, M. M. J. Mol. Catal. A: Chem. 1999, 144, 47.

solution was extracted with EtOAc, washed with brine (15 mL), and dried over anhydrous Na2SO4. After filtration and evaporation of the solvent, the residue was purified by flash column chromatography with EtOAc/petroleum ether to give the desired product.



ASSOCIATED CONTENT

S Supporting Information *

The Supporting Information is available free of charge on the ACS Publications website at DOI: 10.1021/acsomega.7b00713. Experimental procedures and product characterization (PDF)



AUTHOR INFORMATION

Corresponding Author

*E-mail: [email protected] (Z.H.). ORCID

Zheng Huang: 0000-0001-7524-098X Notes

The authors declare no competing financial interest.



ACKNOWLEDGMENTS This work was supported by the MOST of China (2015CB856600 and 2016YFA0202900), the National Natural Science Foundation of China (21432011, 21422209, 21421091, and 21572255), and the Chinese Academy of Science (XDB20000000).



REFERENCES

(1) (a) Noyori, R. Asymmetric Catalysis in Organic Synthesis; Wiley: New York, 1994. (b) Noyori, R.; Ohkuma, T. Angew. Chem., Int. Ed. 2001, 40, 40. (c) Ikariya, T.; Blacker, A. J. Acc. Chem. Res. 2007, 40, 1300. (2) (a) Brunner, H.; Nishiyama, H.; Itoh, K. In Catalytic Asymmetric Synthesis; Ojima, I., Ed.; Wiley-VCH: New York, 1993; Chapter 6. (b) Nishiyama, H. In Comprehensive Asymmetric Catalysis; Jacobsen, E. N., Pfaltz, A., Yamamoto, H., Eds.; Springer: New York, 1999; Chapter 6.3. (3) For representative examples, see: (a) Nishiyama, H.; Sakaguchi, H.; Nakamura, T.; Horihata, M.; Kondo, M.; Itoh, K. Organometallics 1989, 8, 846. (b) Sawamura, M.; Kuwano, R.; Ito, Y. Angew. Chem., Int. Ed. 1994, 33, 111. (c) Tao, B.; Fu, G. C. Angew. Chem., Int. Ed. 2002, 41, 3892. (d) Gade, L. H.; César, V.; Bellemin-Laponnaz, S. Angew. Chem., Int. Ed. 2004, 43, 1014. (e) Zhu, G.; Terry, M.; Zhang, X. J. Organomet. Chem. 1997, 547, 97. (f) Nishibayashi, Y.; Takei, I.; Uemura, S.; Hidai, M. Organometallics 1998, 17, 3420. (g) Chianese, A. R.; Crabtree, R. H. Organometallics 2005, 24, 4432. (h) Malacea, R.; Poli, R.; Manoury, E. Coord. Chem. Rev. 2010, 254, 729. (i) DíezGonzález, S.; Nolan, S. P. Org. Prep. Proced. Int. 2007, 39, 523. (j) Hydrosilylation: A Comprehensive Review on Recent Advances; Marciniec, B., Ed.; Springer: Netherlands, 2009. (4) For recent examples, see: (a) Nishiyama, H.; Furuta, A. Chem. Commun. 2007, 760. (b) Shaikh, N. S.; Enthaler, S.; Junge, K.; Beller, M. Angew. Chem., Int. Ed. 2008, 47, 2497. (c) Langlotz, B. K.; Wadepohl, H.; Gade, L. H. Angew. Chem., Int. Ed. 2008, 47, 4670. (d) Gaillard, S.; Renaud, J.-L. ChemSusChem 2008, 1, 505. (e) Tondreau, A. M.; Darmon, J. M.; Wile, B. M.; Floyd, S. K.; Lobkovsky, E.; Chirik, P. J. Organometallics 2009, 28, 3928. (f) Morris, R. H. Chem. Soc. Rev. 2009, 38, 2282. (g) Addis, D.; Shaikh, N.; Zhou, S.; Das, S.; Junge, K.; Beller, M. Chem.Asian J. 2010, 5, 1687. (h) Inagaki, T.; Phong, L. T.; Furuta, A.; Ito, J.-I.; Nishiyama, H. Chem.Eur. J. 2010, 16, 3090. (i) Junge, K.; Schröder, K.; Beller, M. Chem. Commun. 2011, 47, 4849. (j) Flückiger, M.; Togni, A. Eur. J. Org. Chem. 2011, 4353. (k) Bleith, T.; Wadepohl, H.; Gade, L. H. J. 4691

DOI: 10.1021/acsomega.7b00713 ACS Omega 2017, 2, 4688−4692

ACS Omega

Article

(e) Cavanaugh, M.; Gregg, B. T.; Cutler, A. R. Organometallics 1996, 15, 2764. (f) Son, S. U.; Paik, S.-J.; Lee, I. S.; Lee, Y.-A.; Chung, Y. K.; Seok, W. K.; Lee, H. N. Organometallics 1999, 18, 4114. (g) Son, S. U.; Paik, S.-J.; Chung, Y. K. J. Mol. Catal. A: Chem. 2000, 151, 87. (12) Magnus, P.; Fielding, M. R. Tetrahedron Lett. 2001, 42, 6633. (13) (a) Chidara, V. K.; Du, G. Organometallics 2013, 32, 5034. (b) Vijjamarri, S.; Chidara, V. K.; Du, G. ACS Omega 2017, 2, 582. (14) (a) Valyaev, D. A.; Wei, D.; Elangovan, S.; Cavailles, M.; Dorcet, V.; Sortais, J.-B.; Darcel, C.; Lugan, N. Organometallics 2016, 35, 4090. (b) Zheng, J.; Elangovan, S.; Valyaev, D. A.; Brousses, R.; César, V.; Sortais, J.-B.; Darcel, C.; Lugan, N.; Lavigne, G. Adv. Synth. Catal. 2014, 356, 1093. (15) (a) Mukhopadhyay, T. K.; Flores, M.; Groy, T. L.; Trovitch, R. J. J. Am. Chem. Soc. 2014, 136, 882. (b) Ghosh, C.; Mukhopadhyay, T. K.; Flores, M.; Groy, T. L.; Trovitch, R. J. Inorg. Chem. 2015, 54, 10398. (c) Mukhopadhyay, T. K.; Rock, C. L.; Hong, M.; Ashley, D. C.; Groy, T. L.; Baik, M.-H.; Trovitch, R. J. J. Am. Chem. Soc. 2017, 139, 4901. (16) (a) Zhang, L.; Zuo, Z.; Wan, X.; Huang, Z. J. Am. Chem. Soc. 2014, 136, 15501. (b) Zuo, Z.; Zhang, L.; Leng, X.; Huang, Z. Chem. Commun. 2015, 51, 5073. (17) (a) Chen, J.; Xi, T.; Lu, Z. Org. Lett. 2014, 16, 6452. (b) Chen, J.; Xi, T.; Ren, X.; Cheng, B.; Guo, J.; Lu, Z. Org. Chem. Front. 2014, 1, 1306. (c) Chen, J.; Cheng, B.; Cao, M.; Lu, Z. Angew. Chem., Int. Ed. 2015, 54, 4661. (d) Guo, J.; Chen, J.; Lu, Z. Chem. Commun. 2015, 51, 5725. (e) Guo, J.; Lu, Z. Angew. Chem., Int. Ed. 2016, 55, 10835. (18) NaBHEt3 can react with the precatalyst 3 to generate the catalytically active manganese monohydride species in situ. A plausible mechanism is proposed on the basis of a precedent of Mn-catalyzed ketone hydrosilylation described by Trovitch and co-workers (see ref 15c and the proposed mechanism in Supporting Information).

4692

DOI: 10.1021/acsomega.7b00713 ACS Omega 2017, 2, 4688−4692