Cobalt(II) Coordination Polymer as a Precatalyst for Selective

Jun 22, 2018 - Department of Sciences, John Jay College and Ph.D. Program in Chemistry, The Graduate Center of the City University of New York, New Yo...
0 downloads 0 Views 975KB Size
Note Cite This: J. Org. Chem. 2018, 83, 9442−9448

pubs.acs.org/joc

Cobalt(II) Coordination Polymer as a Precatalyst for Selective Hydroboration of Aldehydes, Ketones, and Imines Jing Wu,†,‡ Haisu Zeng,†,‡ Jessica Cheng,† Shengping Zheng,*,‡ James A. Golen,§ David R. Manke,§ and Guoqi Zhang*,† †

J. Org. Chem. 2018.83:9442-9448. Downloaded from pubs.acs.org by UNIV OF SUSSEX on 08/17/18. For personal use only.

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 § Department of Chemistry and Biochemistry, University of Massachusetts Dartmouth, North Dartmouth, Massachusetts 02747, United States S Supporting Information *

ABSTRACT: Highly effective hydroboration precatalyst is developed based on a cobalt(II)-terpyridine coordination polymer (CP). The hydroboration of ketones, aldehydes, and imines with pinacolborane (HBpin) has been achieved using the recyclable CP catalyst in the presence of an air-stable activator. A wide range of substrates containing polar CO or CN bonds have been hydroborated selectively in excellent yields under ambient conditions.

C

hydroboration of unsaturated bonds. In 2016, Lin and coworkers revealed that single-site magnesium alkyl supported on the metal nodes of TPHN-MOF [UiO-69 topology, TPHN = 4, 4-bis(carboxyphenyl)-2-nitro-1,1′-biphenyl] cataylzed the hydroboration of carbonyl compounds and imines with high TONs, where the catalyst was post-synthesized by reacting active Me2Mg with as-synthesized MOF materials.11 More recently, Delferro et al. reported a MOF-supported single-site Ti(IV) catalyst for the hydroboration of carbonyl compounds.12 This catalyst also utilizes the post-synthetic modification on an existing Zr-MOF with Ti(OiPr)4 to generate catalytically active Ti(IV) sites. Typically, in those limited examples the active hydroboration catalysts were generated by modifying MOF materials (constructed by complex organic linkers) with an air- and moisture-sensitive metal reagent for immobilization of catalytically active sites. Herein, we report the first example of a readily available cobalt(II) coordination polymer as a precatalyst for selective hydroboration of aldehydes, ketones, and imines under mild conditions with high efficiency, in the absence of air- and moisture-sensitive organometallic reagents. Previously, well-defined cobalt(II) complexes based on tridentate NNN ligands have found applications in the hydroboration of a variety of nonpolar and polar multiple bonds, in the presence of activators (NaHB(OEt)3 or LiCH2SiMe3).8e,9b,13 We also reported a manganese(II) dialkyl complex of 2,2′;6′,2″-terpyridine (tpy) that catalyzed the

atalytic hydroboration of carbonyl compounds and imines provides an important and convenient approach to functionalized alcohols and amines, respectively, comparing to the conventional stoichiometric reduction using hydride reagents (such as LiAlH4 or NaBH4), which usually suffer from poor functional group tolerance and product selectivity (aldehydes vs ketones or over-reduction), modest reaction rates (with hindered ketones and imines), and often harsh conditions.1 Recent efforts have been made to develop efficient catalysts for the hydroboration of carbonyl compounds and imines with relatively less-active reductants, such as HBpin. In a recent review, Kinjo and co-worker have documented the catalytic methods for this transformation.2 A number of discrete metal complexes have been utilized as catalysts for this reaction.3,4 Among the nonprecious metals explored, cobalt is quite attractive in catalytic hydrogenation and dehydrogenation processes.5,6 Major progress has been achieved for the observation of well-defined homogeneous cobalt hydrogenation catalysts under mild conditions, including the Co-PNP pincer complex studied by our group,7,8 and some have also found applications in catalytic hydroboration of alkenes, alkynes, and nitriles.8e,9 While homogeneous metal catalysts for hydroboration of carbonyl compounds and imines have been extensively explored, heterogeneous approaches for this conversion were little addressed. Mesoporous silica-supported amidozirconium catalysts have been previously reported for the hydroboration of carbonyl compounds.10 It was recently realized that incorporating active metal sites within a metal−organic framework (MOF) could be a viable strategy for catalytic © 2018 American Chemical Society

Received: May 7, 2018 Published: June 22, 2018 9442

DOI: 10.1021/acs.joc.8b01094 J. Org. Chem. 2018, 83, 9442−9448

Note

The Journal of Organic Chemistry

interchain π-stacking interaction (the shortest interatomic C··· C distance between the adjacent chains is 3.495 (5) Å), thanks to the perfect coplanarity of the aromatic region. In contrast, 4 crystallized in the orthorhombic space group Pbcn. CoII is hexacoordinate with four N atoms of four distinct ligands and two Cl atoms to form an octahedral coordination geometry. The structural expansion around CoII centers results in the formation of a 1D coordination framework composed of cagelike subunits formed by a [4 + 4] metal−ligand assembly (Figure 1). It is worth mentioning that although numerous CPs derived from related tpy ligands have been reported,18 3 and 4 are the first examples of structurally characterized CPs from the assembly of tpy with CoCl2. To demonstrate the capability of CPs 3 and 4 in catalytic hydroboration of carbonyl compounds with HBpin, initially we employed similar reaction conditions from the known cobaltcatalyzed hydroboration involving discrete complexes.15 The results are summarized in Table 1. When only 0.1 mol % of 3

Markovnikov-selective hydroboration of alkenes.14 Significantly, Thomas and co-workers have revealed that air-stable cobalt(II) or iron(II) dihalide complexes of NNN ligands could act as effective precatalysts for alkene hydroboration and hydrosilylation while being activated by sodium tert-butoxide.15 To design extended coordination polymers as heterogeneous precatalysts, we envisioned that anchoring an additional N coordination site on the tpy backbone would form a ditopic ligand (1, Scheme 1) that is suitable for Scheme 1. Ligands 1 and 2 Derived from 2,2′;6′,2″Terpyridine

Table 1. Condition Screening for Cobalt(II) CP-Catalyzed Hydroboration of Acetophenone with Pinacolboranea construction of extended coordination framework upon coordinating with CoII dihalide.16 Thus, the metal centers within the framework should adopt a coordination environment similar to that reported previously in discrete complexes.15 To compare, the divergent terpyridine ligand, 4′-(4-dimethylaminophenyl)-4,2′;6′,4″-tpy (2), was also employed for similar coordination framework.17 Coordination-driving self-assembly of known 1 and 2 with cobalt dichloride by a layering technique has afforded goodquality crystals of 3 and 4 in high yields, respectively (see SI). Both crystalline compounds were characterized by IR, elemental analysis, and X-ray crystallography. Structural analysis revealed that 3 crystallized in the orthorhombic space group Pcca. As expected, the CoII center resides within the N3 cavity of one ligand, and an extra Npyridyl atom from another ligand links to the CoII center to extend the structure. Two coordinated chloride anions are located in the axial direction of CoN4 coordination sphere, leading to a pseudooctahedral geometry around CoII (Figure 1). The assembled structure is polymeric and propagates along the crystallographic c axis to form a one-dimensional (1D) coordination polymeric chain. The chains are further packed by a strong

entry 1 2 3 4 5 6 7 8c 9d 10 11 12 13e 14f 15g

catalyst 3 4 3 3 3 3 CoCl2 Co(tpy)Cl2 3 3 3 3 3 3

activator t

KO Bu KOtBu NaOtBu KOH NaHB(OEt)3 KOtBu KOtBu KOtBu KOtBu KOtBu KOtBu KOtBu KOtBu KOtBu

solvent

yield (%)b

THF THF THF THF THF THF THF THF THF Et2O toluene pentane THF THF THF

>99 95 98 97 70 0 8 54 71 88 68 65 25 87 86

a

Conditions: Acetophenone (1.0 mmol), HBpin (1.1 mmol), cobalt CP (0.1 mol %), activator (2 mol %), and solvent (1 mL), 25 °C, 1 h, N2. bDetermined by GC analysis with hexamethylbenzene as an internal standard. c2 mol % catalyst was used. d1 mol % catalyst was used. eCatecholborane was used as the boron source. fReaction run in the air. g0.01 mol % 3 was used and reaction run for 24 h.

and potassium tert-butoxide (KOtBu, 2 mol %) were used, the hydroboration of acetophenone proceeded well in THF at room temperature, affording the hydroborated product 5 in >99% GC yield after 4 h. In comparison, 4 is also an active precatalyst under the same conditions, even though the yield was slightly lower. Similar results were found when using NaOtBu or KOH as an activator, while NaHB(OEt)3 acts as a poorer activator for this reaction. Control experiments show that the use of CP 3 was crucial for the excellent reactivity, whereas the combination of CoCl2 (2 mol %) and KOtBu also exhibited modest catalytic activity. Interestingly, the discrete complex, Co(tpy)Cl2, displayed inferior efficiency compared to the CP analogue. In addition, solvent effects were studied, and the results indicated that THF is more suitable than several

Figure 1. ORTEP structures of 3 and 4 plotted at 50% thermal ellipsoid probability level and the ball−stick representation of the extended network in 3 and 4, respectively. H-atoms are omitted for clarity. 9443

DOI: 10.1021/acs.joc.8b01094 J. Org. Chem. 2018, 83, 9442−9448

Note

The Journal of Organic Chemistry

yields were obtained for 2-acetylnaphthalene (6i) and diaryl ketones (6j and 6k). However, the cyclic ketone, tetralone, was hydroborated completely (6l) to afford the corresponding cyclic alcohol in 97% yield. In addition, the steric hindered ketone, 2′-methylacetophenone also proceeded well with 6m being isolated in 90% yield. Unfortunately, ketones containing a reducible functional group were not suitable substrates for effective ketone hydroboration under standard conditions. For example, when 4′-nitroacetophenone was used, a complicated mixture of hydroborated products was detected (∼50% conversion of ketone) and no desired alcohol product could be isolated. Finally, more challenging alkyl ketones proved to be suitable substrates (6n−q). The heterogeneous hydroboration of ketones catalyzed by 3 allows us to study the recycling and reuse of the catalyst. In a recycling reaction using acetophenone as a model substrate (Figure S3, see SI), catalyst 3 (2 mol %) was filtered out of the reaction mixture and reused for 5 more reaction cycles (the activator was reloaded in each reuse cycle). The results showed no obvious drop in the yields determined by GC, indicating excellent recyclability of the CP precatalyst. Under the same conditions, we further investigated the hydroboration of aldehydes to prepare primary alcohols. Generally, aldehyde hydroboration proceeded smoothly in 16 h, with the alcohol products being isolated in good to excellent yields after column chromatography (Scheme 3). Benzalde-

other solvents tested for this reaction. When catecholborane was used as a boron source under the optimized conditions (as in entry 1, Table 1), the corresponding hydroborated product was found in only 25% GC yield (entry 13, Table 1). It was interesting to note that the hydroboration even proceeded smoothly in the air, giving 5a in 87% yield (entry 14, Table 1). Finally, when 0.01 mol % of 3 was used as the precatalyst, product 5 was detected in 86% yield, corresponding to a turnover number (TON) of 8600, comparable to the MOFsupported single site catalysts reported previously.11,12 Next, we sought to investigate the application of 3-catalyzed hydroboration under optimized conditions. A variety of ketones were first examined, and the results are summarized in Scheme 2. For all reactions indicated here, the hydroborated Scheme 2. CP-Catalyzed Hydroboration of Ketonesa

Scheme 3. CP-Catalyzed Hydroboration of Aldehydesa

a

Conditions: Ketone (1.0 mmol), HBpin (1.1 mmol), cobalt CP (0.1 mol %), KOtBu (2 mol %), and THF (1 mL), 25 °C, 4 h, N2. Yields of isolated products of alcohols (GC yields of boronate esters are shown in parentheses). b10.0 mmol acetophenone was used.

a

Conditions: Aldehydes (1.0 mmol), HBpin (1.1 mmol), cobalt CP (0.1 mol %), KOtBu (2 mol %), and THF (1 mL), 25 °C, 16 h, N2. Yields of isolated products of alcohols (GC yields of boronate intermediates are shown in parentheses). b2.0 equiv of HBpin was used, and ∼34% starting material was detected by GC. c∼20% starting material was detected by GC. d2.0 equiv of HBpin was used. en.r.: no reaction.

products were confirmed by GC-MS analysis. Purification of the products through silica gel column chromatography afforded hydrolyzed products, the alcohols (6) in high yields, consistent with the results reported previously.14 The hydroboration of acetophenone was readily approached in a gram scale, yielding 1-phenylethanol (6a) in 94% isolated yield. Functionalized aryl ketones were smoothly hydroborated under the standard conditions. Halogenated acetophenones were completely converted to the corresponding boronate esters and isolated as alcohols in excellent yields (6b−d). Methyl phenyl ketones containing electron-donating groups and a bulkier isopropyl phenyl ketone also proceeded well (6e−g). Cyclopropyl phenyl ketone was used to produce 6h in a reasonable yield without ring-opening, indicating that no radical mechanism was involved. In addition, relatively lower

hydes bearing halogen, electron-withdrawing, or electrondonating groups were hydroborated to the corresponding boronate esters in almost quantitative conversions, and alcohols were isolated in high yields (7a−i), except that in the case of p-cyanobenzaldehyde the reaction was a little sluggish and the desired product 7e was obtained in a modest yield. In addition, benzaldehyde containing an ester functional group was selectively hydroborated on the aldehyde, affording 9444

DOI: 10.1021/acs.joc.8b01094 J. Org. Chem. 2018, 83, 9442−9448

Note

The Journal of Organic Chemistry

In summary, we report the first cobalt-based coordination polymer-catalyzed selective hydroboration of ketones, aldehydes, and imines under mild, heterogeneous conditions. In comparison with other nonprecious metal catalysts, the CoII CP catalysis system is easy-to-make, air-stable, highly efficient, reusable, and tolerant of a wide range of substrates. This work implies new catalytic applications of simple coordination polymers based on earth-abundant metals, paving the way for development of practical and sustainable catalysts for key organic transformations.

7j in a reasonable yield. Furthermore, trans-cinnamaldehyde containing conjugated CC and CO bonds was selectively hydroborated on the CO part (7k), even though 2.0 equiv of HBpin was used. While the heterocyclic aldehyde, 2thiophenecarboxaldehyde is a suitable substrate for the hydroboration (7l), 2-pyridinecarboxaldehyde did not react at all (7m), probably due to the strong binding ability of the substrate to metal centers, which deactivates the catalyst. Finally, aliphatic aldehydes were found to be readily hydroborated in high efficiencies (7n and 7o). Encouraged by these results, we studied the hydroboration of imines with the cobalt CP catalyst. While there was no reaction observed when N-benzylideneaniline was mixed with HBpin under the standard condition at room temperature, at elevated temperature (70 °C) complete hydroboration of the imine occurred, with 8a being isolated in 96% yield (Scheme 4). Accordingly, several substituted aromatic imines containing



EXPERIMETAL SECTION

General Information. The catalytic reactions were carried out under a nitrogen atmosphere using standard glovebox technique. Anhydrous grade solvents and liquid reagents used were obtained from Aldrich or Fisher Scientific and stored over 4 Å molecular sieves. FT-IR spectra were recorded on a Shimadzu 8400S instrument with solid samples under N2 using a Golden Gate ATR accessory. Elemental analyses were performed by Midwest Microlab LLC in Indianapolis, IN. 1H NMR and 13C NMR spectra were obtained at room temperature on a Bruker AV 500 or 600 MHz NMR spectrometer, with chemical shifts (δ) referenced to the residual solvent signal. GC-MS analysis was obtained using a Shimadzu GCMS-QP2010S gas chromatograph mass spectrometer. 2,2′;6′,2″Terpyridine was purchased from Sigma-Aldrich. 4′-(4-Pyridyl)2,2′;6′,2′′-tpy (1), 4′-(4-dimethylaminophenyl)-4,2′;6′,4″-tpy (2), and Co(tpy)Cl 2 were prepared according to the literature procedure.8d,16,17 Experimental Procedure. Synthesis of 3. A solution of 1 (62.0 mg, 0.200 mmol) in MeOH/CH2Cl2 (10 mL, 1:3, v/v) was placed in a test tube. A blank solution of MeOH/CH2Cl2 (6 mL, 1:1, v/v) was layered on the top of the ligand solution, followed by a solution of CoCl2·6H2O (47.4 mg, 0.200 mmol) in MeOH (10 mL). The tube was sealed and allowed to stand at room temperature for a week, during which time X-ray quality light brown crystals grew on the wall and bottom of the tube. The crystals were collected by decanting the solvent and washed with MeOH and then dried in vacuo. Yield: 55.0 mg (89%). FT-IR (solid, cm−1) 3481br, 3057m, 1599s, 1569w, 1539s, 1472m, 1408s, 1346w, 1247s, 1169m, 1079m, 1012s, 897w, 842s, 792s, 748m, 730m. Anal. calcd for C20H14Cl2CoN4·2H2O: C 50.44, H 3.81, N 11.77%. Found C 50.31, H 3.79, N 11.52. Synthesis of 4. A solution of 2 (70.4 mg, 0.200 mmol) in MeOH/ CH2Cl2 (10 mL, 1:3, v/v) was placed in a test tube. A blank solution of MeOH/CH2Cl2 (6 mL, 1:1, v/v) was layered on the top of the ligand solution, followed by a solution of CoCl2·6H2O (47.4 mg, 0.200 mmol) in MeOH (10 mL). The tube was sealed and allowed to stand at room temperature for a week, during which time X-ray quality light brown crystals grew on the bottom of the tube. The crystals were collected by decanting the solvent and washed with MeOH and then dried in vacuo. Yield: 76.6 mg (92% based on 2). FT-IR (solid, cm−1) 3050m, 2900m, 1608s, 1589s, 1557w, 1525s, 1497m, 1434w, 1399s, 1363s, 1216m, 1168m, 1063s, 1012s, 946m, 844s, 817s, 727s, 674s. Anal. calcd for C46H40Cl2CoN8: C 66.19, H 4.83, N 13.42%. Found C 66.05, H 4.90, N 13.30. General Procedure for CP-Catalyzed Hydroboration of Ketones. In a glovebox under N2 atmosphere, catalyst 3 (0.44 mg, 1.0 μmol) and KOtBu (2.2 mg, 20 μmol) were dissolved in THF (1.0 mL) in a 3.8 mL glass vial equipped with a stir bar. Ketone (1.0 mmol) and pinacolborane (140.8 mg, 1.1 mmol) were then added. The reaction mixture was allowed to stir at room temperature for 4 h. After completion of the reaction, the solvent was evaporated, the crude reaction mixture was first analyzed by GC-MS using a dilute CH2Cl2 solution, and then the product was isolated by flash column chromatography with SiO2 using ethyl acetate/hexane as eluent. The pure products were characterized by 1H and 13C NMR spectroscopies. General Procedure for CP-Catalyzed Hydroboration of Aldehydes. In a glovebox under N2 atmosphere, catalyst 3 (0.44 mg, 1.0

Scheme 4. CP-Catalyzed Hydroboration of Iminesa

a

Conditions: Imines (1.0 mmol), HBpin (1.1 mmol), cobalt CP (0.1 mol %), KOtBu (2 mol %), and THF (1 mL), 70 °C, 16 h, N2. Yields of isolated products of amines (GC conversion of imines are shown in parentheses).

fluoro, nitro, or ester groups have been converted to the corresponding amines (8b−d) in good yields by the same protocol, demonstrating better applicability of the CP precatalyst. The chemoselectivity of CP-catalyzed hydroboration of ketones and aldehydes was further examined. First, the hydroboration of equimolar acetophenone and benzaldehyde with HBpin (1.0 equiv) was conducted in a competition study (Scheme S1), and the results indicated that the hydroboration of aldehyde is much preferred, providing only hydroborated product of benzaldehyde. In contrast, the discrete catalyst, Co(tpy)Cl2/KOtBu is much less active, whereas the trend of chemoselectivity remains the same (Scheme S1A). This trend was further confirmed in a similar experiment using substituted ketone and aldehyde as substrates when 3 was the precatalyst. It was also observed that hydroboration of acetophenone was not affected by the presence of equimolar N-benzylideneaniline at room temperature, showing good chemoselectivity for ketones over imines. In addition, 4-acetylbenzaldehyde was used to study the intramolecular selectivity. It was found that the hydroboration reaction mainly occurred at the aldehyde, affording 9 in 81% yield after column chromatography. The chemoselectivity on aldehydes over ketones observed here is similar to those found in ruthenium-, lanthanide-, or aluminum-catalyzed hydroboration of carbonyl compounds.4,19 9445

DOI: 10.1021/acs.joc.8b01094 J. Org. Chem. 2018, 83, 9442−9448

Note

The Journal of Organic Chemistry μmol) and KOtBu (2.2 mg, 20 μmol) were dissolved in THF (1.0 mL) in a 3.8 mL glass vial equipped with a stir bar. Aldehyde (1.0 mmol) and pinacolborane (140.8 mg, 1.1 mmol) were then added. The reaction mixture was allowed to stir at room temperature for 16 h. After completion of the reaction, the solvent was evaporated, the crude reaction mixture was first analyzed by GC-MS using a dilute CH2Cl2 solution, and then the product was isolated by flash column chromatography with SiO2 using ethyl acetate/hexane as eluent. The pure products were characterized by 1H and 13C NMR spectroscopies. General Procedure for CP-Catalyzed Hydroboration of Imines. In a glovebox under N2 atmosphere, catalyst 3 (0.44 mg, 1.0 μmol) and KOtBu (2.2 mg, 20 μmol) was dissolved in THF (1.0 mL) in a 20 mL Schlenk tube equipped with a stir bar. Imine (1.0 mmol) and pinacolborane (140.8 mg, 1.1 mmol) were then added. The reaction mixture was sealed and allowed to heat to 70 °C for 16 h. After completion of the reaction, the solvent was evaporated , the crude reaction mixture was first analyzed by GC-MS using a dilute CH2Cl2 solution, and then the product was isolated by flash column chromatography with SiO2 using ethyl acetate/hexane as eluent. The pure products were characterized by 1H and 13C NMR spectroscopies. Competing Experiment for Selective Hydroboration of Ketones vs Aldehydes. In a glovebox under N2 atmosphere, catalyst 3 (0.44 mg, 1.0 μmol) and KOtBu (2.2 mg, 20 μmol) were dissolved in THF (1.0 mL) in a 3.8 mL glass vial equipped with a stir bar. The ketone (1.0 mmol), aldehyde (1.0 mmol), and pinacolborane (128.0 mg, 1.0 mmol) were then added sequentially. The reaction mixture was allowed to stir at room temperature for 16 h. The reaction was exposed to the air, and the solvent was evaporated. The products were analyzed by GC-MS using hexamethylbenzene as an internal reference. Recycling and Reusing Experiment for Hydroboration of Acetophenone. In a glovebox under N2 atmosphere, catalyst 3 (4.4 mg, 10 μmol) and KOtBu (2.2 mg, 20 μmol) were dissolved in THF (1.0 mL) in a small vial equipped with a stir bar. Acetophenone (120.0 mg, 1.0 mmol) and pinacolborane (140.8 mg, 1.1 mmol) were then added. The reaction mixture was allowed to stir at room temperature for 4 h. Then, the solid was centrifuged out of suspension and extracted with hexane three times. The combined organic extracts were analyzed by GC-MS using hexamethylbenzene as an internal reference. The recovered solid catalyst in 1.0 mL THF was charged into a small vial, to which KOtBu (2.2 mg, 20 μmol) was added. Then, acetophenone (120.0 mg, 1.0 mmol) and pinacolborane (140.8 mg, 1.1 mmol) were added. The resultant mixture was stirred at room temperature in the glovebox for 4 h. The solid was centrifuged out of suspension and extracted with hexane three times. The combined organic extracts were analyzed by GC-MS. The results of five recycling experiments are summarized in Figure S1. Spectroscopic Data for Isolated Products. 1-Phenylethanol (6a).14 Colorless oil. Yield: 117.0 mg (96%). 1H NMR (600 MHz, CDCl3) δ 7.47−7.31 (m, 4H), 7.29 (ddd, J = 8.8, 4.8, 3.4 Hz, 1H), 4.85 (q, J = 6.5 Hz, 1H), 2.63 (s, 1H), 1.47 (d, J = 6.4 Hz, 3H) ppm; 13 C NMR (151 MHz, CDCl3) δ 145.9, 128.5, 127.4, 125.5, 70.3, 25.2 ppm. 1-(4-Fluorophenyl)ethanol (6b).14 Colorless oil. Yield: 133 mg (95%). 1H NMR (600 MHz, CDCl3) δ 7.35−7.33 (m, 2H), 7.05− 7.01 (m, 2H), 4.89 (q, J = 6.5 Hz, 1H), 1.82 (br, 1 H), 1.48 (d, J = 6.5 Hz, 3H) ppm; 13C NMR (151 MHz, CDCl3) δ 162.4 (d, J = 244.9 Hz), 141.9 (d, J = 2.4 Hz), 127.4(d, J = 8.1 Hz), 115.6 (d, J = 21.2 Hz), 70.1 (d, J = 2.3 Hz), 25.6 ppm. 1-(4-Chlorophenyl)ethanol (6c).14 Colorless oil. Yield: 150.3 mg (96%). 1H NMR (600 MHz, CDCl3) δ 7.33−7.28 (m, 4H), 4.86 (q, J = 6.5 Hz, 1H), 2.23 (br, 1H), 1.47 (d, J = 6.5 Hz, 3H) ppm; 13C NMR (151 MHz, CDCl3) δ 144.3, 133.0, 128.6, 126.8, 69.7, 25.2 ppm. 1-(4-Bromophenyl)ethanol (6d).14 Colorless oil. Yield: 191.0 mg (95%). 1H NMR (600 MHz, CDCl3) δ 7.46−7.43 (m, 2H), 7.22− 7.20 (m, 2H), 4.81 (q, J = 6.6 Hz, 1 H), 2.34 (br, 1H), 1.43 (d, J = 6.6 Hz, 3 H) ppm; 13C NMR (151 MHz, CDCl3) δ 145.1, 131.8, 127.4, 121.4, 70.0, 25.5 ppm.

1-(4-Methoxyphenyl)ethanol (6e).14 Yellowish oil. Yield: 149 mg (98%). 1H NMR (600 MHz, CDCl3) δ 7.30 (d, J = 8.7 Hz, 2H), 6.89 (d, J = 8.7 Hz, 2H), 4.84 (q, J = 6.5 Hz, 1 H), 3.81 (s, 1H), 2.27 (br, 1H), 1.45 (d, J = 6.5 Hz, 3 H) ppm. 13C NMR (126 MHz, CDCl3) δ 158.9, 138.1, 126.7, 113.8, 69.9, 55.3, 25.1 ppm. 1-(3-Methoxyphenyl)ethanol (6f).7a Yellowish oil. Yield: 144.6 mg (95%). 1H NMR (600 MHz, CDCl3) δ 7.29−7.27 (m, 1H), 6.97− 6.96 (m, 2H), 6.84−6.82 (m, 1H), 4.88 (q, J = 6.7 Hz, 1H), 3.83 (s, 3H), 2.07 (br, 1H), 1.50 (d, J = 6.7 Hz, 3 H) ppm; 13C NMR (151 MHz, CDCl3) δ 160.1, 147.9, 129.8, 118.0, 113.2, 111.2, 70.6, 55.5, 25.4 ppm. 2-Methyl-1-phenylethanol (6g).19a Colorless oil. Yield: 120 mg (80%). 1H NMR (500 MHz, CDCl3) δ 7.42−7.32 (m, 4H), 7.32− 7.28 (m, 1H), 4.39 (d, J = 6.7 Hz, 1H), 1.99 (qd, J = 6.8, 1.5 Hz, 1H), 1.89 (s, 1H), 1.03 (d, J = 6.7 Hz, 3H), 0.83 (d, J = 6.9 Hz, 3H) ppm; 13 C NMR (126 MHz, CDCl3) δ 143.7, 128.3, 127.5, 126.7, 80.2, 35.4, 19.1, 18.4 ppm. 1-Cyclopropyl Phenylmethanol (6h).14 Yellowish oil. Yield: 106.7 mg (72%). 1H NMR (600 MHz, CDCl3) δ 7.43−7.42 (m, 2H), 7.37−7.35 (m, 2H), 7.31−7.28 (m, 1H), 4.00 (d, J = 8.3 Hz, 1H), 2.21 (br, 1H), 1.25−1.19 (m, 1H), 0.66−0.61 (m, 1 H), 0.58−0.53 (m, 1 H), 0.49−0.45 (m, 1 H), 0.40−0.35 (m, 1 H) ppm; 13C NMR (151 MHz, CDCl3) δ 144.2, 128.6, 127.8, 126.3, 78.8, 19.4, 3.9, 3.1 ppm. 1-(2-Naphthalenyl)ethanol (6i).20 White solid. Mp: 69−70 °C. Yield: 114 mg (66%). 1H NMR (600 MHz, CDCl3) δ 7.85−7.83 (m, 3H), 7.80 (s, 1H), 7.51−7.47 (m, 3H), 5.04 (q, J = 6.5 Hz, 1H), 2.22 (br, 1H), 1.58 (d, J = 6.5 Hz, 3H) ppm; 13C NMR (151 MHz, CDCl3) δ 143.5, 133.6, 133.2, 128.6, 128.2, 128.0, 126.4, 126.1, 124.1, 124.1, 70.8, 25.4 ppm. Diphenylmethanol (6j).4f White solid. Mp: 64−65 °C. Yield: 147.4 mg (80%). 1H NMR (600 MHz, CDCl3) δ 7.43−7.41 (m, 4H), 7.40−7.37 (m, 4H), 7.33−7.30 (m, 2H), 5.85 (s, 1H), 2.38 (br, 1H) ppm; 13C NMR (151 MHz, CDCl3) δ 143.9, 128.5, 127.6, 126.6, 76.3 ppm. Di-p-tolylmethanol (6k).20 White solid. Mp: 66−68 °C. Yield: 157 mg (74%). 1H NMR (600 MHz, CDCl3) δ 7.28 (d, J = 8.4 Hz, 4H), 7.17 (q, J = 8.4 Hz, 4H), 5.78 (br, 1H), 2.36 (s, 7H) ppm; 13C NMR (151 MHz, CDCl3) δ 141.4, 137.4, 129.4, 126.8, 76.2, 21.4 ppm. 1,2,3,4-Tetrahydronaphthalen-1-ol (6l).20 Colorless oil. Yield: 143.7 mg (97%). 1H NMR (600 MHz, CDCl3): δ 7.44−7.42 (m, 1H), 7.22−7.20 (m, 2H), 7.12−7.11 (m, 1H), 4.77−4.75 (m, 1H), 2.86−2.81 (m, 1H), 2.76−2.71 (m, 1H), 2.11 (br, 1H), 2.02−1.95 (m, 1H), 1.92−1.87 (m, 1H), 1.82−1.76 (m, 1H). 13C NMR (151 MHz, CDCl3): δ 139.1, 127.4, 129.2, 128.9, 127.8, 126.4, 68.4, 32.5, 29.5, 19.0 ppm. 1-(2-Methylphenyl)ethanol (6m).20 Colorless oil. Yield: 122.0 mg (90%). 1H NMR (600 MHz, CDCl3) δ 7.54 (dd, J = 7.8, 1.7 Hz, 1H), 7.27 (td, J = 7.5, 1.7 Hz, 1H), 7.21 (td, J = 7.4, 1.6 Hz, 1H), 7.17 (dd, J = 7.7, 1.8 Hz, 1H), 5.12 (q, J = 6.5 Hz, 1H), 2.37 (s, 3H), 2.26 (s, 1H), 1.49 (d, J = 6.5 Hz, 3H) ppm; 13C NMR (151 MHz, CDCl3) δ 143.9, 134.2, 130.4, 127.2, 126.4, 124.6, 66.8, 24.0, 19.0 ppm. 1,3-Diphenylpropan-2-ol (6n).14 Colorless oil. Yield: 201.7 mg (95%). 1H NMR (600 MHz, CDCl3) δ 7.39−7.36 (m, 4H), 7.31− 7.28 (m, 6H), 4.13−4.09 (m, 1H), 2.93−2.90 (m, 2H), 2.84−2.80 (m, 2H), 1.80 (br, 1H) ppm; 13C NMR (151 MHz, CDCl3) δ 138.8, 129.7, 128.8, 126.8, 73.8, 43.7 ppm. 2-Heptanol (6o).20 Colorless oil. Yield: 72 mg (62%). 1H NMR (500 MHz, CDCl3) δ 3.82 (m, 1 H), 1.47 (m, 4 H), 1.33 (m, 5 H), 1.21 (d, J = 6.0 Hz, 3 H), 0.91 (t, J = 7.0 Hz, 3 H) ppm; 13C NMR (126 MHz, CDCl3) δ 68.2, 39.3, 31.9, 25.5, 23.5, 22.7, 14.1 ppm. Benzyl Alcohol (7a).14 Colorless oil. Yield: 102 mg (94%). 1H NMR (600 MHz, CDCl3) δ 7.46−7.38 (m, 2H), 7.38−7.30 (m, 2H), 4.63 (s, 2H), 3.35 (s, 1H); 13C NMR (151 MHz, CDCl3) δ 140.9, 128.5, 127.5, 127.0, 64.9 ppm. 4-Chlorobenzyl Alcohol (7b).14 Colorless oil. Yield: 135.4 mg (95%). 1H NMR (600 MHz, CDCl3) δ 7.35−7.33 (m, 2H), 7.30− 7.28 (m, 2H), 4.65 (s, 1H), 2.45 (s, 1 H), 2.12 (br, 1H) ppm; 13C NMR (151 MHz, CDCl3) δ 139.6, 133.6, 129.0, 128.6, 64.7 ppm. 9446

DOI: 10.1021/acs.joc.8b01094 J. Org. Chem. 2018, 83, 9442−9448

Note

The Journal of Organic Chemistry 2-Chlorobenzyl Alcohol (7c).21 Colorless oil. Yield: 136.9 mg (96%). 1H NMR (600 MHz, CDCl3) δ 7.49 (d, J = 7.7 Hz, 1H), 7.38 (d, J = 7.7 Hz, 2H), 7.30 (t, J = 7.4 Hz, 1H), 7.25 (t, J = 7.4 Hz, 1H), 4.79 (s, 2H), 2.29 (s, 1H) ppm; 13C NMR (151 MHz, CDCl3) δ 138.5, 133.0, 129.6, 129.1, 129.0, 127.3, 63.1 ppm. 4-Nitrobenzyl Alcohol (7d).14 Yellowish oil. Yield: 140.9 mg (92%). 1H NMR (600 MHz, CDCl3) δ 8.21 (d, J = 8.4 Hz, 2H), 7.53 (d, J = 8.4 Hz, 2H), 4.83 (s, 2H), 1.97 (br, 1H) ppm; 13C NMR (151 MHz, CDCl3) δ 148.5, 147.6, 127.3, 124.1, 64.3 ppm. 4-(Hydroxymethyl)benzonitrile (7e).21 Yellowish oil. Yield: 79.8 mg (60%). 1H NMR (600 MHz, CDCl3) δ 7.64 (d, J = 8.0 Hz, 2H), 7.47 (d, J = 8.0 Hz, 2H), 4.77 (s, 2H), 1.96 (br, 1H) ppm; 13C NMR (151 MHz, CDCl3) δ 146.6, 132.6, 127.3, 119.2, 111.5, 64.5 ppm. 4-Methoxybenzyl Alcohol (7f).4f Colorless oil. Yield: 131.2 mg (95%). 1H NMR (600 MHz, CDCl3) δ 7.28 (d, J = 8.6 Hz, 2H), 6.89 (d, J = 8.6 Hz, 2H), 4.60 (s, 2H), 3.80 (s, 3H), 1.81 (br, 1H) ppm; 13 C NMR (151 MHz, CDCl3) δ 159.5, 133.5, 129.0, 114.3, 65.3, 55.6 ppm. 4-(Dimethylamino)benzyl Alcohol (7g).21 Colorless oil. Yield: 137.6 mg (91%). 1H NMR (500 MHz, CDCl3) δ 7.27 (d, J = 8.5 Hz, 2 H), 6.77 (d, J = 9.0 Hz, 2 H), 4.58 (s, 2 H), 2.98 (s, 6 H), 1.97 (br., 1 H) ppm; 13C NMR (126 MHz, CDCl3) δ 150.3, 129.0, 128.7, 112.7, 65.3, 40.8 ppm. Piperonol (7h).14 Colorless oil. Yield: 147.6 mg (97%). 1H NMR (600 MHz, CDCl3) δ 6.85 (s, 1H), 6.79−6.76 (m, 2H), 5.94 (s, 2H), 4.55 (s, 2H), 1.95 (br, 1H) ppm; 13C NMR (151 MHz, CDCl3) δ 148.1, 147.4, 135.2, 120.8, 108.5, 108.2, 101.3, 65.5 ppm. 2-(Methylthio)benzyl Alcohol (7i).22 Colorless oil. Yield: 137.2 mg (89%). 1H NMR (500 MHz, CDCl3) δ 7.15 (d, J = 7.5 Hz, 1 H), 7.05 (m, 2 H), 6.96 (m, 1 H), 4.52 (d, J = 2 Hz, 2 H), 2.26 (d, J = 1.5 Hz, 3 H), 2.18 (br., 1 H) ppm; 13C NMR (125 MHz, CDCl3) δ 138.8, 136.6, 128.4, 127.9, 126.4, 125.5, 63.4, 16.0 ppm. Methyl 4-(Hydroxymethyl)benzoate (7j).21 Colorless oil. Yield: 118 mg (71%). 1H NMR (500 MHz, CDCl3) δ 8.00 (d, J = 8.0 Hz, 2 H), 7.41 (d, J = 8.0 Hz, 2 H), 4.74 (s, 2 H), 3.91 (s, 3 H), 2.52 (br., 1 H) ppm; 13C NMR (126 MHz, CDCl3) δ 167.1, 146.2, 129.8, 129.1, 126.5, 64.5, 52.2 ppm. Cinnamyl Alcohol (7k).14 Yellowish oil. Yield: 104.6 mg (78%). 1H NMR (500 MHz, CDCl3) δ 7.41−7.40 (m, 2H), 7.35−7.33 (m, 2H), 7.28−7.26 (m, 1H), 6.63 (d, J = 16 Hz, 1H), 6.97 (dt, J = 16, 5.8 Hz, 1H), 4.33 (dd, J = 5.8, 1.7 Hz, 2H), 2.49 (br, 1H) ppm; 13C NMR (126 MHz, CDCl3) δ 137.0, 131.2, 128.8, 127.9, 126.7, 63.8 ppm. 2-Thiophenylmethanol (7l).14 Yellowish oil. Yield: 107.3 mg (94%). 1H NMR (600 MHz, CDCl3) δ 7.28−7.27 (m, 1H), 7.01− 6.97 (m, 2H), 4.81 (s, 2H), 2.07 (br, 1H) ppm; 13C NMR (150 MHz, CDCl3) δ 144.3, 127.2, 125.9, 125.8, 60.2 ppm. N-Benzylaniline (8a).7a Yellowish oil. Yield: 176 mg (96%). 1H NMR (600 MHz, CDCl3) δ 7.55−7.43 (m, 4H), 7.42−7.36 (m, 1H), 7.29 (dd, J = 9.0, 7.2 Hz, 2H), 6.84 (t, J = 7.4 Hz, 1H), 6.78−6.70 (m, 2H), 4.42 (s, 2H) ppm; 13C NMR (151 MHz, CDCl3) δ 148.2, 139.5, 129.4, 128.7, 127.6, 127.3, 117.7, 113.0, 48.4 ppm. N-(4-Fluorobenzyl)aniline (8b)7f. Yellow oil. Yield: 145 mg (72%). 1 H NMR (600 MHz, CDCl3) δ 7.34 (ddd, J = 8.6, 5.4, 2.9 Hz, 2H), 7.27 (d, J = 5.3 Hz, 1H), 7.19 (td, J = 7.5, 2.7 Hz, 2H), 7.02 (td, J = 8.6, 3.1 Hz, 2H), 6.77 (t, J = 7.3 Hz, 1H), 6.68 (dd, J = 8.3, 5.1 Hz, 2H), 4.31 (s, 2H) ppm; 13C NMR (151 MHz, CDCl3) δ163.1, 161.5, 129.5, 129.4, 129.4, 118.7, 115.7, 115.5, 113.7, 48.3 ppm. N-(4-Nitrobenzyl)aniline (8c).23 Yellow oil. Yield: 180 mg (79%). 1 H NMR (600 MHz, CDCl3) δ 8.23−8.17 (m, 2H), 7.54 (d, J = 8.6 Hz, 2H), 7.24−7.15 (m, 2H), 6.77 (s, 1H), 6.61 (d, J = 8.0 Hz, 2H), 4.48 (s, 2H) ppm; 13C NMR (151 MHz, CDCl3) δ 147.3, 129.5, 128.0, 124.0, 118.8, 113.5, 100.1, 48.0 ppm. Methyl 4-((Phenylamino)methyl)benzoate (8d).24 Yellowish oil. Yield: 169 mg (70%). 1H NMR (600 MHz, CDCl3) δ 8.08−7.99 (m, 2H), 7.45 (d, J = 8.4 Hz, 2H), 7.18 (dd, J = 8.7, 7.2 Hz, 2H), 6.79− 6.70 (m, 1H), 6.66−6.57 (m, 2H), 4.41 (s, 2H), 3.92 (s, 3H) ppm; 13 C NMR (151 MHz, CDCl3) δ167.1, 147.8, 145.0, 130.0, 129.4, 129.1, 127.2, 117.9, 113.0, 52.2, 48.0 ppm.

4-Acetylbenzyl Alcohol (9).25 Colorless oil. Yield: 121.5 mg (81%). H NMR (600 MHz, CDCl3) δ 7.90 (d, J = 8.2, 2H), 7.42 (d, J = 8.2, 2H), 4.73 (s, 2H), 2.56 (s, 3H), 2.50 (br, 1H) ppm; 13C NMR (151 MHz, CDCl3) δ 198.5, 146.8, 136.5, 128.9, 126.9, 64.7, 26.9 ppm. 1



ASSOCIATED CONTENT

* Supporting Information S

The Supporting Information is available free of charge on the ACS Publications website at DOI: 10.1021/acs.joc.8b01094. Crystallographic data (CIF) Crystallographic data (CIF) Additional experimental details and copies of NMR spectra (PDF)



AUTHOR INFORMATION

Corresponding Authors

*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 donors of the American Chemical Society Petroleum Research Fund for partial support of this work (54247-UNI3). We also acknowledge the support from the CUNY Collaborative Research Incentive Program, the PSC− CUNY awards (69069-0047, 60328-0048), and the Seed grant from the Office for Advancement of Research at John Jay College. The National Science Foundation (CHE-1429086) is acknowledged for the X-ray diffractometer.



REFERENCES

(1) (a) Cho, B. T. Recent Eevelopment and Improvement for Boron Hydride-Based Catalytic Asymmetric Reduction of Unsymmetrical Ketones. Chem. Soc. Rev. 2009, 38, 443−452. (b) Togni, A.; Grützmacher, H. Catalytic Heterofunctionalization; Wiley- VCH: Weinheim, 2001. (c) Smith, M. B.; March, J. March’s Advanced Organic Chemistry; 6th ed.; Wiley-Interscience: Hoboken, NJ, 2007; pp 1703−1869. (2) Chong, C. C.; Kinjo, R. Catalytic Hydroboration of Carbonyl Derivatives, Imines, and Carbon Dioxide. ACS Catal. 2015, 5, 3238− 3259. (3) (a) Khalimon, A. Y.; Farha, P.; Kuzmina, L. G.; Nikonov, G. I. Catalytic Hydroboration by an Imido-Hydrido Complex of Mo(IV). Chem. Commun. 2012, 48, 455. (b) Blake, A. J.; Cunningham, A.; Ford, A.; Teat, S. J.; Woodward, S. Chem. - Eur. J. 2000, 6, 3586− 3594. (4) For representative examples: Ti catalysts: (a) Oluyadi, A. A.; Ma, S.; Muhoro, C. N. Titanocene(II)-Catalyzed Hydroboration of Carbonyl Compounds. Organometallics 2013, 32, 70−78. Zn: (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. Mg: (c) Arrowsmith, M.; Hadlington, T. J.; Hill, M. S.; Kociok-Kohn, G. Magnesium-catalysed Hydroboration of Aldehydes and Ketones. Chem. Commun. 2012, 48, 4567. (d) Mukherjee, D.; Ellern, A.; Sadow, A. D. MagnesiumCatalyzed Hydroboration of Esters: Evidence for a New Zwitterionic Mechanism. Chem. Sci. 2014, 5, 959−964. Co: (e) Guo, J.; Chen, J.; Lu, Z. Cobalt-Catalyzed Asymmetric Hydroboration of Aryl Ketones with PinacolBorane. Chem. Commun. 2015, 51, 5725−5727. Fe: (f) Tamang, S. R.; Findlater, M. Iron Catalyzed Hydroboration of 9447

DOI: 10.1021/acs.joc.8b01094 J. Org. Chem. 2018, 83, 9442−9448

Note

The Journal of Organic Chemistry Aldehydes and Ketones. J. Org. Chem. 2017, 82, 12857−12862. Cu: (g) Bagherzadeh, S.; Mankad, N. P. Extremely Efficient Hydroboration of Ketones and Aldehydes by Copper Carbene Catalysis. Chem. Commun. 2016, 52, 3844−3846. Al: (h) 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. (i) Jakhar, V. K.; Barman, M. K.; Nembenna, S. Aluminum Monohydride Catalyzed Selective Hydroboration of Carbonyl Compounds. Org. Lett. 2016, 18, 4710−4713. (5) Friedfeld, M. R.; Shevlin, M.; Hoyt, J. M.; Krska, S. W.; Tudge, M. T.; Chirik, P. J. Cobalt Precursors for High-Throughput Discovery of Base Metal Asymmetric Alkene Hydrogenation Catalysts. Science 2013, 342, 1076−1080. (6) Zhang, G.; Hanson, S. K. Cobalt-Catalyzed Acceptorless Alcohol Dehydrogenation: Synthesis of Imines from Alcohols and Amines. Org. Lett. 2013, 15, 650−653. (7) (a) 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. (b) 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. (c) Yin, Z.; Zeng, H.; Wu, J.; Zheng, S.; Zhang, G. CobaltCatalyzed Synthesis of Aromatic, Aliphatic, and Cyclic Secondary Amines via a “Hydrogen-Borrowing” Strategy. ACS Catal. 2016, 6, 6546−6550. (d) Zhang, G.; Yin, Z.; Zheng, S. Cobalt-Catalyzed NAlkylation of Amines with Alcohols. Org. Lett. 2016, 18, 300−303. (e) Zhang, G.; Wu, J.; Zeng, H.; Zhang, S.; Yin, Z.; Zheng, S. CobaltCatalyzed α-Alkylation of Ketones with Primary Alcohols. Org. Lett. 2017, 19, 1080−1083. (8) (a) Palmer, W. N.; Diao, T.; Pappas, I.; Chirik, P. J. HighActivity Cobalt Catalysts for Alkene Hydroboration with Electronically Responsive Terpyridine and α-Diimine Ligands. ACS Catal. 2015, 5, 622−626. (b) Zhang, L.; Zuo, Z.; Wan, X.; Huang, Z. Cobalt-Catalyzed Enantioselective Hydroboration of 1,1-Disubstituted Aryl Alkenes. J. Am. Chem. Soc. 2014, 136, 15501−15504. (c) 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. (d) Peng, J.; Docherty, J. H.; Dominey, A. P.; Thomas, S. P. Chem. Commun. 2017, 53, 4726−4729. (e) 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. (9) (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) 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, 19107−19110. (c) Scheuermann, M. L.; Johnson, E. J.; Chirik, P. J. Alkene Isomerization−Hydroboration Promoted by Phosphine-Ligated Cobalt Catalysts. Org. Lett. 2015, 17, 2716−2719. (d) 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. (10) 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. (11) 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. (12) Huang, Z.; Liu, D.; Camacho-Bunquin, J.; Zhang, G.; Yang, D.; López-Encarnación, J. M.; Xu, Y.; Ferrandon, J. K.; Niklas, J.; 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. (13) (a) Zhang, L.; Zuo, Z.; Leng, X.; Huang, Z. A Cobalt-Catalyzed Alkene Hydroboration with Pinacolborane. Angew. Chem., Int. Ed. 2014, 53, 2696−2700. (b) Zhang, L.; Zuo, Z.; Wan, X.; Huang, Z. Cobalt-Catalyzed Enantioselective Hydroboration of 1,1-Disubstituted Aryl Alkenes. J. Am. Chem. Soc. 2014, 136, 15501−15504. (14) 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. (15) 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. (16) Beves, J. E.; Constable, E. C.; Housecroft, C. E.; Kepert, C.; Price, D. J. The First Example of a Coordination Polymer from the Expanded 4,4′-Bipyridine Ligand [Ru(pytpy)2]2+ (pytpy = 4′-(4pyridyl)-2,2′:6′,2″-terpyridine). CrystEngComm 2007, 9, 456−459. (17) Constable, E. C.; Zhang, G.; Housecroft, C. E.; Zampese, J. A. Zinc(II) Coordination Polymers, Metallohexacycles and Metallocapsules-Do We Understand Self-assembly in Metallosupramolecular Chemistry: Algorithms or Serendipity? CrystEngComm 2011, 13, 6864−6870. (18) (a) Housecroft, C. E. 4,2’:6’,4”-Terpyridines: Diverging and Diverse building blocks in Coordination Polymers and Metallomacrocycles. Dalton Trans. 2014, 43, 6594−6604. (b) Housecroft, C. E. Divergent 4,2’:6’,4”- and 3,2’:6’,3”-Terpyridines as Linkers in 2- and 3Dimensional Architectures. CrystEngComm 2015, 17, 7461−7468. (19) (a) Kaithal, A.; Chatterjee, B.; Gunanathan, C. Ruthenium Catalyzed Selective Hydroboration of Carbonyl Compounds. Org. Lett. 2015, 17, 4790−4793. (b) Weidner, V. L.; Barger, C. J.; Delferro, M.; Lohr, T. L.; Marks, T. J. Rapid, Mild, and Selective Ketone and Aldehyde Hydroboration/Reduction Mediated by a Simple Lanthanide Catalyst. ACS Catal. 2017, 7, 1244−1247. (c) Chen, S.; Yan, D.; Xue, M.; Hong, Y.; Yao, Y.; Shen, Q. Tris(cyclopentadienyl)lanthanide Complexes as Catalysts for Hydroboration Reaction toward Aldehydes and Ketones. Org. Lett. 2017, 19, 3382−3385. (20) Ruddy, A. J.; Kelly, C. M.; Crawford, S. M.; Wheaton, C. A.; Sydora, O. L.; Small, B. L.; Stradiotto, M.; Turculet, L. (NPhosphinoamidinate)Iron Pre-Catalysts for the Room Temperature Hydrosilylation of Carbonyl Compounds with Broad Substrate Scope at Low Loadings. Organometallics 2013, 32, 5581−5588. (21) Wang, R.; Tang, Y.; Xu, M.; Meng, C.; Li, F. Transfer Hydrogenation of Aldehydes and Ketones with Isopropanol under Neutral Conditions Catalyzed by a Metal−Ligand Bifunctional Catalyst [Cp*Ir(2,2′-bpyO)(H2O)]. J. Org. Chem. 2018, 83, 2274− 2281. (22) Huynh, H. V.; Yeo, C. H.; Chew, Y. X. Syntheses, Structures, and Catalytic Activities of Hemilabile Thioether-Functionalized NHC Complexes. Organometallics 2010, 29, 1479−1486. (23) Bisai, M. K.; Pahar, S.; Das, T.; Vanka, K.; Sen, S. S. Transition Metal Free Catalytic Hydroboration of Aldehydes and Aldimines by Amidinato Silane. Dalton Trans. 2017, 46, 2420−2424. (24) Kato, T.; Matsuoka, S.; Suzuki, M. Transfer Hydrogenation Promoted by N-heterocyclic Carbene and Water. Chem. Commun. 2015, 51, 13906−13909. (25) Mazza, S.; Scopelliti, R.; Hu, X. Chemoselective Hydrogenation and Transfer Hydrogenation of Aldehydes Catalyzed by Iron(II) PONOP Pincer Complexes. Organometallics 2015, 34, 1538−1545.

9448

DOI: 10.1021/acs.joc.8b01094 J. Org. Chem. 2018, 83, 9442−9448