Copper(II)-Catalyzed Selective Hydroboration of Ketones and Aldehydes

2 days ago - A novel nonanuclear copper(II) complex obtained by a facile one-pot self-assembly was found to catalyze the hydroboration of ketones and ...
4 downloads 0 Views 1MB Size
Letter pubs.acs.org/OrgLett

Cite This: Org. Lett. XXXX, XXX, XXX−XXX

Copper(II)-Catalyzed Selective Hydroboration of Ketones and Aldehydes Haisu Zeng,†,‡ Jing Wu,†,‡ Sihan Li,†,‡ Christina Hui,† Anita Ta,† Shu-Yuan Cheng,*,† Shengping Zheng,*,‡ and Guoqi Zhang*,† †

Downloaded via IOWA STATE UNIV on January 9, 2019 at 22:45:19 (UTC). See https://pubs.acs.org/sharingguidelines for options on how to legitimately share published articles.

Department of Sciences, John Jay College and Ph.D. Program in Chemistry, The Graduate Center of the City University of New York, New York, New York 10019, United States ‡ Department of Chemistry, Hunter College, The City University of New York, New York, New York 10065, United States S Supporting Information *

ABSTRACT: A novel nonanuclear copper(II) complex obtained by a facile one-pot self-assembly was found to catalyze the hydroboration of ketones and aldehydes with the absence of an activator under mild, solvent-free conditions. The catalyst is air- and moisture-stable, displaying high efficiency (1980 h−1 turnover frequency, TOF) and chemoselectivity on aldehydes over ketones and ketones over imines. This represents a rare example of divalent copper catalyst for the hydroboration of carbonyls.

using a CuI−carbene catalyst ((IPr)CuOtBu) with high efficiency in C6D6.12 Although this catalyst was tolerant of some reducible functional groups, it was unselective for ketone vs aldehyde hydroboration (Scheme 1). In addition, CuI− carbene catalysts often suffer from poor air and moisture stability and high expenses in ligand synthesis. In this work, we wish to report a novel, air- and moisturestable nonanuclear copper(II) complex assembled readily from

I

n sustainable chemistry, an urgency for replacement of precious metals with Earth-abundant elements toward fundamental catalytic processes has prompted enormous research efforts leading to observation of effective nonprecious metal catalysts.1 Recent progress in industrially important catalytic processes including (transfer) hydrogenation, dehydrogenation, and hydrofunctionalization using well-defined iron and cobalt molecular catalysts has shown great promise for the future utilization of Earth-abundant metals in important organic transformations.2,3 Catalytic hydroboration represents a highly valuable approach toward synthesizing functionalized organoborates that play a key role in organic synthesis.4,5 A variety of base metal (Fe, Co, Mn, and Cu) catalysts have been recently developed to carry out the hydroboration of alkenes and alkynes with pinacolborane (HBpin).5,6 Hydroboration of carbonyl compounds with HBpin provides an alternative route to access alcohols. The introduction of a Bpin group can also serve as versatile directing and protecting group in synthetic chemistry.7 To date, the hydroboration of carbonyls has been achieved by various metal catalysts including transition metal (Ti, Mo, Ru, Co, Fe, Cu, Ni, and Zn), main group (Li, Mg, Ca, Al, Ga, Ge, Sn, and P), and lanthanide catalysts.8 In addition, several heterogeneous catalysts based on Zr, Co, and Fe coordination frameworks are also known.9 In some examples, good chemoselectivity between ketones and aldehydes was also reported.7,9 It was noted that although monovalent copper catalysts based on phosphine or carbene ligands have been extensively exploited for the regio- and/or enantioselective hydroboration of alkene and alkynes,10 copper-catalyzed hydroboration of carbonyls is surprisingly underexplored.11 The only copper(I)-catalyzed hydroboration of carbonyls was reported by the Mankad group © XXXX American Chemical Society

Scheme 1. Copper-Catalyzed Hydroboration of Ketones and Aldehydes

Received: November 8, 2018

A

DOI: 10.1021/acs.orglett.8b03583 Org. Lett. XXXX, XXX, XXX−XXX

Letter

Organic Letters

self-assembly process, we found that when 4-tert-butyl-2,6diformylphenol was utilized in the one-pot reaction (ratio of aldehyde/amine/Cu = 1:2:6), green needle-like crystals of a novel nonanuclear CuII complex, namely (R)-Cu9 (2), were afforded in high yield (91%) after slow evaporation of solvents in air within 6 days (Scheme 2, see the SI). Compound 2 was characterized by IR spectroscopy, elemental analysis, X-ray crystallography, and X-ray photoelectron spectroscopy (XPS) (see the SI). Microanalytic data of vacuum-dried bulk sample of 2 are well consistent with a formula of Cu9(L)2(CH3COO−)8(CH3OH)2(μ3-OH)4 containing a Cu9 core (where L represents the fully deprotonated bis-Schiff base ligand formed in situ). The oxidation states of all copper centers are confirmed to be +2 by XPS analysis, in agreement with the charge balance in the formula. The IR spectrum showed a strong absorption at ∼1600 cm−1, assigned to the imine bonds in 2. The in vitro cytotoxicity measurements indicate that 2 is moderately toxic toward human breast cancer (MCF-7) and human leukemia (K562) cells, displaying significantly lower LD50 values than the tetranuclear 1 against both cells (see the SI). Single-crystal structural analysis of 2 confirmed the formation of a nonanuclear supramolecular complex composed of two Schiff base ligands and nine CuII centers (Figure 1a). Compound 2 crystallizes in the monoclinic space group P21, and two independent cluster molecules are included in the asymmetric unit, except for a number of crystalline solvent molecules. One of the Cu9 cluster structure with an N,Ocoordination environment is depicted in Figure 1. In each molecule, there contain nine CuII centers surrounded by two

Schiff base ligands and its extremely high activity in catalytic hydroboration of carbonyls without the use of extra activators.13 Notably, very high TOFs exceeding 1980 h−1 and excellent chemoselectivity for aldehydes over ketones were revealed, in contrast to the known CuI−carbene catalyst. This represents a rare example of a divalent copper complex that catalyzes hydroboration of carbonyl compounds.8 In search of readily available, air-stable CuII complexes for catalysis, we have previously reported the one-pot synthesis of a tetranuclear CuII cluster, namely (R)-Cu4 (1), resulting from the mixture of 3,5-di-tert-butyl-2-hydroxybenzaldehyde, (R)-2amino-2-phenylethanol, and Cu(OAc)2·2H2O in a 1:1:1 ratio in a CH2Cl2−MeOH solution and explored its catalytic activity in aerobic alcohol oxidation (Scheme 2).14 We were curious Scheme 2. One-Pot Synthesis of Complexes (R)-Cu4 (1) and (R)-Cu9 (2)

whether such a multinuclear CuII complex would be effective for reduction catalysis, such as hydroboration of unsaturated bonds. Hence, we examined the catalytic hydroboration of acetophenone with HBpin by using 1 as a catalyst in THF; to our delight, 30 turnover numbers (TONs) were found after 16 h (entry 1, Table 1). Encouraged by the moderate catalytic turnover, we sought to modify the structure of the CuII cluster complex and improve the catalytic efficiency. Interestingly, by varying the starting aldehyde in the CuII-mediated Schiff base Table 1. Reactivity Test of Copper(II)-Catalyzed Hydroboration of Acetophenone with HBpina

entry

catalyst

loading (mol %)

time (h)

solvent

yieldb (%)

1 2 3 4 5 6 7 8 9 10 11c

1 2 Cu(OAc)2 none 2 2 2 2 2 2 2

1.0 1.0 5.0 − 1.0 1.0 1.0 1.0 0.05 0.001 0.05

16 16 16 16 16 16 16 1 1 24 16

THF THF THF THF Et2O toluene pentane neat neat neat neat

30 81 0 0 75 60 62 >99 99 35 20

Figure 1. (a) ORTEP representation of 2 plotted at 50% thermal ellipsoid probability level, (b) ball−stick model of the Cu9 core showing the coordination environment, and (c) side view of the cluster molecule. Crystalline solvent molecules and H-atoms are omitted for clarity.

a

Conditions: acetophenone (1.0 mmol), HBpin (1.2 mmol), catalyst (indicated loading amount), and solvent (1 mL), rt, 1−24 h, N2. b Determined by GC analysis with hexamethylbenzene as an internal standard. cReaction run in the air. B

DOI: 10.1021/acs.orglett.8b03583 Org. Lett. XXXX, XXX, XXX−XXX

Letter

Organic Letters Table 2. Substrate Scope of Copper(II)-Catalyzed Ketone Hydroborationa

a

Conditions: ketone (1.0 mmol), HBpin (1.2 mmol), and 2 (0.05 mol %), neat, rt, 1 h, N2. bYields of hydroborated products (3) were determined by GC−MS analysis. Isolated yields of alcohols (4) are shown in parentheses. cReaction run at a 10 mmol scale. dReaction run for 16 h. e0.5 mol % of 2 was used.

fully deprotonated bis-Schiff base ligands, eight CH3COO−, and two CH3OH groups and four bridging −OH groups (through μ3-O coordination) to form a {Cu9N4O25} core, where all CuII centers are bridged by O atoms (Figure 1b). Interestingly, a side view of one independent molecule reveals a saddle-like shape and accessible CuII sites are available, indicating its potential in activating substrates in catalysis (vide infra). Next, complex 2 was examined for catalytic hydroboration of acetophenone with HBpin under the same conditions used for 1, without the use of an activator. The results for condition screening are summarized in Table 1. It was delightful to find that 2 (1 mol %) led to much more effective hydroboration of acetophenone than 1 without the presence of any additives, and 81% GC yield of boronate ester 3a was detected (entry 2, Table 1). In contrast, the control experiments show that Cu(OAc)2 alone (5 mol %) does not catalyze this reaction, and no product was observed in the absence of any catalysts (entries 3 and 4, Table 1). These results indicate the unique role of the nonanuclear CuII cluster in enabling catalytic reactivity. We further tested other solvents and solvent-free conditions for the 2-catalyzed hydroboration. The results reveal that the reaction proceeded most efficiently under neat conditions (entry 8, Table 1). To this end, quantitative yield was observed for 3a within 1 h under solvent-free conditions. It was further found that even when the the catalyst loading was lowered to 0.05 mol % the reaction could be completed in 1 h,

achieving a TOF of 1980 h−1 (entry 9, Table 1). The reaction was also tested with an extremely low loading of catalyst (0.001 mol %) under neat conditions, and 35% yield of 3a was detected by using GC after 24 h, corresponding to a TON of 35000 (entry 10, Table 1), which is comparable to the most effective metal catalysts including precious metals for HBpinbased reduction.2−8 Finally, it was noted that when the reaction was performed open to the air, a very low yield of 3a was observed (entry 11, Table 1). The applicability of 2-catalyzed hydroboration was then studied by using a variety of functionalized carbonyl compounds. First, a series of aryl and aliphatic ketones were employed under the optimized conditions (0.05 mol % of 2, neat). The results are shown in Table 2. The hydroborated products (3) were analyzed by GC−MS, and the subsequent purification by column chromatography on silica led to hydrolysis of boronate esters, affording the corresponding alcohols (4).5 It is worth mentioning that the current copper(II) precatalyst promoted the hydroboration of acetophenone smoothly on a gram scale (entry 1, Table 2). Differently substituted acetophenones with groups including methyl, halo, and methoxy (electron-donating) all proceeded well, affording the hydroborated products in high yields (entries 2−6, Table 2). However, acetophenones bearing 4trifluoromethyl and 4-nitro (electron-withdrawing) groups were less active substrates, with moderate yields of the desired products being isolated even after elongated reaction time C

DOI: 10.1021/acs.orglett.8b03583 Org. Lett. XXXX, XXX, XXX−XXX

Letter

Organic Letters (entries 7 and 8, Table 2). Isobutyrophenone, cyclopropyl phenyl ketone, and benzophenone were suitable substrates for the hydroboration, and the corresponding alcohols were isolated in appreciable yields (entries 9−11, Table 2). The reaction was also tolerant of cyclic or acyclic aliphatic ketones, affording boronate esters 3l−n with moderate to good yields when 0.5 mol % of catalyst was used (entries 12−14, Table 2). Interestingly, the heteroatomic aryl ketone 2-acetylpyrazine was fully hydroborated with high isolated yield of alcohol 4o, whereas 2- or 3-acetylpyridine and α,β-unsaturated ketone were challenging substrates under standard conditions (entries 15−18, Table 2). We also examined a variety of aldehyde substrates under the optimized conditions (Table 3). It was found that

benzaldehyde and functionalized benzaldehyde were suitable for copper(II)-catalyzed hydroboration with HBpin. Halogenated benzaldehydes (4-chloro- and pentafluoro-) are suitable substrates, and the corresponding primary alcohols were obtained in 81 and 90% yields, respectively (entries 2 and 3, Table 3). While benzaldehyde bearing an electron-donating group is a less active substrate affording only a moderate yield of alcohol, benzaldehydes containing strongly electron-withdrawing groups (−CF3 and −NO2) were converted to the corresponding alcohols in high yields (entries 4−6, Table 3). Ester-functionalized aldehyde was selectively hydroborated in a reasonable yield (entry 7, Table 3). In addition, furfural showed complete conversion, although 2-pyridinecarboxaldehyde was almost inactive (entries 8 and 9, Table 3). transCinnamaldehyde containing both CC and CO bonds was selectively hydroborated on the CO (entry 10, Table 3), giving the desired allylic alcohol in 80% yield. Finally, two aliphatic aldehydes were tested, and high conversions to boronate esters were found based on GC analysis (entries 11 and 12, Table 3). Chemoselectivity is a critical and challenging issue in developing practical methodology for carbonyl hydroboration.7 Previously, the Mankad’s CuI−carbene catalyst was found to show poor selectivity toward ketone vs aldehyde hydroboration.12 Our CuII catalyst was therefore examined for the chemoselective hydroboration (Scheme 3). Thus, an equi-

Table 3. Substrate Scope of Copper(II)-Catalyzed Aldehyde Hydroborationa

Scheme 3. Intermolecular and Intramolecular Chemoselective Hydroboration of Different Functionalities

molar mixture of acetophenone and benzaldehyde was first tested under standard conditions, and the quantitative hydroboration of benzaldehyde was observed, with acetophenone being fully recovered. Likewise, the same selectivity was observed in the competing hydroboration between 1-octanal and 2-heptanone. In addition, good selectivity on the ketone hydroboration over imine was also confirmed. Toward this end, the intramolecular competing hydroboration of ketone vs aldehyde was conducted using 4-acetylbenzaldehyde and selective hydroboration occurred only on the aldehyde part, offering alcohol 7 with 83% isolated yield. These results indicate excellent chemoselectivity of 2 toward aldehyde vs ketone and ketone vs imine hydroboration. To investigate whether 2-catalyzed hydroboration proceeded through a homogeneous or heterogeneous mechanism, we conducted a mercury poisoning experiment (see the SI).15 It was found that adding elemental Hg to the reaction of

a

Conditions: aldehyde (1.0 mmol), HBpin (1.2 mmol), and 2 (0.05 mol %), neat, rt, 1 h, N2. bYields of hydroborated products (5) were determined by GC−MS analysis. Isolated yields of primary alcohols (6) are shown in parentheses. cReaction run for 16 h. D

DOI: 10.1021/acs.orglett.8b03583 Org. Lett. XXXX, XXX, XXX−XXX

Organic Letters



acetophenone with HBpin under standard conditions resulted in complete suppression of the reaction, indicating that the reaction was likely to take place through a heterogeneous mechanism. We also noticed that the reaction between 2 and HBpin (10 equiv) in THF gave immediately a black precipitate which was insoluble in common organic solvents. We assume that an active copper cluster hydride species which is responsible for the effective hydroboration might have formed,12 although the nature of this species remains to be determined. We are currently in the progress of a detailed mechanistic investigation as well as the potential of this and relevant chiral copper catalysts in asymmetric hydroboration. In conclusion, we report a divalent copper-catalyzed chemoselective hydroboration of ketones and aldehydes using a novel nonanuclear copper(II) cluster complex. The method features low catalyst loading, mild, solvent-free and activator-free conditions, a broad scope of substrates, and good functional group compatibility, which imply promising practical applicability of this catalyst. The observation of the easy-to-make, bench-stable CuII−Schiff base complex as a highly efficient reduction catalyst would pave the way to future catalyst design based on Earth-abundant and inexpensive copper.



REFERENCES

(1) (a) Zuo, W.; Morris, R. H. Synthesis and Use of an Asymmetric Transfer Hydrogenation Catalyst Based on Iron(II) for the Synthesis of Enantioenriched Alcohols and Amines. Nat. Protoc. 2015, 10, 241− 257. (b) Mikhailine, A.; Maishan, M. I.; Lough, A. J.; Morris, R. H. The Mechanism of Efficient Asymmetric Transfer Hydrogenation of Acetophenone Using an Iron(II) Complex Containing an (S,S)Ph2PCH2CH = NCHPhCHPh-N = CHCH2PPh2 Ligand: Partial Ligand Reduction Is the Key. J. Am. Chem. Soc. 2012, 134, 12266− 12280. (c) Zell, T.; Milstein, D. Hydrogenation and Dehydrogenation Iron Pincer Catalysts Capable of Metal Ligand Cooperation by Aromatization/Dearomatization. Acc. Chem. Res. 2015, 48, 1979− 1994. (d) Chakraborty, S.; Bhattacharya, P.; Dai, H.; Guan, H. Nickel and Iron Pincer Complexes as Catalysts for the Reduction of Carbonyl Compounds. Acc. Chem. Res. 2015, 48, 1995−2003. (e) Garbe, M.; Junge, K.; Beller, M. Homogeneous Catalysis by Manganese-Based Pincer Complexes. Eur. J. Org. Chem. 2017, 4344− 4362. (2) (a) Chirik, P. J. Iron- and Cobalt-Catalyzed Alkene Hydrogenation: Catalysis with Both Redox-Active and Strong Field Ligands. Acc. Chem. Res. 2015, 48, 1687−1695. (b) Zhang, G.; Scott, B. L.; Hanson, S. K. Mild and Homogeneous Cobalt-Catalyzed Hydrogenation of C = C, C = O, and C = N Bonds. Angew. Chem., Int. Ed. 2012, 51, 12102−12106. (c) Zhang, G.; Vasudevan, K. V.; Scott, B. L.; Hanson, S. K. Understanding the Mechanisms of Cobalt-Catalyzed Hydrogenation and Dehydrogenation Reactions. J. Am. Chem. Soc. 2013, 135, 8668−8681. (d) Tokmic, K.; Markus, C. R.; Zhu, L.; Fout, A. R. Well-Defined Cobalt(I) Dihydrogen Catalyst: Experimental Evidence for a Co(I)/Co(III) Redox Process in Olefin Hydrogenation. J. Am. Chem. Soc. 2016, 138, 11907−11913. (e) Nerush, A.; Vogt, M.; Gellrich, U.; Leitus, G.; Ben-David, Y.; Milstein, D. Template Catalysis by Metal−Ligand Cooperation. C−C Bond Formation via Conjugate Addition of Non-activated Nitriles under Mild, Base-free Conditions Catalyzed by a Manganese Pincer Complex. J. Am. Chem. Soc. 2016, 138, 6985−6997. (3) (a) Mukhopadhyay, T. K.; Flores, M.; Groy, T. L. R.; Trovitch, J. A Highly Active Manganese Precatalyst for the Hydrosilylation of Ketones and Esters. J. Am. Chem. Soc. 2014, 136, 882−885. (b) Palmer, W. N.; Diao, T.; Pappas, I.; Chirik, P. J. High-Activity Cobalt Catalysts for Alkene Hydroboration with Electronically Responsive Terpyridine and α-Diimine Ligands. ACS Catal. 2015, 5, 622−626. (c) Zhang, G.; Yin, Z.; Tan, J. Cobalt(II)-catalysed transfer hydrogenation of olefins. RSC Adv. 2016, 6, 22419−22423. (d) Yin, Z.; Zeng, H.; Wu, J.; Zheng, S.; Zhang, G. Cobalt-Catalyzed Synthesis of Aromatic, Aliphatic, and Cyclic Secondary Amines via a “Hydrogen-Borrowing” Strategy. ACS Catal. 2016, 6, 6546−6550. (e) Zhang, G.; Yin, Z.; Zheng, S. Cobalt-Catalyzed N-Alkylation of Amines with Alcohols. Org. Lett. 2016, 18, 300−303. (f) Zhang, G.; Wu, J.; Zeng, H.; Zhang, S.; Yin, Z.; Zheng, S. Cobalt-Catalyzed αAlkylation of Ketones with Primary Alcohols. Org. Lett. 2017, 19, 1080−1083. (g) Zhang, G.; Wu, J.; Wang, M.; Zeng, H.; Cheng, J.; Neary, M. C.; Zheng, S. Cobalt-Catalyzed Regioselective Hydroboration of Terminal Alkenes. Eur. J. Org. Chem. 2017, 2017, 5814− 5818. (4) Arai, N.; Ohkuma, T. Modern Reduction Methods; Wiley-VCH: Weinheim, 2008; pp 159−181. (5) (a) Zhang, G.; Zeng, H.; Wu, J.; Yin, Z.; Zheng, S.; Fettinger, J. C. Highly Selective Hydroboration of Alkenes, Ketones and Aldehydes Catalyzed by a Well-Defined Manganese Complex. Angew. Chem., Int. Ed. 2016, 55, 14369−14372. (b) Vasilenko, V.; Blasius, C. K.; Wadepohl, H.; Gade, L. H. Mechanism-Based Enantiodivergence in Manganese Reduction Catalysis: A Chiral Pincer Complex for the Highly Enantioselective Hydroboration of Ketones. Angew. Chem., Int. Ed. 2017, 56, 8393−8397. (c) Obligacion, J. V.; Chirik, P. J. Highly Selective Bis(imino)pyridine Iron-Catalyzed Alkene Hydroboration. Org. Lett. 2013, 15, 2680−2683. (d) Obligacion, J. V.; Chirik, P. J. Bis(imino)pyridine Cobalt-Catalyzed Alkene Isomerization−Hydroboration: A Strategy for Remote Hydrofunctionalization with Terminal Selectivity. J. Am. Chem. Soc. 2013, 135,

ASSOCIATED CONTENT

* Supporting Information S

The Supporting Information is available free of charge on the ACS Publications website at DOI: 10.1021/acs.orglett.8b03583. Additional experimental details and copies of NMR spectra (PDF) Accession Codes

CCDC 1870930 contains the supplementary crystallographic data for this paper. These data can be obtained free of charge via www.ccdc.cam.ac.uk/data_request/cif, or by emailing [email protected], or by contacting The Cambridge Crystallographic Data Centre, 12 Union Road, Cambridge CB2 1EZ, UK; fax: +44 1223 336033.



Letter

AUTHOR INFORMATION

Corresponding Authors

*E-mail: [email protected]. *E-mail: [email protected]. *E-mail: [email protected]. ORCID

Guoqi Zhang: 0000-0001-6071-8469 Notes

The authors declare no competing financial interest.



ACKNOWLEDGMENTS We are grateful to the donors of the American Chemical Society Petroleum Research Fund for partial support of this work (54247-UNI3). We also acknowledge the support from the PSC−CUNY awards (69069-0047, 60328-0048), the Seed grant from the Office for Advancement of Research, and the PRISM program at John Jay College, CUNY. Partial support from National Science Foundation (CHE-1464543) is gratefully acknowledged. E

DOI: 10.1021/acs.orglett.8b03583 Org. Lett. XXXX, XXX, XXX−XXX

Letter

Organic Letters 19107−19110. (e) Tseng, K.-N. T.; Kampf, J. W.; Szymczak, N. K. Regulation of Iron-Catalyzed Olefin Hydroboration by Ligand Modifications at a Remote Site. ACS Catal. 2015, 5, 411−415. (f) Greenhalgh, M. D.; Thomas, S. P. Chemo-, regio-, and stereoselective iron-catalysed hydroboration of alkenes and alkynes. Chem. Commun. 2013, 49, 11230−11232. (6) (a) Nakajima, K.; Kato, T.; Nishibayashi, Y. Hydroboration of Alkynes Catalyzed by Pyrrolide-Based PNP Pincer−Iron Complexes. Org. Lett. 2017, 19, 4323−4326. (b) Ben-Daat, H.; Rock, C. L.; Flores, M.; Groy, T. L.; Bowman, A. C.; Trovitch, R. J. Hydroboration of alkynes and nitriles using an α-diimine cobalt hydride catalyst. Chem. Commun. 2017, 53, 7333−7336. (7) Das, U. K.; Higman, C. S.; Gabidullin, B.; Hein, J. E.; Baker, R. T. Efficient and Selective Iron-Complex-Catalyzed Hydroboration of Aldehydes. ACS Catal. 2018, 8, 1076−1081 and references cited therein . (8) For a recent review, see: (a) Chong, C. C.; Kinjo, R. Catalytic Hydroboration of Carbonyl Derivatives, Imines, and Carbon Dioxide. ACS Catal. 2015, 5, 3238−3259. For selected examples, see: (b) Lummis, P. A.; Momeni, M. R.; Lui, M. W.; McDonald, R.; Ferguson, M. J.; Miskolzie, M.; Brown, A.; Rivard, E. Accessing zinc monohydride cations through coordinative interactions. Angew. Chem., Int. Ed. 2014, 53, 9347−9351. (c) Arrowsmith, M.; Hadlington, T. J.; Hill, M. S.; Kociok-Kohn, G. Magnesium-catalysed hydroboration of aldehydes and ketones. Chem. Commun. 2012, 48, 4567−4569. (d) Mukherjee, D.; Ellern, A.; Sadow, A. D. Magnesiumcatalyzed hydroboration of esters: evidence for a new zwitterionic mechanism. Chem. Sci. 2014, 5, 959−964. (e) Guo, J.; Chen, J.; Lu, Z. Cobalt-catalyzed asymmetric hydroboration of aryl ketones with pinacolborane. Chem. Commun. 2015, 51, 5725−5727. (f) Oluyadi, A. A.; Ma, S.; Muhoro, C. N. Titanocene(II)-Catalyzed Hydroboration of Carbonyl Compounds. Organometallics 2013, 32, 70−78. (g) Yang, Z.; Zhong, M.; Ma, X.; De, S.; Anusha, C.; Parameswaran, P.; Roesky, H. W. An Aluminum Hydride That Functions like a Transition-Metal Catalyst. Angew. Chem., Int. Ed. 2015, 54, 10225−10229. (h) Jakhar, V. K.; Barman, M. K.; Nembenna, S. Aluminum Monohydride Catalyzed Selective Hydroboration of Carbonyl Compounds. Org. Lett. 2016, 18, 4710−4713. (i) Wang, W.; Shen, X.; Zhao, F.; Jiang, H.; Yao, W.; Pullarkat, S. A.; Xu, L.; Ma, M. Ytterbium-Catalyzed Hydroboration of Aldehydes and Ketones. J. Org. Chem. 2018, 83, 69−74. (j) Tamang, S. R.; Findlater, M. Iron Catalyzed Hydroboration of Aldehydes and Ketones. J. Org. Chem. 2017, 82, 12857− 12860. (k) Baishya, A.; Baruah, S.; Geetharani, K. Efficient Hydroboration of Carbonyls by an Iron(II) Amide Catalyst. Dalton Trans. 2018, 47, 9231−9236. (l) Kaithal, A.; Chatterjee, B.; Gunanathan, C. Ruthenium Catalyzed Selective Hydroboration of Carbonyl Compounds. Org. Lett. 2015, 17, 4790−4793. (9) (a) Wu, J.; Zeng, H.; Cheng, J.; Zheng, S.; Golen, J. A.; Manke, D. R.; Zhang, G. Cobalt(II) Coordination Polymer as a Precatalyst for Selective Hydroboration of Aldehydes, Ketones, and Imines. J. Org. Chem. 2018, 83, 9442−9448. (b) Li, L.; Liu, E.; Cheng, J.; Zhang, G. Iron(II) coordination polymer catalysed hydroboration of ketones. Dalton. Trans. 2018, 47, 9579−9584. (c) Eedugurala, N.; Wang, Z.; Chaudhary, U.; Nelson, N.; Kandel, K.; Kobayashi, T.; Slowing, I. I.; Pruski, M.; Sadow, A. D. Mesoporous Silica-Supported Amidozirconium-Catalyzed Carbonyl Hydroboration. ACS Catal. 2015, 5, 7399− 7414. (d) Manna, K.; Ji, P.; Greene, F. X.; Lin, W. Metal−Organic Framework Nodes Support Single-Site Magnesium−Alkyl Catalysts for Hydroboration and Hydroamination Reactions. J. Am. Chem. Soc. 2016, 138, 7488−7491. (e) Huang, Z.; Liu, D.; Camacho-Bunquin, J.; Zhang, G.; Yang, D.; López-Encarnación, J. M.; Xu, Y.; Ferrandon, J. K.; Poluektov, O. G.; Jellinek, J.; Lei, A.; Bunel, E. E.; Delferro, M. Supported Single-Site Ti(IV) on a Metal−Organic Framework for the Hydroboration of Carbonyl Compounds. Organometallics 2017, 36, 3921−3930. (f) Manna, K.; Ji, P.; Lin, Z.; Greene, F. X.; Urban, A.; Thacker, N. C.; Lin, W. Chemoselective Single-Site Earth-Abundant Metal Catalysts at Metal-Organic Framework Nodes. Nat. Commun. 2016, 7, 12610−12621.

(10) (a) Noh, D.; Chea, H.; Ju, J.; Yun, J. Highly regio- and enantioselective copper-catalyzed hydroboration of styrenes. Angew. Chem., Int. Ed. 2009, 48, 6062−6064. (b) Iwamoto, H.; Kubota, K.; Ito, H. Highly selective Markovnikov hydroboration of alkylsubstituted terminal alkenes with a phosphine−copper(I) catalyst. Chem. Commun. 2016, 52, 5916−5919. (c) Romero, E. A.; Jazzar, R.; Bertrand, G. (CAAC)CuX-catalyzed hydroboration of terminal alkynes with pinacolborane directed by the X-ligand. J. Organomet. Chem. 2017, 829, 11−13. (d) Wyss, C.; Bitting, J.; Bacsa, J.; Gray, T. G.; Sadighi, J. P. Bonding and Reactivity of a Dicopper(I) μ−Boryl Cation. Organometallics 2016, 35, 71−74. (e) Zeng, X.; Gong, C.; Guo, H.; Xu, H.; Zhang, J.; Xie, J. Efficient Heterogeneous Hydroboration of Alkynes: Enhancing the Catalytic Activity by Cu(0) Incorporated CuFe2O4 nanoparticles. New J. Chem. 2018, 42, 17346−17350. (f) Zhang, Q.; Liu, Y.; Wang, T.; Zhang, X.; Long, C.; Wu, Y.-D.; Wang, M.-X. Mechanistic Study on Cu(II)-Catalyzed Oxidative Cross-Coupling Reaction between Arenes and Boronic Acids under Aerobic Conditions. J. Am. Chem. Soc. 2018, 140, 5579− 5587. (11) To the best of our knowledge, only one example on the hydrosilylation of ketones was reported using copper(II) acetate complexes of biphosphine ligands. See: Zhou, J.-N.; Fang, Q.; Hu, Y.H.; Yang, L.-Y.; Wu, F.-F.; Xie, L.-J.; Wu, J.; Li, S. Copper(II)catalyzed enantioselective hydrosilylation of halo-substituted alkyl aryl and heteroaryl ketones: asymmetric synthesis of (R)-fluoxetine and (S)-duloxetine. Org. Biomol. Chem. 2014, 12, 1009−1017. (12) Bagherzadeh, S.; Mankad, N. P. Extremely efficient hydroboration of ketones and aldehydes by copper carbene catalysis. Chem. Commun. 2016, 52, 3844−3846. (13) Docherty, J. H.; Peng, J.; Dominey, A. P.; Thomas, S. P. Activation and discovery of earth-abundant metal catalysts using sodium tert-butoxide. Nat. Chem. 2017, 9, 595−600. (14) Zhang, G.; Proni, G.; Zhao, S.; Constable, E. C.; Housecroft, C. E.; Zampese, J. A.; Neuburger, M. Chiral tetranuclear and dinuclear copper(II) complexes for TEMPO-mediated aerobic oxidation of alcohols: are four metal centres better than two? Dalton Trans. 2014, 43, 12313−12320. (15) Cho, W. K.; Lee, J. K.; Kang, S. M.; Chi, Y. S.; Lee, H.-S.; Choi, I. S. Gold-Catalyzed Cyanosilylation Reaction: Homogeneous and Heterogeneous Pathways. Chem. - Eur. J. 2007, 13, 6351−6358.

F

DOI: 10.1021/acs.orglett.8b03583 Org. Lett. XXXX, XXX, XXX−XXX