You are on page 1of 6

Note

Cite This: J. Org. Chem. 2019, 84, 4429−4434 pubs.acs.org/joc

Asymmetric Synthesis of γ‑Hydroxy Pinacolboronates through


Copper-Catalyzed Enantioselective Hydroboration of α,β-
Unsaturated Aldehydes
Won Jun Jang,† Seung Min Song,† Yeji Park, and Jaesook Yun*
Department of Chemistry, Sungkyunkwan University, Suwon 16419, Korea
*
S Supporting Information
See https://pubs.acs.org/sharingguidelines for options on how to legitimately share published articles.

ABSTRACT: We report a copper-catalyzed enantioselective


hydroboration of α,β-unsaturated aldehydes with pinacolbor-
ane. α,β-Unsaturated aldehydes were converted to the
corresponding γ-pinacolboronate alcohols in good yields and
Downloaded via UNIV D'ORLEANS on March 15, 2022 at 08:06:40 (UTC).

enantioselectivities through consecutive hydroboration of the


CO and CC bonds. This process provides simple access
to the hydroborated product of allylic alcohols, and the
resulting γ-pinacolboronate alcohols could be utilized in
various transformations.

T he hydroboration of alkenes is one of the fundamental


methods to produce organoboron compounds, which are
recognized as versatile synthetic intermediates in organic
Scheme 1. Asymmetric Hydroboration Reactions of Allylic
Alcohol Derivatives

synthesis.1 In recent decades, there have been significant


advances in the transition-metal-catalyzed hydroboration field,
allowing highly regio- and enantioselective syntheses of chiral
alkylboron compounds.2 Moreover, a range of stereospecific
transformations of carbon−boron bonds to other function-
alities have increased the synthetic utility of organoboranes,
which has further encouraged investigations on the efficient
synthesis of chiral organoboron compounds.3
Controlling regioselectivity and enantioselectivity in the
asymmetric hydroboration of terminal olefins has been an
important issue, as most transition-metal catalytic systems
produce anti-Markovnikov addition products, linear products
with no chiral center, or a mixture of products.4 While
regioselective hydroboration of internal alkenes is more
challenging than that of terminal alkenes, our group previously
disclosed that copper(I)−bisphosphine catalysts could hydro-
borate vinyl arenes and disubstituted internal alkenes with high
regio- and enantioselectivities with pinacolborane as the
hydroborating reagent.5 In particular, copper(I) chloride and
DTBM-Segphos (L*) as the chiral ligand was the most
effective catalytic combination for the asymmetric hydro-
boration of both terminal and internal alkenes, producing
secondary alkylboronates with high enantioselectivity.5b,c
Studies on the asymmetric hydroboration of internal allylic chiral copper(I)−bisphosphine complexes, we decided to
alcohols are rare, and only a few examples of hydroborations of investigate this catalytic system in the asymmetric hydro-
protected terminal allylic alcohols have been reported in Rh- boration of internal allylic alcohols. With the copper-(R)-
catalyzed directed hydroborations6 and hydroboration of an
allylic ester with a copper catalyst5b (Scheme 1, A). Based on Received: November 29, 2018
our copper-catalyzed enantioselective hydroboration with Published: March 14, 2019

© 2019 American Chemical Society 4429 DOI: 10.1021/acs.joc.8b03045


J. Org. Chem. 2019, 84, 4429−4434
The Journal of Organic Chemistry Note

Scheme 2. Copper-Catalyzed Enantioselective Hydroborationa

a
General reaction conditions: 1 (0.5 mmol), pinacolborane (1.25 mmol), CuCl (0.025 mmol), ligand L* ((R)-DTBM-Segphos) (0.0275 mmol),
NaOt-Bu (0.05 mmol) in toluene (1 mL); see the Experimental Section for details. bIsolated yield. cDetermined by chiral HPLC analysis of the
corresponding diol 3.

DTBM-Segphos catalyst, however, hydroboration reaction of and 2j) were not very stable on SiO2, and thus, the
cinnamyl alcohol failed to give the desired product with no corresponding diols were obtained after oxidation. This
conversion (Scheme 1, B). We thought this problem could be protocol was not suitable for α,β-unsaturated aldehydes with
circumvented by using α,β-unsaturated aldehydes, hoping that β-alkyl substituents either, as the reaction produced fully
1,2-hydroboration of the aldehyde carbonyl would occur faster reduced, nonborylated alcohol products.8
than 1,4-hydroboration. Herein, we describe a copper- The hydroboration protocol was suitable for large-scale
catalyzed enantioselective hydroboration of α,β-unsaturated
synthesis (Scheme 3a). A 1.0 mol % loading of copper catalyst
aldehydes with pinacolborane to produce γ-pinacolboronate
alcohols. was sufficient to perform the hydroboration reaction on a 3.0
We started our investigation by examining the hydro-
boration of cinnamaldehyde 1a with copper catalyst (Scheme Scheme 3. Gram-Scale Synthesis and Transformation of
2). The reaction proceeded smoothly to afford target product Chiral Boron Products
2a in good yield and good enantioselectivity in toluene at 60
°C.7 Then, we investigated the substrate scope of cinnamalde-
hyde derivatives. Most of the desired products were obtained
in good yields and good enantioselectivities. A naphthyl moiety
(2b) was accommodated with no problem. Substrates with an
electron-donating or electron-withdrawing substituent at the
para position afforded the desired products (2c and 2d) in
good yields and enantioselectivities. The hydroboration
reaction was amenable to enal substrates with a methyl or
halogen substituent at the ortho- or meta-position of the aryl
group (1e−1h). Heteroaryl compounds (1i and 1j) such as 2-
benzofuranyl and 2-thienyl reacted with good conversion to
the desired products, but with slightly decreased enantiose-
lectivities. However, γ-hydroxy pinacolboronate products with
ortho-substitution (2g and 2h) or heteroaryl substituent (2i
4430 DOI: 10.1021/acs.joc.8b03045
J. Org. Chem. 2019, 84, 4429−4434
The Journal of Organic Chemistry Note

mmol scale of 1a. Despite the reduced amounts of catalyst and effective in this transformation, furnishing useful chiral
ligand, 2a was produced in 82% yield and 90% ee. Next, the hydroxyboronates with high enantioselectivities via consecutive
chiral boronates (2a) were stereospecifically converted to hydroboration of CO and CC bonds. Moreover, the
compounds containing a new C−C bond (Scheme 3b). Suzuki protocol could be easily scaled up to gram scale, and
coupling9 of 2a with 4-bromotoluene stereospecifically stereospecific transformations of the resulting hydroxyboro-
produced the arylated product 410 in the presence of a nates to various compounds demonstrate their synthetic utility.


palladium catalyst. Both reactions occurred in high yields and
with complete conservation of the enantiomeric excess of the EXPERIMENTAL SECTION
original hydroborated products. General Methods. All reactions were carried out under a nitrogen
To investigate the mechanism of the hydroboration, an atmosphere using oven-dried glassware. CuCl, NaOt-Bu, pinacolbor-
NMR study was carried out with 1.1 equiv of pinacolborane to ane, and other commercial reagents were purchased from Aldrich and
detect a reaction intermediate (Scheme 4). The 1,2-hydro- were used as received. (R)-DTBM-Segphos was purchased from TCI.
Unsaturated aldehydes 1b−1j were prepared following literature
Scheme 4. NMR Study of Chemoselective Hydroboration of procedures.11 All liquid aldehydes were distilled from calcium hydride,
Cinnamaldehyde and solid aldehydes were purified by recrystallization. Deactivated
silica gel was prepared by stirring a slurry of the silica gel in a 5 mol %
NaOAc aqueous solution for 30 min. Toluene was purified using the
PureSolv solvent purification system, from Innovative Technology,
Inc. All 1H NMR spectra were obtained on Bruker at 500 systems and
reported in parts per million (ppm) downfield from tetramethylsilane.
13
C NMR spectra were reported in ppm referenced to deuteriochloro-
form (77.16 ppm). 11B NMR spetra were obtained on Bruker at 400
systems at Kyonggi University (Suwon, Korea). High-performance
liquid chromatography (HPLC) was performed using Younglin Acme
9100 series. Infrared spectra (IR) were obtained on Nicolet 205 FT-
IR and were recorded in cm−1. High-resolution mass spectra (HRMS)
were obtained by the electrospray ionization (ESI) method using an
ion trap mass analyzer at Sogang Center for Research Facilities of
boration of the aldehyde carbonyl occurred faster than Sogang University (Seoul, Korea) and reported in the form of m/z
conjugate addition, resulting in the formation of II, which (intensity relative to peak = 100).
was chemically compatible to the yield of 2a by reaction with General Procedure for the Copper-Catalyzed Enantioselec-
HBpin. tive Hydroboration of α,β-Unsaturated Aldehydes. A mixture
A proposed catalytic cycle for the hydroboration of enals is of CuCl (5 mol %, 0.025 mmol), (R)-DTBM-Segphos (5.5 mol %,
0.0275 mmol), and NaOt-Bu (10 mol %, 0.05 mmol) in toluene (0.5
shown in Scheme 5. The active catalyst L*Cu−H reacts with mL) was stirred for 5 min in a Schlenk tube under an atmosphere of
nitrogen. Pinacolborane (2.5 equiv, 1.25 mmol) was added to the
Scheme 5. Proposed Catalytic Cycle reaction mixture, and stirring was continued for another 15 min at
room temperature. Substrate 1 (1 equiv, 0.5 mmol) dissolved in
toluene (0.5 mL) was added to the reaction mixture. The reaction
mixture was stirred at 60 °C and monitored by TLC. Upon
completion of the reaction, the reaction mixture was diluted with
ethyl acetate (10 mL) and water (10 mL). The aqueous layer was
extracted with ethyl acetate, and the combined organic layers were
dried over Na2SO4 and concentrated in vacuo. The product was
purified by deactivated silica gel chromatography. Rapid silica gel
chromatography (5 min) is imperative for reproducible results
because the γ-pinacolboronate alcohol decomposes readily on silica
gel.
For determination of enantiomeric excess (ee), the γ-pinacolbor-
onate alcohol was transformed to the corresponding diol by the
following procedure: To a solution of γ-pinacolboronate alcohol (1
equiv) in THF:H2O (2 mL, 1:1 (v/v)) was added sodium perborate
(3 equiv). The reaction mixture was stirred for 6 h at room
temperature. The reaction was quenched with water and extracted
with ethyl acetate. The product was purified by silica gel
chromatography.
General Procedure for the Gram-Scale Synthesis of γ-
Hydroxy Pinacolboronates. A mixture of CuCl (1 mol %, 0.03
mmol), (R)-DTBM-Segphos (1.1 mol %, 0.033 mmol), and NaOt-Bu
(2 mol %, 0.06 mmol) in toluene (0.5 mL) was stirred for 5 min in a
the aldehyde to form the copper alkoxide I, which Schlenk tube under an atmosphere of nitrogen. Pinacolborane (2.5
subsequently reacts with HBpin to generate II and the active equiv, 7.5 mmol) was added to the reaction mixture, and stirring was
L*Cu−H catalyst. Further hydroboration of intermediate II continued for another 15 min at room temperature. Substrate 1 (1
equiv, 3 mmol) dissolved in toluene (0.5 mL) was added to the
with the catalyst forms an alkylcopper intermediate III, which reaction mixture. The reaction mixture was stirred at 60 °C and
reacts with HBpin to form the final borylated product. monitored by TLC. Upon completion of the reaction, the reaction
In summary, we have developed a copper-catalyzed mixture was diluted with ethyl acetate (10 mL) and water (10 mL).
enantioselective hydroboration of α,β-unsaturated aldehydes The aqueous layer was extracted with ethyl acetate, and the combined
with pinacolborane. The DTBM-Segphos−copper catalyst was organic layers were dried over Na2SO4 and concentrated in vacuo. The

4431 DOI: 10.1021/acs.joc.8b03045


J. Org. Chem. 2019, 84, 4429−4434
The Journal of Organic Chemistry Note

product was purified by deactivated silica gel chromatography. Rapid 7.03−7.01 (m, 2H), 6.97−6.95 (m, 1H), 3.71−3.65 (m, 1H), 3.60−
silica gel chromatography (5 min) is imperative for reproducible 3.54 (m, 1H), 2.40 (t, J = 7.5 Hz, 1H), 2.31 (s, 3H), 2.15−2.07 (m,
results because the γ-pinacolboronate alcohol decomposes readily on 1H), 1.94−1.88 (m, 1H), 1.54 (brs, 1H), 1.21 (s, 6H), 1.20 (s, 6H).
silica gel. 13
C{1H} NMR (125 MHz, CDCl3) δ 142.6, 137.9, 129.3, 128.3,
(S)-3-Phenyl-3-(4,4,5,5-tetramethyl-1,3,2-dioxaborolan-2-yl)- 126.2, 125.4, 83.5, 62.5, 35.7, 24.6, 21.5. The carbon bonded to boron
propan-1-ol (2a). Following the general procedure, 2a was obtained was not detected due to quadrupolar relaxation. IR (neat) 3347, 2968,
in 84% yield (110.3 mg, 0.42 mmol, colorless oil). Rf = 0.4 (EtOAc/ 1467, 1371, 1311, 1134, 1008 cm−1. HRMS (ESI-Ion Trap) m/z [M
hexane = 1/1). 1H NMR (500 MHz, CDCl3) δ 7.28−7.21 (m, 4H), + H]+ calcd for C16H26BO3 277.1975, found 277.1976. 91% ee was
7.16−7.13 (m, 1H), 3.71−3.64 (m, 1H), 3.63−3.55 (m, 1H), 2.44 (t, measured by chiral HPLC on an AD-H column with the
J = 7.5 Hz, 1H), 2.16−2.09 (m, 1H), 1.95−1.89 (m, 1H), 1.51 (brs, corresponding diol obtained after oxidation (i-PrOH:hexanes =
1H), 1.21 (s, 6H), 1.19 (s, 6H); 13C{1H} NMR (125 MHz, CDCl3) δ 5:95, 0.5 mL/min); tR = 30.01 min (major), tR = 28.29 min (minor).
142.7, 128.41, 128.40, 125.4, 83.5, 62.5, 35.6, 24.61, 24.58. The (S)-3-(3-Fluorophenyl)-3-(4,4,5,5-tetramethyl-1,3,2-dioxaboro-
carbon bonded to boron was not detected due to quadrupolar lan-2-yl)propan-1-ol (2f). Following the general procedure, 2f was
relaxation. 11B{1H} NMR (128 MHz, CDCl3) δ 33.9. IR (neat) 3437, obtained in 81% yield (113.6 mg, 0.41 mmol, colorless oil). Rf = 0.4
2977, 1458, 1382, 1324, 1144 cm−1. HRMS (ESI-Ion Trap) m/z [M (EtOAc/hexane = 1/1). 1H NMR (500 MHz, CDCl3) δ 7.23−7.19
+ H]+ calcd for C15H24BO3 263.1819, found 263.1820. 90% ee was (m, 1H), 6.99−6.93 (m, 2H), 6.85−6.82 (m, 1H), 3.71−3.65 (m,
measured by chiral HPLC on an AD-H column with the 1H), 3.60−3.55 (m, 1H), 2.46 (t, J = 7.5 Hz, 1H), 2.15−2.08 (m,
corresponding diol obtained after oxidation (i-PrOH:hexanes = 1H), 1.93−1.87 (m, 1H), 1.53 (s, 1H), 1.21 (s, 6H), 1.19 (s, 6H).
5:95, 0.5 mL/min); tR = 29.70 min (major), tR = 27.97 min (minor). 13
C{1H} NMR (125 MHz, CDCl3) δ 163.0 (d, J = 244 Hz), 145.4 (d,
(S)-3-(Naphthalen-2-yl)-3-(4,4,5,5-tetramethyl-1,3,2-dioxaboro- J = 6.3 Hz), 129.7 (d, J = 8.8 Hz), 124.1 (d, J = 2.5 Hz), 115.1 (d, J =
lan-2-yl)propan-1-ol (2b). Following the general procedure, 2b was 21.3 Hz), 112.3 (d, J = 21.3 Hz), 83.7, 62.2, 35.4, 24.59, 24.56. The
obtained in 78% yield (121.6 mg, 0.39 mmol, colorless oil). Rf = 0.4 carbon bonded to boron was not detected due to quadrupolar
(EtOAc/hexane = 1/1). 1H NMR (500 MHz, CDCl3) δ 7.79−7.74 relaxation. IR (neat) 3460, 2965, 1455, 1383, 1321, 1118 cm−1.
(m, 3H), 7.66 (s, 1H), 7.45−7.38 (m, 3H), 3.74−3.69 (m, 1H), HRMS (ESI-Ion Trap) m/z [M + H]+ calcd for C15H23BFO3
3.66−3.60 (m, 1H), 2.63 (t, J = 7.5 Hz, 1H), 2.28−2.19 (m, 1H), 281.1724, found 281.1722. 88% ee was measured by chiral HPLC
2.05−1.99 (m, 1H), 1.56 (brs, 1H), 1.21 (s, 6H), 1.19 (s, 6H). on an AS-H column with the corresponding diol obtained after
13
C{1H} NMR (125 MHz, CDCl3) δ 140.3, 133.8, 131.8, 127.9, oxidation (i-PrOH:hexanes = 10:90, 0.5 mL/min); tR = 16.33 min
127.6, 127.5, 127.4, 126.3, 125.8, 125.0, 83.6, 62.5, 35.4, 24.62, 24.60. (major), tR = 18.92 min (minor).
The carbon bonded to boron was not detected due to quadrupolar (S)-1-(2-Chlorophenyl)propane-1,3-diol (3g). Following the gen-
relaxation. IR (neat) 3474, 3052, 2970, 1457, 1390, 1314, 1147 cm−1. eral procedure, 3g was obtained in 74% yield (69.3 mg, 0.37 mmol,
HRMS (ESI-Ion Trap) m/z [M + H]+ calcd for C19H26BO3 313.1975, colorless oil). The characterization data for 3g were concordant with
found 313.1977. 90% ee was measured by chiral HPLC on an AS-H that previously reported in the literature.12a Rf = 0.2 (EtOAc/hexane
column with the corresponding diol obtained after oxidation (i- = 1/1). 1H NMR (500 MHz, CDCl3) δ 7.65−7.63 (m, 1H), 7.33−
PrOH:hexanes = 10:90, 0.5 mL/min); tR = 24.60 min (major), tR = 7.30 (m, 2H), 7.23−7.20 (m, 1H), 5.38−3.36 (m, 1H), 3.96−3.89
27.20 min (minor). (m, 2H), 3.09 (brs, 1H), 2.28 (brs, 1H), 2.08−2.04 (m, 1H), 1.97−
(S)-3-(4-Methoxyphenyl)-3-(4,4,5,5-tetramethyl-1,3,2-dioxaboro- 1.89 (m, 1H). 13C{1H} NMR (125 MHz, CDCl3) δ 141.5, 131.4,
lan-2-yl)propan-1-ol (2c). Following the general procedure, 2c was 129.4, 128.5, 127.2, 127.1, 71.2, 61.8, 38.5. 87% ee was measured by
obtained in 83% yield (121.0 mg, 0.41 mmol, colorless oil). Rf = 0.4 chiral HPLC on an OJ-H column (i-PrOH:hexanes = 5:95, 0.5 mL/
(EtOAc/hexane = 1/1). 1H NMR (500 MHz, CDCl3) δ 7.14−7.12 min); tR = 47.78 min (major), tR = 53.22 min (minor).
(m, 2H), 6.82−6.80 (m, 2H), 3.78 (s, 3H), 3.69−3.63 (m, 1H), (S)-1-(o-Tolyl)propane-1,3-diol (3h). Following the general
3.61−3.55 (m, 1H), 2.38 (t, J = 7.5 Hz, 1H), 2.11−2.04 (m, 1H), procedure, 3h was obtained in 79% yield (61.7 mg, 0.37 mmol,
1.91−1.85 (m, 1H), 1.50 (brs, 1H), 1.21 (s, 6H), 1.19 (s, 6H). colorless oil). The characterization data for 3h were concordant with
13
C{1H} NMR (125 MHz, CDCl3) δ 157.5, 134.6, 129.3, 113.9, 83.5, that previously reported in the literature.12b Rf = 0.2 (EtOAc/hexane
62.4, 55.2, 35.9, 24.63, 24.59. The carbon bonded to boron was not = 1/1). 1H NMR (500 MHz, CDCl3) δ 7.54−7.53 (m, 1H), 7.25−
detected due to quadrupolar relaxation. IR (neat) 3401, 2978, 1608, 7.24 (m, 1H), 7.20−7.13 (m, 2H), 5.22−5.20 (m, 1H), 3.95−3.85
1511, 1447, 1380, 1324, 1216, 1141 cm−1. HRMS (ESI-Ion Trap) m/ (m, 2H), 2.55 (brs, 1H), 2.38 (brs, 1H), 2.33 (s, 3H), 2.01−1.90 (m,
z [M + H]+ calcd for C16H26BO4 293.1924, found 293.1926. 91% ee 1H). 13C{1H} NMR (125 MHz, CDCl3) δ 142.3, 134.1, 130.5, 127.3,
was measured by chiral HPLC on an AD-H column with the 126.4, 125.2, 71.0, 61.8, 39.2, 18.9. IR (neat) 3397, 2960, 1487, 1380,
corresponding diol obtained after oxidation (i-PrOH:hexanes = 5:95, 1326, 1124 cm−1. 90% ee was measured by chiral HPLC on an AD-H
0.5 mL/min); tR = 43.42 min (major), tR = 41.52 min (minor). column (i-PrOH:hexanes = 5:95, 0.5 mL/min); tR = 44.90 min
(S)-3-(4,4,5,5-Tetramethyl-1,3,2-dioxaborolan-2-yl)-3-(4- (major), tR = 38.59 min (minor).
(trifluoromethyl)phenyl)propan-1-ol (2d). Following the general (S)-1-(Benzofuran-2-yl)propane-1,3-diol (3i). Following the gen-
procedure, 2d was obtained in 82% yield (140.3 mg, 0.43 mmol, eral procedure, 3i was obtained in 82% yield (78.6 mg, 0.41 mmol,
colorless oil). Rf = 0.4 (EtOAc/hexane = 1/1). 1H NMR (500 MHz, colorless oil). The characterization data for 3i were concordant with
CDCl3) δ 7.52−7.50 (m, 2H), 7.34−7.32 (m, 2H), 3.70−3.65 (m, that previously reported in the literature.12b Rf = 0.2 (EtOAc/hexane
1H), 3.60−3.55 (m, 1H), 2.53 (t, J = 7.5 Hz, 1H), 2.19−2.11 (m, = 1/1). 1H NMR (500 MHz, CDCl3) δ 7.55 (d, J = 7.5 Hz, 1H), 7.45
1H), 1.94−1.88 (m, 1H), 1.54 (brs, 1H), 1.20 (s, 6H), 1.19 (s, 6H). (d, J = 8.0 Hz, 1H), 7.28−7.25 (m, 1H), 7.23−7.20 (m, 1H), 6.67 (s,
13
C{1H} NMR (125 MHz, CDCl3) δ 147.2, 128.6, 127.7 (q, J = 30 1H), 5.14−5.11 (m, 1H), 3.97−3.90 (m, 2H), 3.12 (brs, 1H), 2.22−
Hz), 125.3 (q, J = 3.8 Hz), 124.5 (q, J = 270 Hz), 83.8, 62.1, 35.3, 2.18 (m, 2H), 2.16 (brs, 1H). 13C{1H} NMR (125 MHz, CDCl3) δ
24.60, 24.57. The carbon bonded to boron was not detected due to 159.0, 154.8, 128.1, 124.2, 122.8, 121.1, 111.2, 102.6, 68.0, 61.0, 36.9.
quadrupolar relaxation. IR (neat) 3444, 2958, 1445, 1381, 1325, 1124, 87% ee was measured by chiral HPLC on an AS-H column (i-
1015 cm−1. HRMS (ESI-Ion Trap) m/z [M + H]+ calcd for PrOH:hexanes = 5:95, 0.5 mL/min); tR = 29.30 min (major), tR =
C16H23BF3O3 331.1692, found 331.1690. 91% ee was measured by 34.97 min (minor).
chiral HPLC on an OD-H column with the corresponding diol (S)-1-(Thiophen-2-yl)propane-1,3-diol (3j). Following the general
obtained after oxidation (i-PrOH:hexanes = 10:90, 0.5 mL/min); tR = procedure, 3j was obtained in 85% yield (67.1 mg, 0.42 mmol,
13.85 min (major), tR = 6.45 min (minor). colorless oil). The characterization data for 3j were concordant with
(S)-3-(4,4,5,5-Tetramethyl-1,3,2-dioxaborolan-2-yl)-3-(m-tolyl)- that previously reported in the literature.12c Rf = 0.2 (EtOAc/hexane
propan-1-ol (2e). Following the general procedure, 2e was obtained = 1/1). 1H NMR (500 MHz, CDCl3) δ 7.26−7.25 (m, 1H), 6.99−
in 80% yield (110.7 mg, 0.40 mmol, colorless oil). Rf = 0.4 (EtOAc/ 6.97 (m, 2H), 5.24−5.21 (m, 1H), 3.94−3.86 (m, 2H), 2.97 (brs,
hexane = 1/1). 1H NMR (500 MHz, CDCl3) δ 7.16−7.14 (m, 1H), 1H), 2.19 (brs, 1H), 2.16−2.04 (m, 2H). 13C{1H} NMR (125 MHz,

4432 DOI: 10.1021/acs.joc.8b03045


J. Org. Chem. 2019, 84, 4429−4434
The Journal of Organic Chemistry Note

CDCl3) δ 148.3, 126.7, 124.6, 123.5, 70.1, 61.2, 40.7. 88% ee was
measured by chiral HPLC on an AS-H column (i-PrOH:hexanes =
5:95, 0.5 mL/min); tR = 45.76 min (major), tR = 38.82 min (minor).
■ REFERENCES
(1) For reviews, see: (a) Organic Synthesis via Boranes; Brown, H. C.,
(S)-3-Phenyl-3-(p-tolyl)propan-1-ol (4). A mixture of Pd2(dba)3 (1 Ed.; Wiley Interscience: New York, 1975. (b) Burgess, K.; Ohlmeyer,
mol %, 0.005 mmol), RuPhos (2 mol %, 0.010 mmol), and KOH (3 M. J. Transition-Metal-Promoted Hydroborations of Alkenes,
equiv, 1.5 mmol) in toluene (1 mL) was stirred for 15 min under an Emerging Methodology or Organic Transformations. Chem. Rev.
atmosphere of nitrogen. 2a (1 equiv, 0.5 mmol), 4-bromotoluene (1.5 1991, 91, 1179−1191. (c) Brown, H. C.; Ramachandran, P. V.
equiv, 0.75 mmol), THF (1 mL), and H2O (0.2 mL) were added. The Versatile α-Pinene-Based Borane Reagents for Asymmetric Syntheses.
reaction mixture was stirred for 24 h at 70 °C. Upon completion of J. Organomet. Chem. 1995, 500, 1−19.
the reaction, the reaction mixture was quenched with water and (2) For a review, see: (a) Crudden, C. M.; Edwards, D. Catalytic
extracted with ethyl acetate. The combined organic layers were dried Asymmetric Hydroboration: Recent Advances and Applications in
over MgSO4 and concentrated in vacuo. The product was purified by Carbon−Carbon Bond-Forming Reactions. Eur. J. Org. Chem. 2003,
silica gel chromatography, and 4 was obtained in 80% yield (90.4 mg, 2003, 4695−4712. (b) Carroll, A.-M.; O’Sullivan, T. P.; Guiry, P. J.
0.40 mmol, colorless oil). The characterization data for 3j were The Development of Enantioselective Rhodium-Catalysed Hydro-
concordant with that previously reported in the literature.10 Rf = 0.3 boration of Olefins. Adv. Synth. Catal. 2005, 347, 609−631.
(EtOAc/hexane = 1/5). 1H NMR (500 MHz, CDCl3) δ 7.29−7.24 (3) (a) Contemporary Boron Chemistry; Davidson, M. G., Hughes, A.
(m, 4H), 7.19−7.14 (m, 3H), 7.10−7.09 (m, 2H), 4.10 (t, J = 8.0 Hz, K., Marder, T. B., Wade, K., Eds.; Royal Society of Chemistry:
1H), 3.63−3.60 (m, 2H), 2.33−2.29 (m, 5H), 1.21 (brs, 1H). Cambridge, 2000. (b) Boronic Acids: Preparation and Applications in
13
C{1H} NMR (125 MHz, CDCl3) δ 144.8, 141.5, 135.8, 129.3, Organic Synthesis, Medicine and Materials, 2nd ed.; Hall, D. G., Ed.;
128.5, 127.8, 127.7, 126.2, 61.2, 47.0, 38.3, 21.0. [α]D20 +10.2 (c 0.52, Wiley-VCH: Weinheim, 2011. (c) Crudden, C. M.; Glasspoole, B. W.;
CHCl3); [lit.10 (93% ee): [α]D20 +4.3 (c 1.24, CHCl3)]. 90% ee was Lata, C. J. Expanding the Scope of Transformations of Organoboron
measured by chiral HPLC on an AS-H column (i-PrOH:hexanes = Species: Carbon−Carbon Bond Formation with Retention of
2:98, 1.0 mL/min); tR = 21.59 min (major), tR = 18.99 min (minor). Configuration. Chem. Commun. 2009, 6704−6716. (d) Scott, H. K.;
2-(Cinnamyloxy)-4,4,5,5-tetramethyl-1,3,2-dioxaborolane (II). A Aggarwal, V. K. Highly Enantioselective Synthesis of Tertiary Boronic
mixture of CuCl (5 mol %, 0.025 mmol), (R)-DTBM-Segphos (5.5 Esters and their Stereospecific Conversion to other Functional
mol %, 0.0275 mmol), and NaOt-Bu (10 mol %, 0.05 mmol) in Groups and Quaternary Stereocentres. Chem. - Eur. J. 2011, 17,
benzene-d (0.5 mL) was stirred for 5 min in a Schlenk tube under an 13124−13132. (e) Leonori, D.; Aggarwal, V. K. Stereospecific
atmosphere of nitrogen. Pinacolborane (1.1 equiv, 0.55 mmol) was Couplings of Secondary and Tertiary Boronic Esters. Angew. Chem.,
added to the reaction mixture, and stirring was continued for another Int. Ed. 2015, 54, 1082−1096. (f) Sandford, C.; Aggarwal, V. K.
15 min at room temperature. Substrate 1 dissolved in benzene-d (0.5 Stereospecific Functionalizations and Transformations of Secondary
mL) was added to the reaction mixture. The reaction mixture was and Tertiary Boronic Esters. Chem. Commun. 2017, 53, 5481−5494.
stirred at 60 °C for 1 h. Aliquot was taken by removing a small (4) (a) Morrill, T. C.; D’Souza, C. A.; Yang, L.; Sampognaro, A. J.
amount of the reaction solution and passing it through a filter, eluting Transition-Metal-Promoted Hydroboration of Alkenes: A Unique
with benzene-d. The characterization data for II were concordant with Reversal of Regioselectivity. J. Org. Chem. 2002, 67, 2481−2484.
that previously reported in the literature.13 1H NMR (500 MHz, (b) Yamamoto, Y.; Fujikawa, R.; Umemoto, T.; Miyaura, N. Iridium-
C6D6) δ 7.18−7.16 (m, 2H), 7.09−7.06 (m, 2H), 7.03−7.01 (m, 1H), Catalyzed Hydroboration of Alkenes with Pinacolborane. Tetrahedron
6.61 (d, J = 16.0 Hz, 1H), 6.17 (dt, J = 16.0, 5.5 Hz, 1H), 4.54 (dd, J 2004, 60, 10695−10700. (c) Crudden, C. M.; Hleba, Y. B.; Chen, A.
= 5.5, 1.5 Hz, 2H), 1.06 (s, 12H). 13C{1H} NMR (125 MHz, C6D6) δ C. Regio- and Enantiocontrol in the Room-Temperature Hydro-
137.0, 130.6, 128.4, 127.3, 127.2, 126.5, 82.3, 65.2, 24.4. boration of Vinyl Arenes with Pinacol Borane. J. Am. Chem. Soc. 2004,


126, 9200−9201. (d) Obligacion, J. V.; Chirik, P. J. Highly Selective
Bis(imino)pyridine Iron-Catalyzed Alkene Hydroboration. Org. Lett.
ASSOCIATED CONTENT 2013, 15, 2680−2683. (e) Zhang, L.; Zuo, Z.; Leng, X.; Huang, Z. A
* Supporting Information
S Cobalt-Catalyzed Alkene Hydroboration with Pinacolborane. Angew.
The Supporting Information is available free of charge on the Chem., Int. Ed. 2014, 53, 2696−2700. (f) Kisan, S.; Krishnakumar, V.;
Gunanathan, C. Ruthenium-Catalyzed Anti-Markovnikov Selective
ACS Publications website at DOI: 10.1021/acs.joc.8b03045. Hydroboration of Olefins. ACS Catal. 2017, 7, 5950−5954.
HPLC and NMR spectra of products (PDF) (5) (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) Noh, D.; Yoon, S.
AUTHOR INFORMATION K.; Won, J.; Lee, J. Y.; Yun, J. An Efficient Copper(I)-Catalyst System
for the Asymmetric Hydroboration of β-Substituted Vinylarenes with
Corresponding Author Pinacolborane. Chem. - Asian J. 2011, 6, 1967−1969. (c) Feng, X.;
*E-mail: jaesook@skku.edu. Jeon, H.; Yun, J. Regio- and Enantioselective Copper(I)-Catalyzed
Hydroboration of Borylalkenes: Asymmetric Synthesis of 1,1-
ORCID Diborylalkanes. Angew. Chem., Int. Ed. 2013, 52, 3989−3992.
Won Jun Jang: 0000-0002-0673-7607 (6) Shoba, V. M.; Thacker, N. C.; Bochat, A. J.; Takacs, J. M.
Seung Min Song: 0000-0002-7921-3353 Synthesis of Chiral Tertiary Boronic Esters by Oxime-Directed
Jaesook Yun: 0000-0003-4380-7878 Catalytic Asymmetric Hydroboration. Angew. Chem., Int. Ed. 2016, 55,
1465−1469.
Author Contributions (7) When this reaction was carried out at room temperature or 40

W.J.J. and S.M.S. contributed equally. °C, only allylic alcohols were produced through 1,2-addition without
the formation of the desired product 2a.
Notes (8) α,β-Unsaturated aldehydes with alkyl substituents at the β-
The authors declare no competing financial interest. position such as (E)-but-2-en-1-al and (E)-pent-2-en-1-al under the


hydroboration conditions produced fully reduced alcohols such as
ACKNOWLEDGMENTS butanol and pentanol.
(9) Ohmura, T.; Awano, T.; Suginome, M. Stereospecific Suzuki-
This research was supported by a National Research Miyaura Coupling of Chiral α-(Acylamino)benzylboronic Esters with
Foundation of Korea grant (NRF-2016R1A2B4011719), Inversion of Configuration. J. Am. Chem. Soc. 2010, 132, 13191−
funded by the Korean government (MEST). 13193.

4433 DOI: 10.1021/acs.joc.8b03045


J. Org. Chem. 2019, 84, 4429−4434
The Journal of Organic Chemistry Note

(10) The absolute configuration of 3b was determined by comparing


optical rotation with the literature data. Shintani, R.; Takatsu, K.;
Takeda, M.; Hayashi, T. Copper-Catalyzed Asymmetric Allylic
Substitution of Allyl Phosphates with Aryl- and Alkenylboronates.
Angew. Chem., Int. Ed. 2011, 50, 8656−8659.
(11) Park, J.; Yun, J.; Kim, J.; Jang, D.-J.; Park, C. H.; Lee, K.
Brønsted Acid−Catalyzed Meyer−Schuster Rearrangement for the
Synthesis of α,β-Unsaturated Carbonyl Compounds. Synth. Commun.
2014, 44, 1924−1929.
(12) (a) Hayashi, Y.; Itoh, T.; Aratake, S.; Ishikawa, H. A
Diarylprolinol in an Asymmetric, Catalytic, and Direct Crossed-
Aldol Reaction of Acetaldehyde. Angew. Chem., Int. Ed. 2008, 47,
2082−2084. (b) Gieuw, M. H.; Ke, Z.; Yeung, Y.-Y. Lewis Base-
Promoted Ring-Opening 1,3-Dioxygenation of Unactivated Cyclo-
propanes Using a Hypervalent Iodine Reagent. Angew. Chem., Int. Ed.
2018, 57, 3782−3786. (c) Motloch, P.; Valterová, I.; Kotora, M.
Enantioselective Allylation of Thiophene-2-carbaldehyde: Formal
Total Synthesis of Duloxetine. Adv. Synth. Catal. 2014, 356, 199−204.
(13) 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.

4434 DOI: 10.1021/acs.joc.8b03045


J. Org. Chem. 2019, 84, 4429−4434

You might also like