Subscriber access provided by Bethel University
Article
Catalytic Asymmetric Conjugate Addition of Indoles to para-Quinone Methide Derivatives Jin-Rong Wang, Xiao-Li Jiang, Qing-Qing Hang, Shu Zhang, Guang-Jian Mei, and Feng Shi J. Org. Chem., Just Accepted Manuscript • Publication Date (Web): 22 May 2019 Downloaded from http://pubs.acs.org on May 24, 2019
Just Accepted “Just Accepted” manuscripts have been peer-reviewed and accepted for publication. They are posted online prior to technical editing, formatting for publication and author proofing. The American Chemical Society provides “Just Accepted” as a service to the research community to expedite the dissemination of scientific material as soon as possible after acceptance. “Just Accepted” manuscripts appear in full in PDF format accompanied by an HTML abstract. “Just Accepted” manuscripts have been fully peer reviewed, but should not be considered the official version of record. They are citable by the Digital Object Identifier (DOI®). “Just Accepted” is an optional service offered to authors. Therefore, the “Just Accepted” Web site may not include all articles that will be published in the journal. After a manuscript is technically edited and formatted, it will be removed from the “Just Accepted” Web site and published as an ASAP article. Note that technical editing may introduce minor changes to the manuscript text and/or graphics which could affect content, and all legal disclaimers and ethical guidelines that apply to the journal pertain. ACS cannot be held responsible for errors or consequences arising from the use of information contained in these “Just Accepted” manuscripts.
is published by the American Chemical Society. 1155 Sixteenth Street N.W., Washington, DC 20036 Published by American Chemical Society. Copyright © American Chemical Society. However, no copyright claim is made to original U.S. Government works, or works produced by employees of any Commonwealth realm Crown government in the course of their duties.
Page 1 of 40 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
The Journal of Organic Chemistry
Catalytic Asymmetric Conjugate Addition of Indoles to para-Quinone Methide Derivatives
Jin-Rong Wang,‡,a Xiao-Li Jiang,‡,a Qing-Qing Hang,a Shu Zhang,*,b Guang-Jian Mei*,a and Feng Shi*,a aSchool
of Chemistry and Materials Science, Jiangsu Normal University, Xuzhou, 221116, China
bDepartment
of Radiotherapy, The First Affiliated Hospital of Nanjing Medical University, Nanjing, 210029, China
E-mail:
[email protected];
[email protected];
[email protected] ‡These
authors contributed equally to the work. OH
O t-Bu
t-Bu
t-Bu
G
t-Bu
O
OH R N H
10 mol% CPA, acetone
+ R
MgSO4, -30 oC
1
R1
H
O
NH
OH R 20 examples 54%-97%, 90:10 to 96:4 er
P
O OH
G G = 9-anthracenyl CPA
Abstract: A catalytic asymmetric conjugate addition of indoles to o-hydroxyphenyl substituted p-quinone methides has been established in the presence of chiral phosphoric acid, which afforded chiral indole-containing triarylmethanes in generally high yields (54%-97%) and good enantioselectivities (90:10 to 96:4 er). The control experiments indicated that o-hydroxyphenyl substituted p-quinone methides had a high possibility to transform into o-quinone methides in the presence of chiral phosphoric acid, and the formation of o-quinone methides might be a necessity for the reaction. This reaction will not only contribute to the research field of catalytic asymmetric transformations of p-quinone methides and o-quinone methides, but also provide a useful method for the construction of enantioenriched triarylmethane frameworks.
ACS Paragon Plus Environment
The Journal of Organic Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
Introduction In recent years, quinone methides (QMs) such as para-quinone methides (p-QMs)1-4 and ortho-quinone methides (o-QMs)5-8 have been recognized as versatile reactants or intermediates for catalytic asymmetric reactions. As a result, rapid developments have been achieved in this research field, and chemists have devised a variety of p-QMs and o-QMs or their precursors. Among them, o-hydroxyphenyl substituted p-QMs have been widely utilized in catalytic asymmetric (4+n) cyclizations including (4+1), (4+2) and (4+3) cyclizations (Scheme 1, eq. 1).9-11 However, in sharp contrast, catalytic asymmetric conjugate addition of o-hydroxyphenyl substituted p-QMs has rarely been reported (eq. 2). Until recently, Li and co-workers established an organocatalytic asymmetric conjugate addition of 5H-thiazol-4-ones to o-hydroxyphenyl substituted p-QMs in a diastereodivergent mode (eq. 3).12a In addition, Lu, Weng and coworkers reported a catalytic asymmetric conjugate addition of arylboronic acids to this class of p-QMs in the presence of BINOL-catalyst (eq. 4).12b In spite of these elegant work, the catalytic asymmetric conjugate addition of o-hydroxyphenyl substituted p-QMs is rather limited, which requires rapid development in this research field.
ACS Paragon Plus Environment
Page 2 of 40
Page 3 of 40 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
The Journal of Organic Chemistry
OH R1
Nu E
Cat.* O R1
R1
R1
(4+n) cyclization
Well-established (1)
* Nu
R
O E OH R1
R OH
Cat.*
R1
Nu Rarely reported
conjugate addition R
(2)
* Nu OH
Two examples of catalytic asymmetric conjugate additions: O R1
OH
R1
R1 O
R2 + R
Organo-Cat.* S
R2 O
N
*
R
Ar
OH
OH
O R1
R OH
* S
(3)
N Ar
OH
R1
+
R1
R1
HO
OH B
R1
BINOL-Cat.* R
3
R
(4)
* OH
R3
Scheme 1. Profile of catalytic asymmetric reactions involving o-hydroxyphenyl substituted p-QMs
In order to fulfil this task and based on our long-lasting interests in synthesizing indole derivatives,13 we designed a chiral phosphoric acid14 (CPA)-catalyzed asymmetric conjugate addition of indoles15 to o-hydroxyphenyl substituted p-QMs. As illustrated in Scheme 2, in the presence of CPA, o-hydroxyphenyl substituted p-QMs could transform into o-QMs because of the low isomerization energy (6.7 kcal/mol).10d Then, CPA could form two hydrogen bonds with both o-QMs and indoles, thus promoting an enantioselective conjugate addition between them to generate chiral triarylmethane products. Notably, the analogues of this class of indole-containing triarylmethane frameworks exist in lots of biologically active compounds.16
ACS Paragon Plus Environment
The Journal of Organic Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
O R1
Page 4 of 40
OH
R1
R1
R1 R2
R2
+
R
CPA N H
OH
*
R
NH
OH
p-QM
CPA 6.7 kcal/mol
OH R1
OH R1
R1
R2
R2 N H CPA
R o-QM
O
R1
N H
R O
O
H O P *
Scheme 2. Design of catalytic asymmetric conjugate addition of indoles to o-hydroxyphenyl substituted p-QMs
Results and Discussion Based on this design, we tried the reaction of indole 1a with o-hydroxyphenyl substituted p-QM 2a in the presence of CPA 4a (Table 1, entry 1). As expected, the conjugate addition smoothly occurred to give triarylmethane product 3aa in a high yield of 97% albeit with an extremely low enantioselectivity of 54:46 er. Nevertheless, this preliminary result demonstrated the feasibility of our design. To improve the enantioselectivity, a series of CPAs 4 were screened (entries 1-7), which found that CPA 4e bearing two bulky 9-anthracenyl groups could facilitate the conjugate addition in the highest enantioselectivity of 76:24 er (entry 5). Then, in the presence of the optimal catalyst 4e, some representative solvents were screened (entries 8-12), which discovered that acetone could deliver the reaction in the best enantioselectivity of 89:11 er (entry 9). So, acetone was selected as the most suitable reaction media. Then, the effect of temperature on the reaction was investigated (entries 13-17). It was found that elevating the reaction temperature was detrimental to the enantioselectivity (entry 13 vs entry 9), while lowering the reaction temperature could improve the enantioselectivity to some extent (entries 14-17 vs entry
ACS Paragon Plus Environment
Page 5 of 40 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
The Journal of Organic Chemistry
9). Thus, -30 oC was chosen as the optimal reaction temperature (entry 16). Subsequently, the reagents ratio was carefully modulated (entries 18-22), and the molar ratio of 1.2:1 was set as the suitable reagents ratio (entry 20). Finally, some additives such as molecular sieves (MS) and anhydrous sulfate were tentatively added to the reaction (entries 23-27), which found that the addition of magnesium sulfate could enhance the enantioselectivity to the highest level of 95:5 er (entry 26). In addition, we also utilized magnesium phosphate of 4e as a catalyst in this reaction in the absence or presence of magnesium sulfate (entries 28-29). It was found that product 3aa could be generated in a high yield of 98% and a considerable enantioselectivity (85:15 er and 89:11 er, respectively). But in these cases, the enantioselectivities were lower than those of the reactions catalyzed by phosphoric acid 4e (entry 28 vs entry 20, entry 29 vs entry 26). These results suggested that magnesium phosphate of 4e could also catalyze the reaction, but it was inferior to phosphoric acid 4e in terms of controlling the enantioselectivity. Because the addition of magnesium sulfate to the reaction system might have a possibility to convert the free phosphoric acid 4e into the corresponding Mg-phosphate salt, a Lewis acid-catalyzed process could not be excluded in this reaction. Table 1. Screening of catalysts and optimization of reaction conditionsa G
G
4a, G = 4-ClC6H4 4b, G = 2-naphthyl 4c, G = 1-naphthyl 4d, G = 9-phenanthrenyl 4e, G = 9-anthracenyl 4f, G = 2,4,6-(i-Pr)3C6H2 4g,G = SiPh3
O
t-Bu
O O (S)-4
P
t-Bu
O OH
OH t-Bu
t-Bu
OH N H
10 mol% Cat.
+
*
solvent, T oC
NH
OH 2a
1a
entry
Cat.
solvent
3aa
T (oC)
1a:2a
additives
ACS Paragon Plus Environment
yield (%)b
erc
The Journal of Organic Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
Page 6 of 40
1
4a
ClCH2CH2Cl
25
1:1.2
-
97
54:46
2
4b
ClCH2CH2Cl
25
1:1.2
-
90
55:45
3 4 5 6 7 8
4c 4d 4e 4f 4g 4e
1:1.2 1:1.2 1:1.2 1:1.2 1:1.2 1:1.2 1:1.2
-
75:25 61:39 76:24 52:48 57:43 88:12
4e
25 25 25 25 25 25 25
98 97 95 97 98 98
9
ClCH2CH2Cl ClCH2CH2Cl ClCH2CH2Cl ClCH2CH2Cl ClCH2CH2Cl EtOAc acetone
97
89:11
10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29
4e 4e 4e 4e 4e 4e 4e 4e 4e 4e 4e 4e 4e 4e 4e 4e 4e 4e Mg[4e]2 Mg[4e]2
1,4-dioxane CH3CN toluene acetone acetone acetone acetone acetone acetone acetone acetone acetone acetone acetone acetone acetone acetone acetone acetone acetone
25 25 25 50 0 -15 -30 -40 -30 -30 -30 -30 -30 -30 -30 -30 -30 -30 -30 -30
1:1.2 1:1.2 1:1.2 1:1.2 1:1.2 1:1.2 1:1.2 1:1.2 3:1 2:1 1.2:1 1:2 1:3 1.2:1 1.2:1 1.2:1 1.2:1 1.2:1 1.2:1 1.2:1
3Å MS 4Å MS 5Å MS MgSO4 Na2SO4 MgSO4
98 97 98 94 90 88 94 94 96 95 97 97 98 94 93 95 97 95 98 98
85:15 84:16 68:32 83:17 90:10 92:8 92:8 91:9 92:8 92:8 92:8 87:13 88:12 68:32 65:35 75:25 95:5 92:8 85:15 89:11
aUnless
otherwise indicated, the reaction was carried out at the 0.05 mmol scale in a solvent (1 mL) with or
without additives (50 mg) for 12 h. bIsolated yield. cThe enantiomeric ratio (er) was determined by HPLC.
After establishing the optimal reaction conditions, we then investigated the substrate scope of indoles 1 in the conjugate addition with o-hydroxyphenyl substituted p-QM 2a. As summarized in Table 2, this catalytic asymmetric conjugate addition reaction could be applicable to a wide range of indoles 1 bearing either electron-donating or electron-withdrawing groups at different positions of the phenyl ring (C4-C7), which successfully participated in the reaction to give products 3 in good yields (71%-97%) and high enantioselectivities (90:10 to 96:4 er). It seems that the
ACS Paragon Plus Environment
Page 7 of 40 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
The Journal of Organic Chemistry
electronic nature of the substituents exerted some effect on the enantioselectivity because the electron-withdrawing
groups
generally
showed
higher
capacity
in
controlling
the
enantioselectivity than electron-donating groups at the same position of the indole ring (entries 4-8 vs entry 3; entries 11-14 vs entries 9-10). However, in most cases, the position of the substituents did not show evident influence on the enantioselectivity. For example, C5- and C6-methyl substituted indoles 1c and 1i afforded products 3ca and 3ia in the same enantioselectivity of 90:10 er (entry 3 vs entry 9). The similar phenomena could be found in the cases of C5- and C6- fluoro or chloro-substituted indoles (entry 4 vs 11, entry 5 vs 12). Table 2. Substrate scope of indoles 1a OH
O t-Bu 5
t-Bu
4
t-Bu
OH
R 6
t-Bu
N H
7
10 mol% 4e, acetone
+
2a
1
NH
H
MgSO4, -30 oC OH 3
yield
R
entry
R (1)
3
1
H (1a)
3aa
97
95:5
2
4-OMe (1b)
3ba
87
90:10
3
5-Me (1c)
3ca
91
90:10
4
5-F (1d)
3da
97
95:5
5
5-Cl (1e)
3ea
6
5-Br (1f)
3fa
7
5-I (1g)
3ga
8
5-CN (1h)
3ha
93 90 85 71
93:7 93:7 95:5 90:10
9
6-Me (1i)
3ia
10
6-OMe (1j)
3ja
11d
6-F (1k)
3ka
12
6-Cl (1l)
3la
13
6-Br (1m)
3ma
14
6-CN (1n)
3na
15
7-F (1o)
3oa
98 92 79 82 95 75 83
90:10 90:10 96:4 92:8 90:10 93:7 95:5
aUnless
(%)b
erc
indicated otherwise, the reaction was carried out in 0.1 mmol scale in acetone (2 mL) with MgSO4 (100 mg) at -30 oC for 12 h, and the molar ratio of 1:2a was 1.2:1. bIsolated yield. cThe enantiomeric ratio (er) was determined by HPLC. dWhen the reaction was performed with 50 mg of MgSO4, the yield was 75% and the enantioselectivity was 91:9 er.
ACS Paragon Plus Environment
The Journal of Organic Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
Page 8 of 40
Then, the substrate scope of o-hydroxyphenyl substituted p-QMs 2 was studied by the conjugate addition with indoles 1a and 1k. As listed in Table 3, several p-QMs 2a-2f bearing C5 or C4-substituents on the phenyl ring could be utilized as suitable substrates in the reaction, affording products 3 in moderate to high yields (54%-97%) with good enantioselectivities (90:10 to 95:5 er). The electronic nature of the substituents seems to have no obvious effect on the enantioselectivity due to the fact that substrates 2b-2e bearing either electron-withdrawing groups or electron-donating groups delivered the products in the same enantioselectivity of 90:10 er (entries 2-5). Table 3. Substrate scope of o-hydroxyphenyl substituted p-QMs 2a OH t-Bu
O t-Bu
t-Bu
t-Bu OH
N H
R
1a, R = H 1k, R = F
+
6
5
R
10 mol% 4e, acetone MgSO4, -30 oC
1
4
3
OH 2
NH
H R1 3
R
entry
1
R1 (2)
3
yield (%)b
erc
1
1a
H (2a)
3aa
97
95:5
2
1a
5-F (2b)
3ab
81
90:10
3
1k
5-Cl (2c)
3kc
4
1k
5-Me (2d)
3kd
97 97
90:10 90:10
5
1k
5-OMe (2e)
3ke
6
1k
4-Br (2f)
3kf
60 54
90:10 90:10
aUnless
indicated otherwise, the reaction was carried out in 0.1 mmol scale in acetone (2 mL) with MgSO4 (100 mg) at -30 oC for 12 h, and the molar ratio of 1:2 was 1.2:1. bIsolated yield. cThe enantiomeric ratio (er) was determined by HPLC.
The absolute configuration of product 3ba was determined to be (R)-configuration via its single-crystal X-ray analysis (97:3 er after recrystallization).17 Thus, the absolute configuration of other products 3 was deduced to be (R) by analogy. In order to investigate the possible activation mode of chiral catalyst 4e on substrates 1 and 2, we performed some control experiments (Scheme
ACS Paragon Plus Environment
Page 9 of 40 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
The Journal of Organic Chemistry
3). First, N-methyl-protect indole 1p was used as a substrate instead of N-unprotected indoles under the standard conditions (eq. 5). In this case, indole 1p could undergo conjugate addition with p-QM 2a to give product 3pa in a good yield of 79%. However, nearly no enantio-control was observed during the formation of product 3pa (52:48 er). This result demonstrated that the N-H group of indoles played an important role in controlling the enantioselectivity possibly via forming a hydrogen bond with chiral catalyst 4e. Second, o-methoxyphenyl substituted p-QM 2g and phenyl substituted p-QM 2h were employed as substrates instead of o-hydroxyphenyl substituted p-QMs 2 in the conjugate addition under the standard conditions (eq. 6-7). In the two cases, no reaction occurred, which indicate that the OH group of phenyl-substituted p-QMs played a crucial role in controlling the reactivity via the hydrogen-bonding interaction with catalyst 4e. Moreover, ortho-hydroxybenzyl alcohol 2i as a precursor of o-QM was utilized as a substrate in the reaction. It was found that no reaction occurred under the standard conditions (eq. 8). However, elevating the reaction temperature and changing the solvent led to the generation of desired product 3ai in a high yield of 98% and a considerable enantioselectivity of 87:13 er (eq. 9). These results implied that the reactivity of ortho-hydroxybenzyl alcohols is lower than o-hydroxyphenyl substituted p-QMs 2 in the conjugate addition at low temperature, and the conjugate addition of p-QMs 2 might occur via the formation of o-QM intermediates. Finally, tetrasubstituted p-QM 2j was tentatively employed to the reaction under the standard conditions or at a higher temperature. However, no reaction occurred either at lower temperature or at higher temperature (eq. 10). This result indicated that the reactivity of tetrasubstituted p-QMs is much lower than commonly used trisubstituted p-QMs 2, which still remains to be a challenge in the chemistry of p-QMs.
ACS Paragon Plus Environment
The Journal of Organic Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
Page 10 of 40
OH t-Bu
O t-Bu
t-Bu
t-Bu OH
N Me
10 mol% 4e, acetone
+
*
o
MgSO4, -30 C
N Me (5)
OH
1p
3pa 79%, 52:48 er
2a O t-Bu
N H
t-Bu 10 mol% 4e, acetone
+
(6)
No reaction
MgSO4, -30 oC OMe 2g
1a
O t-Bu
N H
t-Bu 10 mol% 4e, acetone
+
No Reaction
MgSO4, -30 oC
1a
2h
H
10 mol% 4e, acetone Ph MeO N H
No Reaction
OH
Ph 10 mol% 4e, DCE
2i
MeO
o
MgSO4, 50 C O t-Bu N H
(8)
MgSO4, -30 oC
OH
+
1a
(7)
*
(9)
OH NH 3ai 98%, 87:13 er
t-Bu 10 mol% 4e, acetone
+
MgSO4, -30 oC or 50 oC
No Reaction
(10)
1a OH 2j
OMe
Scheme 3. Control experiments
Based on the experimental results, we suggested two possible reaction pathways and activation modes of the catalyst to the substrates (Scheme 4). In pathway A, it is suggested that o-hydroxyphenyl substituted p-QMs 2 transformed into o-QMs 2’ in the presence of CPA 4e. Then, CPA 4e simultaneously formed two hydrogen bonds with the N-H group of indoles 1 and the carbonyl group of o-QMs 2’, thus facilitating an enantioselective 1,4-addition of indoles 1 to o-QMs 2’ and giving rise to products 3 with observed (R)-configuration. In pathway B, it is suggested that o-hydroxyphenyl substituted p-QMs 2 were directly attacked by indoles 1 in the
ACS Paragon Plus Environment
Page 11 of 40
presence CPA 4e to undergo an enantioselective 1,6-addition, thus generating (R)-3. However, in the control experiment (Scheme 3, eq. 6), o-methoxyphenyl substituted p-QM 2g failed to participate in the conjugate reaction. This result indicated that the formation of o-QMs 2’ might be a necessity for the reaction. As illustrated in Scheme 4, in the presence of CPA, o-hydroxyphenyl substituted p-QMs 2 could easily transform into o-QMs 2’ via the proton-transfer process, and CPA could serve as a shuttle for proton transfer. On the contrary, p-QM 2g could hardly transform into o-QMs 2g’ due to the absence of the OH group, which is crucial for proton transfer. So, based on the control experiment, we consider that pathway A has a higher possibility than pathway B, but pathway B should not be totally excluded. OH t-Bu
t-Bu OH
OH
t-Bu
H
1
R
O Pathway A 2' (o-QM)
9-An
O
P
ac hr
O
O
thrac
t-Bu
OH
nt
O
1
9-A
CPA 4e
R1
N H
H
R
t-Bu
en yl
t-Bu
O
H
CPA 4e
enyl
NH
R R1 (R)-3
1,4-addition CPA 4e
O t-Bu
9-Anthracenyl
t-Bu
O
H
H N 1
CPA 4e
R
nyl
H
R
(R)-3
race
OH 2 (p-QM)
O
t-Bu
t-Bu
1 Pathway B
O
nth
R
P
O
O
CPA 4e
9-A
1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
The Journal of Organic Chemistry
1
OH 1,6-addition O
H
t-Bu
t-Bu
O *
O
OH
t-Bu
t-Bu
t-Bu
CPA 4e
P OR
O
H 2 (p-QM)
O
t-Bu
t-Bu
t-Bu
CPA 4e R O 2' (o-QM)
OMe 2g
Scheme 4. Possible reaction pathway
ACS Paragon Plus Environment
R 2g'
OMe
The Journal of Organic Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
Page 12 of 40
Moreover, to examine the utility of this reaction, we performed two larger-scale reactions under the standard conditions (Scheme 5). To our satisfaction, the catalytic asymmetric conjugate addition reaction of indole 1k to 2a at the 0.5 mmol scale successfully afforded product 3ka in a nearly maintained good enantioselectivity of 94:6 er and an excellent yield of 93% (eq. 11), which is higher than the small-scale reaction (Table 2, entry 11). In addition, the catalytic asymmetric reaction between indole 1a and p-QM 2a at the 1 mmol scale also smoothly offered product 3aa in a quantitative yield and a high enantioselectivity of 93:7 er (eq. 12), which are similar to the results of small-scale reaction (Table 2, entry 1). These results demonstrated that the catalytic asymmetric conjugate addition reaction could be used as a practical tool for the synthesis of chiral indole-containing triarylmethanes. OH t-Bu
O t-Bu
t-Bu
OH
10 mol% 4e, acetone
+
N H
F
t-Bu
NH
H
(11)
MgSO4, -30 oC OH 3ka F 93% (207mg), 94:6 er
2a 0.5 mmol
1k 0.6 mmol
OH
O t-Bu
t-Bu
1.2 mmol
*
MgSO4, -30 oC
N H OH 2a 1.0 mmol
t-Bu
OH
10 mol% 4e, acetone
+
1a
t-Bu
NH
(12)
3aa 99% (423 mg), 93:7 er
Scheme 5. Larger-scale reactions. Finally, to do some preliminary derivation of products, we carried out a Suzuki coupling of product 3fa with 4-chlorophenylboronic acid, which generated product 5 in a good yield of 79% and a maintained high enantioselectivity of 94:6 er (Scheme 6).
ACS Paragon Plus Environment
Page 13 of 40 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
The Journal of Organic Chemistry
OH t-Bu
OH t-Bu
t-Bu
OH H
NH
t-Bu
OH 4-ClC6H4B(OH)2 (1.5 equiv.) Cs2CO3 (2 equiv.) Pd(OAc)2 (0.05 equiv.)
H
NH
Butyl di-1-adamantylphosphine (0.06 equiv.) DME (0.6 mL), 80 oC, 12 h
Br 3fa 93:7 er
Cl
5 79%, 94:6 er
Scheme 6. Preliminary derivation of product 3fa
Conclusions
In summary, we have established a catalytic asymmetric conjugate addition of indoles to o-hydroxyphenyl substituted p-QMs in the presence of chiral phosphoric acid, which afforded chiral indole-containing triarylmethanes in generally high yields (54%-97%) and good enantioselectivities (90:10 to 96:4 er). The control experiments indicated that o-hydroxyphenyl substituted p-QMs had a high possibility to transform into o-QMs in the presence of chiral phosphoric acid, and the formation of o-QMs might be a necessity for the reaction. The larger-scale reaction demonstrated that the catalytic asymmetric conjugate addition could be scaled up. This reaction will not only contribute to the research field of catalytic asymmetric transformations of p-QMs and o-QMs, but also provide a useful method for the construction of enantioenriched triarylmethane frameworks.
Experimental Section 1H
and
13C
NMR spectra were measured at 400 and 100 MHz, respectively. The solvents
used for NMR spectroscopy were DMSO-d6, CDCl3 and acetone-d6, using tetramethylsilane as the internal reference. HR MS (ESI) was determined by a HR MS/MS instrument. The X-ray source
ACS Paragon Plus Environment
The Journal of Organic Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
Page 14 of 40
used for the single crystal X-ray diffraction analysis of compound 3ba was GaKα (λ = 1.34139), and the thermal ellipsoid was drawn at the 30% probability level. Analytical grade solvents for the column chromatography were used after distillation, and commercially available reagents were used as received. Methods for the synthesis of substrates 2a-2g: p-QM derivatives 2 could be conveniently synthesized according to the known literature procedures.2a, 10a,18-20 Among them, 2a, 2c-2e10a, 2g18, 2h2a, 2i19 and 2j20 are known compounds. 2,6-di-tert-butyl-4-(2-hydroxybenzylidene)cyclohexa-2,5-dienone10a (2a): 78% yield (1.21 g); yellow solid; m.p. 158.6-159.8 oC; 1H NMR (400 MHz, DMSO-d6) δ 10.21 (s, 1H), 7.60 (s, 1H), 7.47 (d, J = 2.1 Hz, 1H), 7.36 (d, J = 7.6 Hz, 1H), 7.32 – 7.25 (m, 1H), 7.24 (d, J = 2.2 Hz, 1H), 6.95 (d, J = 8.0 Hz, 1H), 6.92 (m, 1H), 1.28 (s, 9H), 1.24 (s, 9H). 2,6-di-tert-butyl-4-(5-fluoro-2-hydroxybenzylidene)cyclohexa-2,5-dienone (2b): 70% yield (1.15 g); yellow solid; m.p. 160.1-161.4 oC; 1H NMR (400 MHz, DMSO-d6) δ 10.17 (s, 1H), 7.50 (s, 1H), 7.42 (d, J = 1.5 Hz, 1H), 7.20 (d, J = 1.7 Hz, 1H), 7.17 – 7.08 (m, 2H), 6.97 – 6.91 (m, 1H), 1.26 (s, 9H), 1.23 (s, 9H); 13C{1H} NMR (100 MHz, DMSO-d6) δ 186.1, 155.5(J = 234 Hz), 153.6, 148.3, 146.8, 139.8, 135.8, 131.0, 128.3, 123.5 (J = 8 Hz), 118.0 (J = 23 Hz), 117.3, 117.2, 117.0, 79.6, 35.4, 35.1, 29.7, 29.5; IR (KBr): 3335, 2957, 1591, 1558, 1541, 1362, 1268, 1177, 737, 668 cm-1; HRMS (ESI-TOF) m/z: [M + H]+ Calcd for C21H26FO2 329.1912, found 329.1907. 2,6-di-tert-butyl-4-(5-chloro-2-hydroxybenzylidene)cyclohexa-2,5-dienone10a
(2c):
75%
yield (1.29 g); yellow solid; m.p. 98.2-99.7 oC; 1H NMR (400 MHz, DMSO-d6) δ 10.52 (s, 1H), 7.50 (s, 1H), 7.40 (d, J = 2.0 Hz, 1H), 7.36 – 7.30 (m, 2H), 7.24 (d, J = 2.1 Hz, 1H), 6.96 (d, J = 8.6 Hz, 1H), 1.27 (s, 9H), 1.24 (s, 9H).
ACS Paragon Plus Environment
Page 15 of 40 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
The Journal of Organic Chemistry
2,6-di-tert-butyl-4-(2-hydroxy-5-methylbenzylidene)cyclohexa-2,5-dienone10a
(2d):
71%
yield (1.15 g); yellow solid; m.p. 116.8-118.3 oC; 1H NMR (400 MHz, DMSO-d6) δ 9.98 (s, 1H), 7.57 (s, 1H), 7.50 (d, J = 2.0 Hz, 1H), 7.22 (d, J = 2.1 Hz, 1H), 7.16 (s, 1H), 7.11 – 7.07 (m, 1H), 6.85 (d, J = 8.2 Hz, 1H), 2.24 (s, 3H), 1.27 (s, 9H), 1.24 (s, 9H). 2,6-di-tert-butyl-4-(2-hydroxy-5-methoxybenzylidene)cyclohexa-2,5-dienone10a (2e): 78% yield (1.33 g); orange solid; m.p. 81.7-83.2 oC; 1H NMR (400 MHz, DMSO-d6) δ 9.79 (s, 1H), 7.58 (s, 1H), 7.54 (d, J = 2.1 Hz, 1H), 7.23 (d, J = 2.1 Hz, 1H), 6.93 – 6.88 (m, 3H), 3.72 (s, 3H), 1.28 (s, 9H), 1.25 (s, 9H). 4-(4-bromo-2-hydroxybenzylidene)-2,6-di-tert-butylcyclohexa-2,5-dienone (2f): 68% yield (1.32 g); yellow solid; m.p. 189.8-190.1 oC; 1H NMR (400 MHz, DMSO-d6) δ 10.67 (s, 1H), 7.49 (s, 1H), 7.39 (d, J = 1.9 Hz, 1H), 7.30 (d, J = 8.2 Hz, 1H), 7.21 (d, J = 2.1 Hz, 1H), 7.14 – 7.07 (m, 2H), 1.26 (s, 9H), 1.23 (s, 9H); 13C{1H} NMR (100 MHz, DMSO-d6) δ 186.2, 158.2, 148.2, 146.6, 140.1, 135.9, 133.2, 130.72, 128.4, 124.2, 122.6, 122.4, 119.0, 79.6, 35.4, 35.1, 29.7, 29.6; IR (KBr): 3566, 2956, 1683, 1558, 1456, 1256, 883, 749, 668, 517 cm-1; HRMS (ESI-TOF) m/z: [M H]- Calcd for C21H24BrO2 387.0965, found 387.0979. 2,6-di-tert-butyl-4-(2-methoxybenzylidene)cyclohexa-2,5-dienone18 (2g): 81% yield (1.31 g); yellow solid; m.p. 138.7-139.3 oC; 1H NMR (400 MHz, CDCl3) δ 7.47 (d, J = 2.2 Hz, 1H), 7.41 (s, 1H), 7.39 (s, 1H), 7.37 (s, 1H), 7.08 (d, J = 2.3 Hz, 1H), 7.06 – 7.01 (m, 1H), 6.96 (d, J = 7.9 Hz, 1H), 3.89 (s, 3H), 1.34 (s, 9H), 1.29 (s, 9H). 4-benzylidene-2,6-di-tert-butylcyclohexa-2,5-dienone2a (2h): 71% yield (1.05 g); yellow solid; m.p. 75.7-76.2 oC; 1H NMR (400 MHz, CDCl3) δ 7.54 (s, 1H), 7.49 – 7.35 (m, 5H), 7.20 (s, 1H), 7.03 (s, 1H), 1.35 (s, 9H), 1.31 (s, 9H).
ACS Paragon Plus Environment
The Journal of Organic Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
Page 16 of 40
2-(hydroxy(phenyl)methyl)-4-methoxyphenol19 (2i): 85% yield (0.98 g); white solid; m.p. 105.6-106.1 oC; 1H NMR (400 MHz, CDCl3) δ 7.53 (s, 1H), 7.41 – 7.28 (m, 5H), 6.79 (d, J = 8.8 Hz, 1H), 6.72 (dd, J = 8.8, 3.0 Hz, 1H), 6.44 (d, J = 3.0 Hz, 1H), 5.89 (s, 1H), 3.67 (s, 3H), 3.37 (s, 1H).
2,6-di-tert-butyl-4-((3-hydroxynaphthalen-2-yl)(4-methoxyphenyl)methylene)cyclohexa-2, 5-dienone20 (2j): 57% yield (1.32g); yellow solid; m.p. 92.6-93.5 oC; 1H NMR (400 MHz, CDCl3) δ 7.86 (d, J = 8.9 Hz, 1H), 7.83 – 7.76 (m, 1H), 7.57 (d, J = 2.6 Hz, 1H), 7.43 – 7.37 (m, 1H), 7.35 – 7.28 (m, 4H), 7.25 – 7.23 (m, 1H), 6.89 (d, J = 8.9 Hz, 2H), 6.75 (d, J = 2.6 Hz, 1H), 5.23 (s, 1H), 3.83 (s, 3H), 1.33 (s, 9H), 1.05 (s, 9H).
General procedure for the synthesis of products 3 To the mixture of indoles 1 (0.12 mmol), p-QM derivatives 2 (0.1 mmol), chiral phosphoric acid 4e (0.01 mmol) and MgSO4 (100 mg) was added acetone (2 mL), which was stirred at -30 oC for 12 h. After the completion of the reaction which was indicated by TLC, the reaction mixture was directly purified by preparative thin layer chromatography to afford pure products 3.
(R)-2,6-di-tert-butyl-4-((2-hydroxyphenyl)(1H-indol-3-yl)methyl)phenol
(3aa):
Preparative thin layer chromatography (petroleum ether/EtOAc = 5/1); 97% yield (41.4 mg); brown solid; m.p. 93.8-95.6 oC; [α]D20 = +41.9 (c 0.83, acetone); 1H NMR (400 MHz, CDCl3) δ 8.01 (s, 1H), 7.38 – 7.29 (m, 2H), 7.23 – 7.13 (m, 2H), 7.11 (s, 2H), 7.06 – 7.00 (m, 2H), 6.90 – 6.82 (m, 2H), 6.72 – 6.66 (m, 1H), 5.70 (s, 1H), 5.41 – 5.15 (m, 1H), 5.15 (s, 1H), 1.38 (s, 18H); 13C{1H}
NMR (100 MHz, CDCl3) δ 154.3, 152.6, 136.9, 135.9, 132.3, 130.2, 130.1, 127.9, 126.9,
125.5, 123.9, 122.5, 120.7, 120.0, 119.7, 118.5, 116.5, 111.2, 43.9, 34.4, 30.4; IR (KBr): 3734, 3709, 3628, 2960, 1456, 742, 668, 649 cm-1; HRMS (ESI-TOF) m/z: [M - H]- Calcd for C29H32NO2 426.2438, found 426.2419; Enantiomeric ratio = 95:5, determined by HPLC (Daicel
ACS Paragon Plus Environment
Page 17 of 40 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
The Journal of Organic Chemistry
Chiralpak AD-H, hexane/isopropanol = 90:10, flow rate 1.0 mL/min, T = 30 °C, 254 nm); tR = 11.217 min (minor), tR = 13.210 min (major).
(R)-2,6-di-tert-butyl-4-((2-hydroxyphenyl)(4-methoxy-1H-indol-3-yl)methyl)phenol (3ba): Preparative thin layer chromatography (petroleum ether/EtOAc = 5/1); 87% yield (39.8 mg); brown solid; m.p. 97.1–98.8 oC; [α]D20 = +47.3 (c 0.80, acetone); 1H NMR (400 MHz, CDCl3) δ 7.98 (s, 1H), 7.16 – 7.03 (m, 4H), 7.00 – 6.91 (m, 2H), 6.89 – 6.75 (m, 2H), 6.62 – 6.54 (m, 1H), 6.44 (d, J = 7.7 Hz, 1H), 6.07 (s, 1H), 5.39 (s, 1H), 5.09 (s, 1H), 3.66 (s, 3H), 1.37 (s, 18H); 13C{1H}
NMR (100 MHz, CDCl3) δ 154.9, 153.9, 152.2, 138.3, 135.6, 133.2, 131.7, 129.9, 127.3,
125.7, 123.2, 122.5, 120.3, 119.2, 117.2, 116.1, 104.5, 100.3, 55.2, 43.9, 34.4, 30.4; IR (KBr): 3735, 3648, 3628, 3420, 2956, 1506, 1435, 1259, 1231, 1087 cm-1; HRMS (ESI-TOF) m/z: [M H]- Calcd for C30H34NO3 456.2544, found 456.2524; Enantiomeric ratio = 90:10, determined by HPLC (Daicel Chiralpak OD-H, hexane/isopropanol = 70:30, flow rate 1.0 mL/min, T = 30 °C, 254 nm); tR = 3.803 min (major), tR = 4.147 min (minor). (R)-2,6-di-tert-butyl-4-((2-hydroxyphenyl)(5-methyl-1H-indol-3-yl)methyl)phenol (3ca): Preparative thin layer chromatography (petroleum ether/EtOAc = 5/1); 91% yield (40.0 mg); brown solid; m.p. 180.0-181.9 oC; [α]D20 = +43.0 (c 0.80, acetone); 1H NMR (400 MHz, CDCl3) δ 7.93 (s, 1H), 7.25 – 7.21 (m, 1H), 7.20 – 7.13 (m, 1H), 7.13 – 7.09 (m, 3H), 7.02 (d, J = 7.7 Hz, 2H), 6.92 – 6.80 (m, 2H), 6.66 (s, 1H), 5.66 (s, 1H), 5.22 (s, 1H), 5.14 (s, 1H), 2.36 (s, 3H), 1.38 (s, 18H); 13C{1H} NMR (100 MHz, CDCl3) δ 154.3, 152.6, 135.9, 135.2, 132.3, 130.2, 130.1, 128.9, 127.9, 127.2, 125.5, 124.1, 124.0, 120.7, 119.5, 117.9, 116.5, 110.9, 43.9, 34.4, 30.4, 21.6; IR (KBr): 3648, 3629, 3420, 2957, 1456, 1435, 1232, 1154, 1090, 753 cm-1; HRMS (ESI-TOF) m/z: [M - H]- Calcd for C30H34NO2 440.2595, found 440.2586; Enantiomeric ratio = 90:10,
ACS Paragon Plus Environment
The Journal of Organic Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
determined by HPLC (Daicel Chiralpak AD-H, hexane/isopropanol = 90:10, flow rate 1.0 mL/min, T = 30 °C, 254 nm); tR = 10.723 min (minor), tR = 13.433 min (major). (R)-2,6-di-tert-butyl-4-((5-fluoro-1H-indol-3-yl)(2-hydroxyphenyl)methyl)phenol (3da): Preparative thin layer chromatography (petroleum ether/EtOAc = 5/1); 97% yield (43.2 mg); brown solid; m.p. 164.6-166.5 oC; [α]D20 = +28.9 (c 0.86, acetone); 1H NMR (400 MHz, CDCl3) δ 8.01 (s, 1H), 7.33 (dd, J = 15.4, 8.1 Hz, 2H), 7.23 – 7.14 (m, 2H), 7.11 (s, 2H), 7.06 – 6.96 (m, 2H), 6.91 – 6.80 (m, 2H), 6.73 – 6.64 (m, 1H), 5.70 (s, 1H), 5.15 (s, 1H), 1.38 (s, 18H); 13C{1H} NMR (100 MHz, CDCl3) δ 157.7 (J = 233 Hz), 154.1, 152.6, 136.0, 133.4, 132.0, 130.0, 129.8, 128.0, 127.3(J = 10 Hz), 125.6, 125.4, 120.8, 118.8 (J = 9 Hz), 118.7, 116.5, 111.9, 111.0, 110.7, 104.9 (J = 23 Hz), 43.6, 34.4, 30.4; 19F NMR (376 MHz, CDCl3) δ (ppm) 124.1; IR (KBr): 3854, 3735, 3629, 2959, 1559, 1540, 1507, 1457, 1165, 1089 cm-1; HRMS (ESI-TOF) m/z: [M - H]Calcd for C29H31FNO2 444.2344, found 444.2327; Enantiomeric ratio = 95:5, determined by HPLC (Daicel Chiralpak AD-H, hexane/isopropanol = 90:10, flow rate 1.0 mL/min, T = 30 °C, 254 nm); tR = 13.540 min (minor), tR = 17.560 min (major). (R)-2,6-di-tert-butyl-4-((5-chloro-1H-indol-3-yl)(2-hydroxyphenyl)methyl)phenol (3ea): Preparative thin layer chromatography (petroleum ether/EtOAc = 5/1); 93% yield (42.8 mg); brown solid; m.p. 114.7-116.5 oC; [α]D20 = +23.8 (c 0.86, acetone); 1H NMR (400 MHz, CDCl3) δ 8.03 (s, 1H), 7.27 (s, 1H), 7.25 – 7.24 (m, 1H), 7.19 – 7.10 (m, 2H), 7.06 (s, 2H), 6.97 (d, J = 7.5 Hz, 1H), 6.89 – 6.80 (m, 2H), 6.73 – 6.65 (m, 1H), 5.65 (s, 1H), 5.14 (s, 1H), 5.02 (s, 1H), 1.36 (s, 18H); 13C{1H} NMR (100 MHz, CDCl3) δ 154.0, 152.7, 136.0, 135.2, 131.9, 129.9, 129.8, 128.0, 125.4, 125.3, 125.2, 122.7, 120.8, 119.4, 118.5, 116.5, 112.2, 43.4, 34.4, 30.4; IR (KBr): 3725, 3648, 3628, 3420, 2958, 1558, 1540, 1506, 1457, 668 cm-1; HRMS (ESI-TOF) m/z: [M - H]-
ACS Paragon Plus Environment
Page 18 of 40
Page 19 of 40 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
The Journal of Organic Chemistry
Calcd for C29H31ClNO2 460.2049, found 460.2040; [M - H]- Calcd for C29H31ClNO2 462.2019, found 462.2030; Enantiomeric ratio = 93:7, determined by HPLC (Daicel Chiralpak AD-H, hexane/isopropanol = 90:10, flow rate 1.0 mL/min, T = 30 °C, 254 nm); tR = 14.083 min (minor), tR = 18.713 min (major). (R)-4-((5-bromo-1H-indol-3-yl)(2-hydroxyphenyl)methyl)-2,6-di-tert-butylphenol (3fa): Preparative thin layer chromatography (petroleum ether/EtOAc = 5/1); 90% yield (45.5 mg); brown solid; m.p. 185.8-187.0 oC; [α]D20 = +41.5 (c 0.91, acetone); 1H NMR (400 MHz, CDCl3) δ 8.04 (s, 1H), 7.43 – 7.40 (m, 1H), 7.25 (d, J = 1.8 Hz, 1H), 7.22 – 7.19 (m, 1H), 7.18 – 7.13 (m, 1H), 7.06 (s, 2H), 7.00 – 6.94 (m, 1H), 6.89 – 6.75 (m, 2H), 6.71 – 6.61 (m, 1H), 5.66 (s, 1H), 5.15 (s, 1H), 5.05 (s, 1H), 1.37 (s, 18H); 13C{1H} NMR (100 MHz, CDCl3) δ 153.9, 152.7, 136.0, 135.5, 132.0, 130.0, 129.9, 128.7, 128.0, 125.4, 125.3, 125.1, 122.5, 120.9, 118.4, 116.5, 112.9, 112.7, 43.3, 34.4, 30.4; IR (KBr): 3735, 3675, 3648, 3628, 2957, 1457, 1233, 753, 668, 649 cm-1; HRMS (ESI-TOF) m/z: [M - H]- Calcd for C29H31BrNO2 504.1543, found 504.1525; [M - H]Calcd for C29H31BrNO2 506.1523, found 506.1527; Enantiomeric ratio = 93:7, determined by HPLC (Daicel Chiralpak AD-H, hexane/isopropanol = 70:30, flow rate 1.0 mL/min, T = 30 °C, 254 nm); tR = 3.933 min (minor), tR = 4.483 min (major). (R)-2,6-di-tert-butyl-4-((2-hydroxyphenyl)(5-iodo-1H-indol-3-yl)methyl)phenol
(3ga):
Preparative thin layer chromatography (petroleum ether/EtOAc = 5/1); 85% yield (47.1 mg); brown solid; m.p. 203.4-204.7 oC; [α]D20 = +41.9 (c 0.94, acetone); 1H NMR (400 MHz, CDCl3) δ 8.03 (s, 1H), 7.61 (s, 1H), 7.49 – 7.35 (m, 1H), 7.20 – 7.09 (m, 2H), 7.05 (s, 2H), 7.01 – 6.91 (m, 1H), 6.90 – 6.79 (m, 2H), 6.70 – 6.60 (m, 1H), 5.65 (s, 1H), 5.15 (s, 1H), 5.00 (s, 1H), 1.37 (s, 18H); 13C{1H} NMR (100 MHz, CDCl3) δ 153.9, 152.6, 136.0, 135.8, 131.9, 130.7, 129.9, 129.8,
ACS Paragon Plus Environment
The Journal of Organic Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
129.5, 128.8, 128.0, 125.1, 124.6, 120.8, 118.0, 116.4, 113.2, 83.1, 43.2, 34.4, 30.3; IR (KBr): 3419, 2959, 1456, 1435, 1261, 1234, 1092, 753, 668, 457 cm-1; HRMS (ESI-TOF) m/z: [M - H]Calcd for C29H31INO2 552.1405, found 552.1393; Enantiomeric ratio = 95:5, determined by HPLC (Daicel Chiralpak AD-H, hexane/isopropanol = 70:30, flow rate 1.0 mL/min, T = 30 °C, 254 nm); tR = 3.943 min (minor), tR = 5.063 min (major). (R)-3-((3,5-di-tert-butyl-4-hydroxyphenyl)(2-hydroxyphenyl)methyl)-1H-indole-5-carbo nitrile (3ha): Preparative thin layer chromatography (petroleum ether/EtOAc = 5/1); 71% yield (32.3 mg); brown solid; m.p. 152.9-154.6 oC; [α]D20 = +17.0 (c 0.65, acetone); 1H NMR (400 MHz, CDCl3) δ 8.30 (s, 1H), 7.61 (s, 1H), 7.43 – 7.36 (m, 2H), 7.20 – 7.08 (m, 1H), 7.03 (s, 2H), 6.94 (d, J = 7.5 Hz, 1H), 6.88 – 6.81 (m, 2H), 6.81 – 6.77 (m, 1H), 5.73 (s, 1H), 5.15 (s, 1H), 4.93 (s, 1H), 1.35 (s, 18H); 13C{1H} NMR (100 MHz, CDCl3) δ 153.6, 152.7, 138.4, 136.0, 131.6, 129.8, 129.5, 128.1, 126.8, 125.9, 125.8, 125.3, 125.2, 120.9, 120.7, 120.2, 116.4, 112.0, 102.6, 42.7, 34.4, 30.3; IR (KBr): 3852, 3734, 3648, 3628, 2956, 1558, 1540, 1506, 668, 650 cm-1; HRMS (ESI-TOF) m/z: [M - H]- Calcd for C30H31N2O2 451.2391, found 451.2372; Enantiomeric ratio = 90:10, determined by HPLC (Daicel Chiralpak AD-H, hexane/isopropanol = 90:10, flow rate 1.0 mL/min, T = 30 °C, 254 nm); tR = 22.480 min (minor), tR = 28.243 min (major). (R)-2,6-di-tert-butyl-4-((2-hydroxyphenyl)(6-methyl-1H-indol-3-yl)methyl)phenol (3ia): Preparative thin layer chromatography (petroleum ether/EtOAc = 5/1); 98% yield (43.2 mg); brown solid; m.p. 81.8-82.4 oC; [α]D20 = +42.5 (c 0.86, acetone); 1H NMR (400 MHz, CDCl3) δ 7.88 (s, 1H), 7.20 – 7.13 (m, 3H), 7.10 (s, 2H), 7.06 – 6.99 (m, 1H), 6.91 – 6.76 (m, 3H), 6.64 – 6.58 (m, 1H), 5.65 (s, 1H), 5.19 (s, 1H), 5.13 (s, 1H), 2.44 (s, 3H), 1.37 (s, 18H); 13C{1H} NMR (100 MHz, CDCl3) δ 154.3, 152.5, 137.4, 135.8, 132.3, 132.2, 130.1, 130.0, 127.8, 125.4, 124.7,
ACS Paragon Plus Environment
Page 20 of 40
Page 21 of 40 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
The Journal of Organic Chemistry
123.1, 121.4, 120.6, 119.5, 118.3, 116.5, 111.1, 44.1, 34.4, 30.3, 21.7; IR (KBr): 3735, 3675, 3649, 3629, 2958, 1558, 1507, 668, 649, 472 cm-1; HRMS (ESI-TOF) m/z: [M - H]- Calcd for C30H34NO2 440.2595, found 440.2579; Enantiomeric ratio = 90:10, determined by HPLC (Daicel Chiralpak OD-H, hexane/isopropanol = 90:10, flow rate 1.0 mL/min, T = 30 °C, 254 nm); tR = 6.387 min (major), tR = 9.207 min (minor). (R)-2,6-di-tert-butyl-4-((2-hydroxyphenyl)(6-methoxy-1H-indol-3-yl)methyl)phenol (3ja): Preparative thin layer chromatography (petroleum ether/EtOAc = 5/1); 92% yield (42.2 mg); brown solid; m.p. 92.5-93.3 oC; [α]D20 = +35.3 (c 0.84, acetone); 1H NMR (400 MHz, CDCl3) δ 7.90 (s, 1H), 7.19 – 7.12 (m, 2H), 7.09 (s, 2H), 7.02 (d, J = 7.3 Hz, 1H), 6.91 – 6.77 (m, 3H), 6.68 (d, J = 8.5 Hz, 1H), 6.58 (s, 1H), 5.63 (s, 1H), 5.25 (s, 1H), 5.13 (s, 1H), 3.82 (s, 3H), 1.37 (s, 18H); 13C{1H} NMR (100 MHz, CDCl3) δ 156.7, 154.4, 152.5, 137.8, 135.9, 132.2, 130.1, 130.0, 127.9, 125.5, 122.6, 121.3, 120.7, 120.6, 118.5, 116.5, 109.7, 94.6, 55.6, 53.5, 44.2, 34.4, 30.4; IR (KBr): 3648, 3628, 3420, 2958, 1506, 1456, 1157, 802, 668, 649 cm-1; HRMS (ESI-TOF) m/z: [M - H]- Calcd for C30H34NO3 456.2544, found 456.2525; Enantiomeric ratio = 90:10, determined by HPLC (Daicel Chiralpak OD-H, hexane/isopropanol = 90:10, flow rate 1.0 mL/min, T = 30 °C, 254 nm); tR = 11.843 min (major), tR = 14.207 min (minor). (R)-2,6-di-tert-butyl-4-((6-fluoro-1H-indol-3-yl)(2-hydroxyphenyl)methyl)phenol (3ka): Preparative thin layer chromatography (petroleum ether/EtOAc = 5/1); 79% yield (35.1 mg); yellow solid; m.p. 209.0-210.6 oC; [α]D20 = +38.8 (c 0.70, acetone); 1H NMR (400 MHz, CDCl3) δ 7.99 (s, 1H), 7.22 – 7.11 (m, 2H), 7.06 (s, 2H), 7.05 – 7.02 (m, 1H), 7.00 – 6.96 (m, 1H), 6.88 – 6.81 (m, 2H), 6.81 – 6.73 (m, 1H), 6.68 – 6.61 (m, 1H), 5.64 (s, 1H), 5.13 (s, 1H), 5.06 (s, 1H), 1.35 (s, 18H); 13C{1H} NMR (100 MHz, CDCl3) δ 160.1 (J = 236 Hz), 154.1, 152.5, 136.8 (J = 12
ACS Paragon Plus Environment
The Journal of Organic Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
Page 22 of 40
Hz), 135.9, 131.9, 130.0, 129.8, 127.9, 125.3, 124.2, 124.1, 123.5, 120.8 (J = 10 Hz), 120.7, 118.7, 116.4, 108.5 (J = 25 Hz), 97.5 (J = 26 Hz), 43.7, 34.4, 30.3; 19F NMR (376 MHz, DMSO-d6) δ (ppm) 122.3; IR (KBr): 3628, 3586, 3566, 3420, 2958, 1652, 668, 649, 472, 457 cm-1; HRMS (ESI-TOF) m/z: [M - H]- Calcd for C29H31FNO2 444.2344, found 444.2334; Enantiomeric ratio = 96:4, determined by HPLC (Daicel Chiralpak AD-H, hexane/isopropanol = 90:10, flow rate 1.0 mL/min, T = 30 °C, 254 nm); tR = 13.690 min (minor), tR = 15.607 min (major). (R)-2,6-di-tert-butyl-4-((6-chloro-1H-indol-3-yl)(2-hydroxyphenyl)methyl)phenol
(3la):
Preparative thin layer chromatography (petroleum ether/EtOAc = 5/1); 82% yield (38.0 mg); brown solid; m.p. 89.5-91.0 oC; [α]D20 = +32.1 (c 0.76, acetone); 1H NMR (400 MHz, CDCl3) δ 7.99 (s, 1H), 7.34 (d, J = 1.7 Hz, 1H), 7.19 – 7.12 (m, 2H), 7.06 (s, 2H), 7.00 – 6.95 (m, 2H), 6.89 – 6.81 (m, 2H), 6.68 – 6.65 (m, 1H), 5.66 (s, 1H), 5.14 (s, 1H), 5.04 (s, 1H), 1.36 (s, 18H); 13C{1H} NMR (100 MHz, CDCl3) δ 154.0, 152.6, 137.2, 136.0, 132.0, 130.0, 129.8, 128.3, 128.0, 125.6, 125.4, 124.5, 120.9, 120.8, 120.4, 118.9, 116.5, 111.1, 43.5, 34.4, 30.4; [M - H]- Calcd for C29H31ClNO2 462.2019, found 462.2032; IR (KBr): 3648, 3628, 3420, 2959, 1456, 1435, 1234, 804, 754, 472 cm-1; HRMS (ESI-TOF) m/z: [M - H]- Calcd for C29H31ClNO2 460.2049, found 460.2030; Enantiomeric ratio = 92:8, determined by HPLC (Daicel Chiralpak OD-H, hexane/isopropanol = 90:10, flow rate 1.0 mL/min, T = 30 °C, 254 nm); tR = 8.120 min (major), tR = 9.760 min (minor). (R)-4-((6-bromo-1H-indol-3-yl)(2-hydroxyphenyl)methyl)-2,6-di-tert-butylphenol (3ma): Preparative thin layer chromatography (petroleum ether/EtOAc = 5/1); 95% yield (47.9 mg); brown solid; m.p. 96.1-97.7 oC; [α]D20 = +12.8 (c 0.96, acetone); 1H NMR (400 MHz, CDCl3) δ 7.99 (s, 1H), 7.50 (s, 1H), 7.19 – 7.08 (m, 3H), 7.06 (s, 2H), 6.99 – 6.95 (m, 1H), 6.89 – 6.80 (m,
ACS Paragon Plus Environment
Page 23 of 40 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
The Journal of Organic Chemistry
2H), 6.67 – 6.64 (m, 1H), 5.66 (s, 1H), 5.14 (s, 1H), 5.03 (s, 1H), 1.36 (s, 18H); 13C{1H} NMR (100 MHz, CDCl3) δ 154.0, 152.6, 137.6, 136.0, 132.0, 130.0, 129.8, 128.0, 125.9, 125.4, 124.5, 123.0, 121.3, 120.8, 119.0, 116.5, 116.0, 114.1, 43.5, 34.4, 30.4; IR (KBr): 3566, 3545, 3446, 2958, 1653, 1558, 1506, 668, 649, 457 cm-1; HRMS (ESI-TOF) m/z: [M - H]- Calcd for C29H31BrNO2 504.1543, found 504.1527; [M - H]- Calcd for C29H31BrNO2 506.1523, found 506.1525; Enantiomeric ratio = 90:10, determined by HPLC (Daicel Chiralpak OD-H, hexane/isopropanol = 90:10, flow rate 1.0 mL/min, T = 30 °C, 254 nm); tR = 8.440 min (major), tR = 9.900 min (minor). (R)-3-((3,5-di-tert-butyl-4-hydroxyphenyl)(2-hydroxyphenyl)methyl)-1H-indole-6-carbo nitrile (3na): Preparative thin layer chromatography (petroleum ether/EtOAc = 5/1); 75% yield (33.9 mg); brown solid; m.p. 220.5-220.8 oC; [α]D20 = +28.9 (c 0.68, acetone); 1H NMR (400 MHz, acetone-d6) δ 8.34 (s, 1H), 7.86 – 7.75 (m, 1H), 7.31 (d, J = 8.3 Hz, 1H), 7.20 – 7.16 (m, 1H), 7.15 (s, 2H), 7.07 – 6.97 (m, 3H), 6.92 – 6.88 (m, 1H), 6.76 – 6.70 (m, 1H), 6.06 (s, 1H), 5.90 (s, 1H), 2.93 (s, 1H), 1.36 (s, 18H); 13C{1H} NMR (100 MHz, acetone-d6) δ 154.5, 152.2, 137.0, 136.0, 134.1, 131.0, 130.2, 129.6, 128.5, 127.1, 125.4, 121.1, 120.7, 120.5, 120.4, 119.3, 116.2, 115.2, 103.5, 78.4, 40.7, 34.3, 29.9; IR (KBr): 3649, 3566, 3545, 3420, 2961, 1020, 798, 752, 668, 472 cm-1; HRMS (ESI-TOF) m/z: [M - H]- Calcd for C30H31N2O2 451.2391, found 451.2372; Enantiomeric ratio = 93:7, determined by HPLC (Daicel Chiralpak OD-H, hexane/isopropanol = 95:5, flow rate 1.0 mL/min, T = 30 °C, 254 nm); tR = 31.570 min (major), tR = 40.180 min (minor). (R)-2,6-di-tert-butyl-4-((7-fluoro-1H-indol-3-yl)(2-hydroxyphenyl)methyl)phenol (3oa): Preparative thin layer chromatography (petroleum ether/EtOAc = 5/1); 83% yield (37.1 mg);
ACS Paragon Plus Environment
The Journal of Organic Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
yellow solid; m.p. 78.8-80.0 oC; [α]D20 = +29.3 (c 0.74, acetone); 1H NMR (400 MHz, CDCl3) δ 8.18 (s, 1H), 7.19 – 7.12 (m, 1H), 7.06 (s, 2H), 7.05 – 7.02 (m, 1H), 7.00 – 6.95 (m, 1H), 6.94 – 6.87 (m, 2H), 6.87 – 6.81 (m, 2H), 6.73 – 6.68 (m, 1H), 5.68 (s, 1H), 5.14 (s, 1H), 5.01 (s, 1H), 1.36 (s, 18H); 13C{1H} NMR (100 MHz, CDCl3) δ 154.0, 152.6, 149.4 (J = 243 Hz), 136.0, 132.0, 130.7, 130.6, 130.0, 129.9, 127.9, 125.4, 125.2 (J = 13 Hz), 124.5, 120.7, 119.9 (J = 6 Hz), 119.6, 116.4, 115.8, 107.3, 107.1, 43.6, 34.4, 30.3; 19F NMR (376 MHz, CDCl3) δ (ppm) 135.4; IR (KBr): 3648, 3628, 3545, 3446, 2960, 1576, 1506, 668, 472, 457 cm-1; HRMS (ESI-TOF) m/z: [M - H]Calcd for C29H31FNO2 444.2344, found 444.2326; Enantiomeric ratio = 95:5, determined by HPLC (Daicel Chiralpak OJ-H, hexane/isopropanol = 95:5, flow rate 1.0 mL/min, T = 30 °C, 254 nm); tR = 43.540 min (major), tR = 67.777 min (minor). (R)-2,6-di-tert-butyl-4-((5-fluoro-2-hydroxyphenyl)(1H-indol-3-yl)methyl)phenol (3ab): Preparative thin layer chromatography (petroleum ether/EtOAc = 5/1); 81% yield (36.1 mg); brown solid; m.p. 102.9-104.1 oC; [α]D20 = +40.1 (c 0.72, acetone); 1H NMR (400 MHz, CDCl3) δ 8.05 (s, 1H), 7.37 (d, J = 8.2 Hz, 1H), 7.32 (d, J = 8.0 Hz, 1H), 7.24 – 7.17 (m, 1H), 7.08 (s, 2H), 7.07 – 7.02 (m, 1H), 6.88 – 6.81 (m, 1H), 6.81 – 6.75 (m, 1H), 6.74 – 6.67 (m, 2H), 5.66 (s, 1H), 5.17 (s, 1H), 4.99 (s, 1H), 1.38 (s, 18H); 13C{1H} NMR (100 MHz, CDCl3) δ 157.3 (J = 236 Hz), 152.8, 150.2, 150.1, 136.9, 136.1, 132.0 (J = 6 Hz), 131.6, 126.7, 125.4, 123.9, 122.6, 119.9, 119.8, 117.8, 117.3 (J = 8 Hz), 116.4 (J = 23 Hz), 114.1 (J = 23 Hz), 111.3, 43.8, 34.4, 30.4; 19F NMR (376 MHz, CDCl3) δ (ppm) 123.5; IR (KBr): 3853, 3838, 3735, 3649, 2960, 1558, 1507, 668, 472, 457 cm-1; HRMS (ESI-TOF) m/z: [M - H]- Calcd for C29H31FNO2 444.2344, found 444.2328; Enantiomeric ratio = 90:10, determined by HPLC (Daicel Chiralpak AD-H,
ACS Paragon Plus Environment
Page 24 of 40
Page 25 of 40 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
The Journal of Organic Chemistry
hexane/isopropanol = 90:10, flow rate 1.0 mL/min, T = 30 °C, 254 nm); tR = 10.990 min (minor), tR = 13.283 min (major). (R)-2,6-di-tert-butyl-4-((5-chloro-2-hydroxyphenyl)(6-fluoro-1H-indol-3-yl)methyl)phen ol (3kc): Preparative thin layer chromatography (petroleum ether/EtOAc = 5/1); 97% yield (46.4 mg); brown solid; m.p. 91.3-92.0 oC; [α]D20 = +34.5 (c 0.93, acetone); 1H NMR (400 MHz, CDCl3) δ 8.03 (s, 1H), 7.20 – 7.09 (m, 2H), 7.05 (s, 2H), 7.03 – 6.97 (m, 2H), 6.85 – 6.72 (m, 2H), 6.65 (s, 1H), 5.60 (s, 1H), 5.18 (s, 1H), 5.14 (s, 1H), 1.37 (s, 18H); 13C{1H} NMR (100 MHz, CDCl3) δ 161.4, 160.2 (J = 237 Hz), 152.9 (J = 8 Hz), 136.8 (J = 12 Hz), 136.1, 131.6, 131.2, 129.7, 127.8, 125.5, 125.2, 124.3, 124.2, 123.3, 120.6 (J = 10 Hz), 118.0, 117.9, 108.7 (J = 24 Hz), 97.6 (J = 26 Hz), 43.8, 34.4, 30.3; IR (KBr): 3735, 3649, 3420, 2960, 1457, 1157, 804, 668, 472, 457 cm-1; HRMS (ESI-TOF) m/z: [M - H]- Calcd for C29H30ClFNO2 478.1954, found 478.1936; [M - H]Calcd for C29H30ClFNO2 480.1925, found 480.1929; Enantiomeric ratio = 90:10, determined by HPLC (Daicel Chiralpak IB, hexane/isopropanol = 90:10, flow rate 1.0 mL/min, T = 30 °C, 254 nm); tR = 8.727 min (minor), tR = 9.263 min (major). (R)-2,6-di-tert-butyl-4-((6-fluoro-1H-indol-3-yl)(2-hydroxy-5-methylphenyl)methyl)phenol (3kd): Preparative thin layer chromatography (petroleum ether/EtOAc = 5/1); 97% yield (44.3 mg); brown solid; m.p. 95.7-96.1 oC; [α]D20 = +43.0 (c 0.89, acetone); 1H NMR (400 MHz, CDCl3) δ 7.98 (s, 1H), 7.18 (dd, J = 8.7, 5.4 Hz, 1H), 7.08 (s, 2H), 7.04 – 6.99 (m, 1H), 6.98 – 6.93 (m, 1H), 6.86 – 6.83 (m, 1H), 6.80 – 6.74 (m, 1H), 6.73 (d, J = 8.1 Hz, 1H), 6.68 – 6.60 (m, 1H), 5.60 (s, 1H), 5.14 (s, 1H), 4.97 (s, 1H), 2.21 (s, 3H), 1.37 (s, 18H); 13C{1H} NMR (100 MHz, CDCl3) δ 160.2 (J = 236 Hz), 152.6, 152.0, 136.9 (J = 12 Hz), 135.9, 132.1, 130.6, 129.8, 129.3, 128.5, 125.4, 124.2, 124.1, 123.6, 120.9 (J = 10 Hz), 118.9, 116.4, 108.5 (J = 24 Hz), 97.6, 97.3, 44.2,
ACS Paragon Plus Environment
The Journal of Organic Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
34.4, 30.4, 20.8; IR (KBr): 3628, 3587, 3566, 3446, 2958, 1653, 1506, 668, 649, 457 cm-1; HRMS (ESI-TOF) m/z: [M - H]- Calcd for C30H33FNO2 458.2501, found 458.2482; Enantiomeric ratio = 90:10, determined by HPLC (Daicel Chiralpak OD-H, hexane/isopropanol = 90:10, flow rate 1.0 mL/min, T = 30 °C, 254 nm); tR = 7.583 min (minor), tR = 8.677 min (major). (R)-2,6-di-tert-butyl-4-((6-fluoro-1H-indol-3-yl)(2-hydroxy-5-methoxyphenyl)methyl)ph enol (3ke): Preparative thin layer chromatography (petroleum ether/EtOAc = 5/1); 60% yield (28.5 mg); brown solid; m.p. 96.2-97.3 oC; [α]D20 = +27.5 (c 0.57, acetone); 1H NMR (400 MHz, CDCl3) δ 8.03 (s, 1H), 7.19 (dd, J = 8.7, 5.4 Hz, 1H), 7.07 (s, 2H), 7.04 – 6.98 (m, 1H), 6.82 – 6.74 (m, 2H), 6.72 – 6.68 (m, 1H), 6.67 – 6.64 (m, 1H), 6.57 (d, J = 3.0 Hz, 1H), 5.62 (s, 1H), 5.14 (s, 1H), 4.78 (s, 1H), 3.66 (s, 3H), 1.35 (s, 18H); 13C{1H} NMR (100 MHz, CDCl3) δ 160.1 (J = 237 Hz), 153.6, 152.6, 148.1, 136.8 (J = 13 Hz), 136.0, 131.9, 131.3, 125.4, 124.2, 124.1, 123.5, 120.8 (J = 10 Hz), 118.5, 117.1, 115.9, 112.5, 108.5 (J = 10 Hz), 97.6, 97.3, 55.6, 44.0, 34.4, 30.4; IR (KBr): 3648, 3628, 3587, 3420, 2957, 1558, 1506, 668, 649, 457 cm-1; HRMS (ESI-TOF) m/z: [M - H]- Calcd for C30H33FNO3 474.2450, found 474.2438; Enantiomeric ratio = 90:10, determined by HPLC (Daicel Chiralpak AD-H, hexane/isopropanol = 90:10, flow rate 1.0 mL/min, T = 30 °C, 254 nm); tR = 21.523 min (minor), tR = 26.657 min (major). (R)-4-((4-bromo-2-hydroxyphenyl)(6-fluoro-1H-indol-3-yl)methyl)-2,6-di-tert-butylphen ol (3kf): Preparative thin layer chromatography (petroleum ether/EtOAc = 5/1); 54% yield (28.1 mg); yellow solid; m.p. 99.8-101.7 oC; [α]D20 = +37.5 (c 0.56, acetone); 1H NMR (400 MHz, CDCl3) δ 8.01 (s, 1H), 7.22 – 7.11 (m, 1H), 7.10 – 6.94 (m, 5H), 6.90 – 6.72 (m, 2H), 6.64 (s, 1H), 5.59 (s, 1H), 5.25 (s, 1H), 5.17 (s, 1H), 1.37 (s, 18H); 13C{1H} NMR (100 MHz, CDCl3) δ 160.2 (J = 237 Hz), 155.0, 152.8, 136.9 (J = 12 Hz), 136.2, 131.4, 131.3, 129.2, 125.3, 124.3, 124.2,
ACS Paragon Plus Environment
Page 26 of 40
Page 27 of 40 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
The Journal of Organic Chemistry
123.8, 123.3, 120.8, 120.7 (J = 10 Hz), 119.7, 118.1, 108.6 (J = 24 Hz), 97.6 (J = 26 Hz), 43.4, 34.4, 30.3; IR (KBr): 3586, 3566, 3545, 3446, 2959, 1652, 1456, 668, 649, 472 cm-1; HRMS (ESI-TOF) m/z: [M - H]- Calcd for C29H30BrFNO2 522.1449, found 522.1430; [M - H]- Calcd for C29H30BrFNO2 524.1429, found 524.1437; Enantiomeric ratio = 90:10, determined by HPLC (Daicel Chiralpak AD-H, hexane/isopropanol = 80:20, flow rate 1.0 mL/min, T = 30 °C, 254 nm); tR = 5.427 min (major), tR = 6.247 min (minor). (R)-2,6-di-tert-butyl-4-((2-hydroxyphenyl)(1-methyl-1H-indol-3-yl)methyl)phenol (3pa): Preparative thin layer chromatography (petroleum ether/EtOAc = 5/1); 79% yield (35.0 mg); brown solid; m.p. 91.4-91.6 oC; 1H NMR (400 MHz, CDCl3) δ 7.33 – 7.27 (m, 2H), 7.25 – 7.19 (m, 1H), 7.19 – 7.12 (m, 1H), 7.16 (s, 2H), 7.05 – 6.98 (m, 2H), 6.90 – 6.82 (m, 2H), 6.57 (s, 1H), 5.67 (s, 1H), 5.18 (s, 1H), 5.14 (s, 1H), 3.73 (s, 3H), 1.37 (s, 18H); 13C{1H} NMR (100 MHz, CDCl3) δ 154.3, 152.5, 137.7, 135.8, 132.3, 130.2, 130.0, 128.3, 127.8, 127.3, 125.4, 122.0, 120.6, 120.0, 119.1, 116.6, 116.5, 109.2, 44.0, 34.4, 32.8, 30.3; IR (KBr):3657, 3650, 3620, 3588, 1701, 1559, 1473, 1457, 1436, 1229, 805, 739 cm-1; HRMS (ESI-TOF) m/z: [M - H]- Calcd for C30H34NO2 440.2595, found 440.2577; Enantiomeric ratio = 52:48, determined by HPLC (Daicel Chiralpak AD-H, hexane/isopropanol = 80:20, flow rate 1.0 mL/min, T = 30 °C, 254 nm); tR = 3.733 min (minor), tR = 4.193 min (major). 2-((1H-indol-3-yl)(phenyl)methyl)-4-methoxyphenol
(3ai):
Preparative
thin
layer
chromatography (petroleum ether/Acetone = 5/1); 98% yield (16.3 mg); white solid; m.p. 164.8-165.1 oC; [α]D20 = +11.7 (c 0.33, acetone); 1H NMR (400 MHz, CDCl3) δ 8.04 (s, 1H), 7.36 (d, J = 8.2 Hz, 1H), 7.34 – 7.26 (m, 5H), 7.26 – 7.23 (m, 1H), 7.22 – 7.15 (m, 1H), 7.06 – 6.98 (m, 1H), 6.79 (d, J = 8.7 Hz, 1H), 6.75 – 6.67 (m, 2H), 6.54 (d, J = 3.0 Hz, 1H), 5.80 (s, 1H), 4.67 (s,
ACS Paragon Plus Environment
The Journal of Organic Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
1H), 3.65 (s, 3H); 13C{1H} NMR (100 MHz, CDCl3) δ 153.7, 147.9, 142.1, 136.9, 131.1, 129.0, 128.6, 126.9, 126.8, 124.0, 122.6, 119.8, 117.5, 117.1, 116.3, 112.3, 111.2, 55.6, 43.6; IR (KBr): 3629,
3567, 3432, 2922, 1647, 1636, 1265, 1041, 745, 663 cm-1; HRMS (ESI-TOF) m/z: [M - H]- Calcd for C22H19NO2 328.1343, found 328.1338; Enantiomeric ratio = 87:13, determined by HPLC (Daicel Chiralpak AD-H, hexane/isopropanol = 70:30, flow rate 1.0 mL/min, T = 30 °C, 254 nm); tR = 8.143 min (major), tR = 29.433 min (minor). Procedure for the larger scale synthesis of product 3aa To the mixture of indole 1a (1.2 mmol), p-QM derivative 2a (1.0 mmol), chiral phosphoric acid 4e (0.1 mmol) and MgSO4 (1 g) was added acetone (20 mL), which was stirred at -30 oC for 12 h. After the completion of the reaction which was indicated by TLC, the reaction mixture was directly purified by flash column chromatography to afford pure product 3aa in 99% yield (423 mg). Procedure for the synthesis of product 5 Under argon atmosphere, compound 3fa (0.05 mmol), 4-chlorophenylboronic acid (0.075 mmol), CsCO3 (0.1 mmol), Pd(OAc)2 (0.0025 mmol) and butyldi-1-adamantylphosphine (0.003 mmol) were added to a dried tube. After adding DME (0.6 mL) to the reaction system, the reaction mixture was stirred at 80 oC for 12 h. After the completion of the reaction which was indicated by TLC, the reaction mixture was directly purified by preparative thin layer chromatography to give pure product 5. (R)-2,6-di-tert-butyl-4-((5-(4-chlorophenyl)-1H-indol-3-yl)(2-hydroxyphenyl)methyl)phe nol (5): Preparative thin layer chromatography (petroleum ether/EtOAc = 5/1); 79% yield (21.3 mg); brown solid; m.p. 102.2-103.7 oC; [α]D20 = +30.1 (c 0.43, acetone); 1H NMR (400 MHz,
ACS Paragon Plus Environment
Page 28 of 40
Page 29 of 40 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
The Journal of Organic Chemistry
CDCl3) δ 8.07 (s, 1H), 7.45 – 7.36 (m, 5H), 7.33 (d, J = 8.5 Hz, 2H), 7.20 – 7.13 (m, 1H), 7.11 (s, 2H), 7.04 (d, J = 7.3 Hz, 1H), 6.90 – 6.81 (m, 2H), 6.75 (s, 1H), 5.73 (s, 1H), 5.15 (s, 1H), 5.09 (s, 1H), 1.36 (s, 18H); 13C NMR (100 MHz, CDCl3) δ 154.1, 152.6, 140.9, 136.5, 136.0, 132.3, 132.2, 131.9, 130.1, 130.0, 128.7, 128.5, 128.0, 127.5, 125.5, 124.7, 122.0, 120.8, 119.1, 118.4, 116.5, 111.5, 43.7, 34.4, 30.4; IR (KBr): 3702, 3670, 3629, 3013, 2787, 1559, 1458, 1093, 1023, 807 cm-1; HRMS (ESI-TOF) m/z: [M - H]- Calcd for C35H35ClNO2 536.2362, found 536.2349; Enantiomeric ratio = 94:6, determined by HPLC (Daicel Chiralpak AD-H, hexane/isopropanol = 90:10, flow rate 1.0 mL/min, T = 30 °C, 254 nm); tR = 17.097 min (minor), tR = 21.770 min (major).
Acknowledgements
We are grateful for financial support from National Natural Science Foundation of China (21772069, 21702077 and 21831007), Natural Science Foundation of Jiangsu Province (BK20160003 and BK20170227), Young Medical Talents Project of Jiangsu Province (QNRC2016571), Six Kinds of Talents Project of Jiangsu Province (SWYY-025) and Undergraduate Project of Jiangsu Province (201810320049Z).
Supporting Information: 1H
and 13C NMR spectra of substrates 2b, 2f and products 3 and 5, HPLC spectra of products 3
and 5, X-ray single-crystal data for product 3ba (PDF) Single-crystal data of product 3ba (CIF)
ACS Paragon Plus Environment
The Journal of Organic Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
References and Footnotes (1) For some reviews on p-QMs: (a) Caruana, L.; Fochi, M.; Bernardi, L. The Emergence of Quinone Methides in Asymmetric Organocatalysis. Molecules 2015, 20, 11733-11764. (b) Parra, A.; Tortosa, M. para‐Quinone Methide: a New Player in Asymmetric Catalysis. ChemCatChem 2015, 7, 1524-1526. (c) Chauhan, P.; Kaya, U.; Enders, D. Advances in Organocatalytic 1,6‐Addition Reactions: Enantioselective Construction of Remote Stereogenic Centers. Adv. Synth. Catal. 2017, 359, 888-912. (d) Li, W.; Xu X.; Zhang P.; Li, P. Recent Advances in the Catalytic Enantioselective Reactions of para‐Quinone Methides, Chem. Asian J. 2018, 13, 2350-2359. (2) For early examples on catalytic asymmetric reactions involving p-QMs: (a) Chu, W.-D.; Zhang, L.-F. Bao, X.; Zhao, X.-H.; Zeng, C.; Du, J.-Y.; Zhang, G.-B.; Wang, F.-X.; Ma, X.-Y.; Fan, C.-A. Asymmetric Catalytic 1,6‐Conjugate Addition/Aromatization of para‐Quinone Methides: Enantioselective Introduction of Functionalized Diarylmethine Stereogenic Centers. Angew. Chem. Int. Ed. 2013, 52, 9229-9233. (b) Caruana, L.; Kniep, F.; Johansen, T. K.; Poulsen, P. H.; Jørgensen, K. A. A New Organocatalytic Concept for Asymmetric α-Alkylation of Aldehydes. J. Am. Chem. Soc. 2014, 136, 15929-15932. (3) For some representative examples on catalytic asymmetric reactions involving p-QMs: (a) Wang, Z.; Wong Y. F.; Sun, J. Catalytic Asymmetric 1,6‐Conjugate Addition of para‐Quinone Methides: Formation of All‐Carbon Quaternary Stereocenters. Angew. Chem. Int. Ed. 2015, 54, 13711-13714. (b) Lou, Y.; Cao, P.; Jia, T.; Zhang, Y.; Wang M.; Liao, J. Copper‐Catalyzed Enantioselective 1,6‐Boration of para‐Quinone Methides and Efficient Transformation of gem‐Diarylmethine Boronates to Triarylmethanes. Angew. Chem. Int. Ed. 2015, 54,
ACS Paragon Plus Environment
Page 30 of 40
Page 31 of 40 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
The Journal of Organic Chemistry
12134-12138. (c) Zhao, K.; Zhi, Y.; Wang A.; Enders, D. Asymmetric Organocatalytic Synthesis of 3-Diarylmethine-Substituted Oxindoles Bearing a Quaternary Stereocenter via 1,6-Conjugate Addition to para-Quinone Methides. ACS Catal. 2016, 6, 657-660. (d) Dong, N.; Zhang, Z.-P.; Xue, X.-S. Li X.; Cheng, J.-P. Phosphoric Acid Catalyzed Asymmetric 1,6‐Conjugate Addition of Thioacetic Acid to para‐Quinone Methides. Angew. Chem. Int. Ed. 2016, 55, 1460-1464. (e) He, F.-S.; Jin, J.-H.; Yang, Z.-T.; Yu, X.; Fossey, J. S.; Deng, W.-P. Direct Asymmetric Synthesis of β-Bis-Aryl-α-Amino Acid Esters via Enantioselective Copper-Catalyzed Addition of p-Quinone Methides. ACS Catal. 2016, 6, 652-656. (f) Li, X.; Xu, X.; Wei, W.; Lin, A.; Yao, H Organocatalyzed Asymmetric 1,6-Conjugate Addition of para-Quinone Methides with Dicyanoolefins. Org. Lett. 2016, 18, 428-431. (g) Deng, Y.-H.; Zhang, X.-Z.; Yu, K.-Y.; Yan, X.; Du, J.-Y.; Huang H.; Fan, C.-A. Bifunctional Tertiary Amine-Squaramide Catalyzed Asymmetric Catalytic 1,6-Conjugate Addition/Aromatization of para-Quinone Methides with Oxindoles. Chem. Commun. 2016, 52, 4183-4186. (h) Ma, C.; Huang, Y.; Zhao, Y. Stereoselective 1,6-Conjugate Addition/Annulation of para-Quinone Methides with Vinyl Epoxides/Cyclopropanes. ACS Catal. 2016, 6, 6408-6412; (4) For some recent examples on catalytic asymmetric reactions involving p-QMs: (a) Li, S.; Liu, Y.; Huang, B.; Zhou, T.; Tao, H.; Xiao, Y.; Liu, L.; Zhang, J. Phosphine-Catalyzed Asymmetric Intermolecular Cross-Vinylogous Rauhut–Currier Reactions of Vinyl Ketones with para-Quinone Methides. ACS Catal. 2017, 7, 2805-2809. (b) Li, W.; Xu, X.; Liu, Y.; Gao, H.; Cheng, Y.; Li, P. Enantioselective Organocatalytic 1, 6-Addition of Azlactones to para-Quinone Methides: An access to α,α-Disubstituted and β,β-Diaryl-α-Amino Acid Esters. Org. Lett. 2018, 20, 1142-1145. (c) Rahman, A.; Zhou, Q.; Lin, X. Asymmetric Organocatalytic
ACS Paragon Plus Environment
The Journal of Organic Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
Synthesis of Chiral 3,3-Disubstituted Oxindoles via a 1,6-Conjugate Addition Reaction. Org. Biomol. Chem. 2018, 16, 5301-5309. (5) For some reviews on o-QMs: (a) van de Water, R. W.; Pettus, T. R. R. o-Quinone Methides: Intermediates Underdeveloped and Underutilized in Organic Synthesis. Tetrahedron 2002, 58, 5367-5405. (b) Pathak, T. P.; Sigman, M. S. Applications of ortho-Quinone Methide Intermediates in Catalysis and Asymmetric Synthesis. J. Org. Chem. 2011, 76, 9210-9215. (c) Willis, N. J.; Bray, C. D. ortho‐Quinone Methides in Natural Product Synthesis. Chem. Eur. J. 2012, 18, 9160-9173. (d) Bai, W.-J.; David, J. G.; Feng, Z.-G.; Weaver, M. G.; Wu, K.-L.; Pettus, T. R. R. the Domestication of ortho-Quinone Methides. Acc. Chem. Res. 2014, 47, 3655-3664. (e) Singh, M. S.; Nagaraju, A.; Anand, N.; Chowdhury, S. ortho-Quinone Methide (o-QM): a Highly Reactive, Ephemeral and Versatile Intermediate in Organic Synthesis. RSC Adv. 2014, 4, 55924-55959. (f) Wang, Z.; Sun, J. Recent Advances in Catalytic Asymmetric Reactions of o-Quinone Methides. Synthesis 2015, 47, 3629-3644. (g) Jaworski, A. A.; Scheidt, K. A. Emerging Roles of in Situ Generated Quinone Methides in Metal-Free Catalysis. J. Org. Chem. 2016, 81, 10145-10153. (h) Yang, B.; Gao, S. Recent Advances in the Application of Diels–Alder Reactions Involving o-Quinodimethanes, Aza-o-Quinone Methides and o-Quinone Methides in Natural Product Total Synthesis. Chem. Soc. Rev. 2018, 47, 7926-7953. (6) For some early examples on catalytic asymmetric reactions involving o-QMs: (a) Yu, Z.-P.; Liu, X.-H.; Dong, Z.-H.; Xie, M.-S.; Feng, X.-M. An N,N′‐Dioxide/In(OTf)3 Catalyst for the Asymmetric Hetero‐Diels–Alder Reaction Between Danishefsky's Dienes and Aldehydes: Application in the Total Synthesis of Triketide. Angew. Chem. Int. Ed. 2008, 47, 1308-1311. (b) Alden-Danforth, E.; Scerba, M. T.; Lectka, T. Asymmetric Cycloadditions of o-Quinone
ACS Paragon Plus Environment
Page 32 of 40
Page 33 of 40 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
The Journal of Organic Chemistry
Methides Employing Chiral Ammonium Fluoride Precatalysts. Org. Lett. 2008, 10, 4951-4953. (c) Luan,Y.; Schaus, S. E. Enantioselective Addition of Boronates to o-Quinone Methides Catalyzed by Chiral Biphenols. J. Am. Chem. Soc. 2012, 134, 19965-19968. (d) Izquierdo, J.; Orue, A.; Scheidt, K. A. A Dual Lewis Base Activation Strategy for Enantioselective Carbene-Catalyzed Annulations. J. Am. Chem. Soc. 2013, 135, 10634-10637. (e) Lv, H.; Jia, W.-Q.; Sun, L.-H.; Ye, S. N‐Heterocyclic Carbene Catalyzed [4+3] Annulation of Enals and o‐Quinone Methides: Highly Enantioselective Synthesis of Benzo‐ε‐Lactones. Angew. Chem. 2013, 125, 8769-8772. (f) El-Sepelgy, O.; Haseloff, S.; Alamsetti, S. K.; Schneider, C. Brønsted Acid Catalyzed, Conjugate Addition of β‐Dicarbonyls to in Situ Generated ortho‐Quinone Methides—Enantioselective Synthesis of 4‐Aryl‐4H‐Chromenes. Angew. Chem. Int. Ed. 2014, 53, 7923-7927. (g) Hsiao, C.-C.; Liao, H.-H.; Rueping, M Enantio‐ and Diastereoselective Access
to
Distant
Stereocenters
Embedded
within
Tetrahydroxanthenes:
Utilizing
ortho‐Quinone Methides as Reactive Intermediates in Asymmetric Brønsted Acid Catalysis. Angew. Chem. Int. Ed. 2014, 53, 13258-13263. (7) For some examples on catalytic asymmetric conjugate additions involving o-QMs: (a) Wilcke, D.; Herdtweck, E.; Bach, T. Enantioselective Brønsted Acid Catalysis in the Friedel-Crafts Reaction of Indoles with Secondary ortho-Hydroxybenzylic Alcohols. Synlett 2011, 1235-1238. (b) Saha, S.; Alamsetti, S. K.; Schneider, C. Chiral Brønsted Acid-Catalyzed Friedel–Crafts Alkylation of Electron-Rich Arenes with in Situ-Generated ortho-Quinone Methides: Highly Enantioselective Synthesis of Diarylindolylmethanes and Triarylmethanes. Chem. Commun. 2015, 51, 1461-1464. (c) Zhao, W.; Wang, Z.; Chu, B.; Sun, J. Enantioselective Formation of all-Carbon Quaternary Stereocenters from Indoles and Tertiary Alcohols Bearing a Directing
ACS Paragon Plus Environment
The Journal of Organic Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
Page 34 of 40
Group. Angew. Chem. Int. Ed. 2015, 54, 1910-1913. (d) Liao, H.-H.; Chatupheeraphat, A.; Hsiao, C.-C.; Atodiresei, I.; Rueping, M. Asymmetric Bronsted Acid Catalyzed Synthesis of Triarylmethanes-Construction of Communesin and Spiroindoline Scaffolds. Angew. Chem., Int. Ed. 2015, 54, 15540-15544. (e) Lai, Z.; Wang, Z.; Sun, J. Organocatalytic Asymmetric Nucleophilic Addition to o-Quinone Methides by Alcohols. Org. Lett. 2015, 17, 6058-6061. For some recent examples on catalytic asymmetric cyclizations involving o-QMs: (f) Hu, H.; Liu, Y.; Guo, J.; Lin, L.; Xu, Y.; Liu, X.; Feng, X. Enantioselective Synthesis of Dihydrocoumarin Derivatives
by
Chiral
Scandium(III)-Complex
Catalyzed
Inverse-Electron-Demand
Hetero-Diels–Alder Reaction. Chem. Commun. 2015, 51, 3835-3837. (g) Zhang, J.; Liu, X.; Guo, S.; He, C.; Xiao, W.; Lin, L.; Feng, X. Enantioselective Formal [4 + 2] Annulation of ortho-Quinone Methides with ortho-Hydroxyphenyl α,β-Unsaturated Compounds. J. Org. Chem. 2018, 83, 10175-10185. (h) Zhang, J.; Lin, L.; He, C.; Xiong, Q.; Liu, X.; Feng, X. A Chiral Scandium-Complex-Catalyzed Asymmetric Inverse-Electron-Demand Oxa-Diels–Alder Reaction of o-Quinone Methides with Fulvenes. Chem. Commun., 2018, 54, 74-77. (8) For some examples on catalytic asymmetric (4+2) cyclizations involving o-QMs: (a) Saha, S.; Schneider, C. Bronsted Acid-Catalyzed, Highly Enantioselective Addition of Enamides to in Situ-Generated
ortho-Quinone
Methides:
A
Domino
Approach
to
Complex
Acetamidotetrahydroxanthenes. Chem. Eur. J. 2015, 21, 2348-2352. (b) Saha, S.; Schneider, C. Directing Group Assisted Nucleophilic Substitution of Propargylic Alcohols via o-Quinone Methide Intermediates: Bronsted Acid Catalyzed, Highly Enantio- and Diastereoselective Synthesis of 7-Alkynyl-12a-Acetamido-Substituted Benzoxanthenes. Org. Lett. 2015, 17, 648-651. (c) Zhao, J.-J.; Sun, S.-B.; He, S.-H.; Wu, Q.; Shi, F. Catalytic Asymmetric
ACS Paragon Plus Environment
Page 35 of 40 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
The Journal of Organic Chemistry
Inverse-Electron-Demand Oxa-Diels-Alder Reaction of in Situ Generated ortho-Quinone Methides with 3-Methyl-2-Vinylindoles. Angew. Chem. Int. Ed. 2015, 54, 5460-5464. (d) Hsiao, C.-C.; Raja, S.; Liao, H.-H.; Atodiresei, I.; Rueping, M. ortho-Quinone Methides as Reactive Intermediates in Asymmetric Bronsted Acid Catalyzed Cycloadditions with Unactivated Alkenes by Exclusive Activation of the Electrophile. Angew. Chem. Int. Ed. 2015, 54, 5762-5765. (e) Tsui, G. C.; Liu, L.; List, B. The Organocatalytic Asymmetric Prins Cyclization. Angew. Chem. Int. Ed. 2015, 54, 7703-7706. (f) Alamsetti, S. K.; Spanka, M.; Schneider, C. Synergistic Rhodium/Phosphoric Acid Catalysis for the Enantioselective Addition of Oxonium Ylides to ortho-Quinone Methides. Angew. Chem. Int. Ed. 2016, 55, 2392-2396. (g) Jeong, H. J.; Kim, D. Y.; Enantioselective Decarboxylative Alkylation of β-Keto Acids to ortho-Quinone Methides as Reactive Intermediates: Asymmetric Synthesis of 2,4-Diaryl-1-Benzopyrans. Org. Lett. 2018, 20, 2944-2047. (h) Spanka, M.; Schneider, C. Phosphoric Acid Catalyzed Aldehyde Addition to in Situ Generated o-Quinone Methides: An Enantio- and Diastereoselective Entry toward cis-3,4-Diaryl Dihydrocoumarins. Org. Lett. 2018, 20, 4769-4772. (i) Göricke, F.; Schneider, C. Palladium-Catalyzed Enantioselective Addition of Chiral Metal Enolates to in Situ Generated ortho-Quinone Methides. Angew. Chem. Int. Ed. 2018, 57, 14736-14741. (9) For a catalytic asymmetric (4+1) cyclization: Liu, L.; Yuan, Z.; Pan, R.; Zeng, Y.; Lin, A.; Yao, H.; Huang, Y. 1,6-Conjugated Addition-Mediated [4 + 1] Annulation: An Approach to 2,3-Dihydrobenzofurans. Org. Chem. Front. 2018, 5, 623-628. (10) For catalytic asymmetric (4+2) cyclizations: (a) Zhao, K.; Zhi, Y.; Shu, T.; Valkonen, A.; Rissanen, K.; Enders, D. Organocatalytic Domino Oxa-Michael/1,6-Addition Reactions:
ACS Paragon Plus Environment
The Journal of Organic Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
Page 36 of 40
Asymmetric Synthesis of Chromans Bearing Oxindole Scaffolds. Angew. Chem. Int. Ed. 2016, 55, 12104-12108. (b) Zhang, L.-L.; Liu, Y.; Liu, K.; Liu, Z.-T.; He, N.-N.; Li, W.-J. Asymmetric Synthesis of Dihydrocoumarins via the Organocatalytic Hetero-Diels-Alder Reaction of ortho-Quinone Methides. Org. Biomol. Chem. 2017, 15, 8743-8747. (c) Zhang, L.-L.; Zhou, X.; Li, P.-F.; Liu, Z.-T.; Liu, Y.; Sun, Y.; Li, W.-J. Asymmetric Synthesis of Chromene Skeletons via Organocatalytic Domino Reactions of in Situ Generated ortho-Quinone Methide with Malononitrile and β-Functionalized Ketone. RSC Adv. 2017, 7, 39216-39220. (d) Zhang, Z.-P.; Xie, K.-X.; Yang, C.; Li, M.; Li, X. Asymmetric Synthesis of Dihydrocoumarins
through
Chiral
Phosphoric
Acid-Catalyzed
Cycloannulation
of
para-Quinone Methides and Azlactones. J. Org. Chem. 2018, 83, 364-373.(e) Jiang, X.-L.; Wu, S.-F.; Wang, J.-R.; Mei, G.-J.; Shi, F. Catalytic Asymmetric [4+2] Cyclization of para-Quinone Methide Derivatives with 3-Alkyl-2-vinylindoles. Adv. Synth. Catal. 2018, 360, 4225-4235. (11) For catalytic asymmetric (4+3) cyclizations: (a) Li, W.-J.; Yuan, H.-J.; Liu, Z.-T.; Zhang, Z.-Y.; Cheng, Y.-Y.; Li, P.-F. NHC-Catalyzed Enantioselective [4+3] Cycloaddition of ortho-Hydroxyphenyl Substituted para-Quinone Methides with Isatin-Derived Enals. Adv. Synth. Catal. 2018, 360, 2460-2464. (b) Liu, Q.; Li, S.; Chen, X.-Y.; Rissanen, K.; Enders, D. Asymmetric Synthesis of Spiro-Oxindole-ε-Lactones through N-Heterocyclic Carbene Catalysis. Org. Lett. 2018, 20, 3622-3626. (c) Jiang, F.; Yuan, F.-R.; Jin, L.-W.; Mei, G.-J.; Shi, F. Metal-Catalyzed (4 + 3) Cyclization of Vinyl Aziridines with para-Quinone Methide Derivatives. ACS Catal. 2018, 8, 10234-10240. (d) Sun, M.; Ma, C.; Zhou, S.-J.; Lou, S.-F.; Xiao, J.; Jiao, Y.; Shi, F. Catalytic Asymmetric (4+3) Cyclizations of In Situ Generated ortho-Quinone Methides with 2-Indolylmethanols, Angew. Chem. Int. Ed. 2019, DOI:
ACS Paragon Plus Environment
Page 37 of 40 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
The Journal of Organic Chemistry
10.1002/anie.201901955. (12) (a) Zhang, L.; Yuan, H.; Lin, W.; Cheng, Y.; Li, P.; Li, W. Catalyst-Controlled Diastereodivergent Construction of Vicinal Sulfur-Functionalized Quaternary and Tertiary Stereocenters. Org. Lett. 2018, 20, 4970-4974. (b) Huang, G.-B.; Huang, W.-H. Guo, J.; Xu, D.-L.; Qu, X.-C.; Zhai, P.-H.; Zheng, X.-H.; Weng, J.; Lu, G. Enantioselective Synthesis of Triarylmethanes via Organocatalytic 1,6‐Addition of Arylboronic Acids to para-Quinone Methides. Adv. Synth. Catal. 2019, 361, 1241-1246. (13) (a) Zhang, H.-H.; Wang, C.-S.; Li, C.; Mei, G.-J.; Li, Y.; Shi, F. Design and Enantioselective Construction of Axially Chiral Naphthyl-Indole Skeletons. Angew. Chem. Int. Ed. 2017, 56, 116-121. (b) Ma, C.; Zhou, J.-Y.; Zhang, Y.-Z.; Mei, G.-J.; Shi, F. Catalytic Asymmetric [2+3] Cyclizations of Azlactones with Azonaphthalenes. Angew. Chem. Int. Ed. 2018, 57, 5398-5402. (c) Ma, C.; Jiang, F.; Sheng, F.-T.; Jiao, Y.; Mei, G.-J.; Shi, F. Design and Catalytic Asymmetric Construction of Axially Chiral 3,3’-Bisindole Skeletons. Angew. Chem. Int. Ed. 2019, 58, 3014-3020. (d) Ma, C.; Zhou, J.-Y.; Zhang, Y.-Z.; Jiao, Y.; Mei, G.-J.; Shi, F. Synergistic-Catalysis-Enabled Reaction of 2-Indolymethanols with Oxonium Ylides for the Construction of 3-Indolyl-3-Alkoxy Oxindole Frameworks. Chem. Asian J. 2018, 13, 2549-2558. (14) For some reviews: (a) Akiyama, T. Stronger Bronsted Acids. Chem. Rev. 2007, 107, 5744-5758. (b) Terada, M. Binaphthol-Derived Phosphoric Acid as a Versatile Catalyst for Enantioselective Carbon-Carbon Bond Forming Reactions. Chem. Commun. 2008, 4097-4112. (c) Terada, M. Chiral Phosphoric Acids as Versatile Catalysts for Enantioselective Transformations. Synthesis 2010, 1929-1982; (d) Zamfir, A.; Schenker, S.; Freund, M.;
ACS Paragon Plus Environment
The Journal of Organic Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
Tsogoeva, S. B. Chiral BINOL-Derived Phosphoric Acids: Privileged Bronsted Acid Organocatalysts for C-C Bond Formation Reactions. Org. Biomol. Chem. 2010, 8, 5262-5276. (e) Yu, J.; Shi, F.; Gong, L.-Z. Bronsted-Acid-Catalyzed Asymmetric Multicomponent Reactions for the Facile Synthesis of Highly Enantioenriched Structurally Diverse Nitrogenous Heterocycles. Acc. Chem. Res. 2011, 44, 1156-1171. (f) Parmar, D.; Sugiono, E.; Raja, S.; Rueping, M. Complete Field Guide to Asymmetric BINOL-Phosphate Derived Bronsted Acid and Metal Catalysis: History and Classification by Mode of Activation; Bronsted Acidity, Hydrogen Bonding, Ion Pairing, and Metal Phosphates. Chem. Rev. 2014, 114, 9047-9153. (g) Wu, H.; He, Y.-P.; Shi, F. Recent Advances in Chiral Phosphoric Acid Catalyzed Asymmetric Reactions for the Synthesis of Enantiopure Indole Derivatives. Synthesis 2015, 47, 1990-2016. (h) Held, F. E.; Grau, D.; Tsogoeva, S. B. Enantioselective Cycloaddition Reactions Catalyzed by BINOL-Derived Phosphoric Acids and N-Triflyl Phosphoramides: Recent Advances. Molecules 2015, 20, 16103-16126. (15) For some recent examples on catalytic asymmetric additions of indoles: (a) Zhang, Y.-C.; Zhao, J.-J.; Jiang, F.; Sun, S.-B.; Shi, F. Organocatalytic Asymmetric Arylative Dearomatization of 2,3-Disubstituted Indoles Enabled by Tandem Reactions, Angew. Chem. Int. Ed. 2014, 53, 13912-13915. (b) Zhang, X.; Zhang, J.; Lin, L.; Zheng, H.; Wu, W.; Liu, X.; Feng, X. Chiral N,N′-Dioxide-Zinc(II) Complex-Catalyzed Asymmetric Aza-Friedel–Crafts Reaction of Isatin-Derived Ketimines with Indoles. Adv. Synth. Catal. 2016, 358, 3021-3026. (c) Jiang, F.; Zhao, D.; Yang, X.; Yuan, F.-R.; Mei, G.-J.; Shi, F. Catalyst-Controlled Chemoselective and Enantioselective Reactions of Tryptophols with Isatin-Derived Imines, ACS Catal. 2017, 7, 6984-6989. (d) Xu, Y.; Chang, F.; Cao, W.; Liu, X.; Feng, X. Catalytic
ACS Paragon Plus Environment
Page 38 of 40
Page 39 of 40 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
The Journal of Organic Chemistry
Asymmetric Chemodivergent C2 Alkylation and [3 + 2]-Cycloaddition of 3-Methylindoles with Aziridines. ACS Catal. 2018, 8, 10261-10266. (e) Wang, H.-Q.; Xu, M.-M.; Wan, Y.; Mao, Y.-J.; Mei, G.-J.; Shi, F. Application of 7-Indolylmethanols in Catalytic Asymmetric Arylations with Tryptamines: Enantioselective Synthesis of 7-indolylmethanes, Adv. Synth. Catal. 2018, 360, 1850-1860. (16) (a) Wang, Z.; Ai, F.; Wang, Z.; Zhao, W.; Zhu, G.; Lin, Z.; Sun, J. Organocatalytic Asymmetric Synthesis of 1,1-Diarylethanes by Transfer Hydrogenation. J. Am. Chem. Soc. 2015, 137, 383-389. (b) Yin, L.; Hu, Q.; Hartmann, R. W. Tetrahydropyrroloquinolinone Type Dual Inhibitors of Aromatase/Aldosterone Synthase as a Novel Strategy for Breast Cancer Patients with Elevated Cardiovascular Risks. J. Med. Chem. 2013, 56, 460−470. (c) Wood, P. M.; Woo, L. W. L.; Labrosse, J. R.; Trusselle, M. N.; Abbate,S.; Longhi, G.; Castiglioni, E.; Lebon, F.; Purohit, A.; Reed, M. J.; Potter, B. V. L. Chiral Aromatase and Dual Aromatase−Steroid Sulfatase Inhibitors from the Letrozole Template: Synthesis, Absolute Configuration, and in vitro Activity. J. Med. Chem. 2008, 51, 4226−4238. (d) Shagufta, Srivastava, A. K.; R. Sharma, R.; Mishra, R.; Balapure, A. K.; Murthy, P. S. R.; Panda, G. Substituted Phenanthrenes with Basic Amino Side Chains: A New Series of Anti-Breast Cancer Agents. Bioorg. Med. Chem. 2006, 14, 1497−1505. (e) Hucke, O.; Gelb, M. H.; Verlinde, C. L.; Buckner, F. S. The Protein Farnesyltransferase Inhibitor Tipifarnib as a New Lead for the Development of Drugs against Chagas Disease. J. Med. Chem. 2005, 48, 5415−5418. (17) CCDC 1902126 for 3ba, see the Supporting Information for details. (18) Wang, L.; Jia, Y.-X.; Zhang, J.-M.; Qian, C.; Chen, X.-Z. Improved Synthesis of 4-Benzylidene-2,6-di-tert-Butylyclohexa-2,5-Dienone and its Derivatives. Monatsh. Chem.
ACS Paragon Plus Environment
The Journal of Organic Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
2014, 145, 1941-1945. (19) Wang, P.-F.; Hu, H.-Y.; Wang, Y. Novel Photolabile Protecting Group for Carbonyl Compounds. Org. Lett. 2007, 9, 1533-1535. (20) Zhou, T.; Li, S.-H.; Huang, B.; Li, C.; Zhao, Y.; Chen, J.-Q.; Chen, A.-L.; Xiao,Y.-J.; Liu L.; Zhang, J.-L. Phosphine-Catalyzed Friedel–Crafts Reaction of Naphthols with para-Quinone Methides: Expedient Access to Triarylmethanes. Org. Biomol. Chem. 2017, 15, 4941–4945.
ACS Paragon Plus Environment
Page 40 of 40