Synlett 2015; 26(01): 95-100
DOI: 10.1055/s-0034-1379494
letter
© Georg Thieme Verlag Stuttgart · New York

2-Aminobenzimidazole Organocatalyzed Asymmetric Amination of Cyclic 1,3-Dicarbonyl Compounds

Paz Trillo
Departamento de Química Orgánica, Facultad de Ciencias, and Instituto de Síntesis Orgánica (ISO), Universidad de Alicante, Apdo 99, 03080 Alicante, Spain   Fax: +34(96)5903549   Email: alex.baeza@ua.es
,
Melania Gómez-Martínez
Departamento de Química Orgánica, Facultad de Ciencias, and Instituto de Síntesis Orgánica (ISO), Universidad de Alicante, Apdo 99, 03080 Alicante, Spain   Fax: +34(96)5903549   Email: alex.baeza@ua.es
,
Diego A. Alonso
Departamento de Química Orgánica, Facultad de Ciencias, and Instituto de Síntesis Orgánica (ISO), Universidad de Alicante, Apdo 99, 03080 Alicante, Spain   Fax: +34(96)5903549   Email: alex.baeza@ua.es
,
Alejandro Baeza*
Departamento de Química Orgánica, Facultad de Ciencias, and Instituto de Síntesis Orgánica (ISO), Universidad de Alicante, Apdo 99, 03080 Alicante, Spain   Fax: +34(96)5903549   Email: alex.baeza@ua.es
› Author Affiliations
Further Information

Publication History

Received: 19 September 2014

Accepted after revision: 18 October 2014

Publication Date:
27 November 2014 (online)

 


Dedicated to the memory of Prof. Carlos F. Barbas III

Abstract

The use of a trans-cyclohexanediamine benzimidazole derivative as a hydrogen-bond catalyst for the electrophilic amination of cyclic 1,3-dicarbonyl compounds is herein presented. High yields and enantioselectivities varying from moderate to excellent are generally obtained using mild reaction conditions and as low as 1 mol% of catalyst loading.


#

The construction of chiral quaternary stereocenters bearing an amine moiety represents an important reaction in synthetic organic chemistry due to the range of compounds possessing such a structure in nature, most of them having biological and pharmaceutical activity.[1] In this sense, a wide variety of methods has been developed to gain access to these motifs. Among them, the asymmetric electrophilic amination of prochiral carbonyl compounds employing diazocarboxylates as nitrogen source is a simple and straightforward method since the latter reagents are bench-stable and readily available.[2]

Particularly interesting is the catalytic asymmetric α-amination of prochiral 1,3-dicarbonyl compounds, since the highly functionalized resulting structures can be further transformed and elaborated.[1] [2] [3] In this regard, several strategies have recently been developed to accomplish this transformation.[3] Thus, since the pioneer work of Jørgensen and co-workers using a copper(II)-box catalytic system[4] different methods, not only metal-catalyzed,[5] but also employing organocatalysts[6] [7] [8] have been reported.

Table 1 Catalyst Screeninga

Entry

Catalyst

Conversion (%)b

ee (%)c

1

I

>95

68

2

II

>95

90

3

III

>95

75

4

IV

>95

85

5

V

90

10

6

VI

80

rac

7

VII

90

50

8

VIII

15

rac

9

IX

70

65

10

X

<15

n.d.

a Reaction conditions: 1a (0.10 mmol), 2a (0.15 mmol), catalyst (10 mol%) in toluene (1 mL).

b Determined by 1H NMR analysis from the reaction crude.

c Determined by chiral HPLC (Daicel Chiralpak IA, see Supporting Information for details).

Recently, we have been interested in the use of trans-­cyclohexanediamine benzimidazole derivatives as hydrogen-bonding organocatalysts in various organic transformations.[9] Therefore, we decided to explore the performance of these catalysts in the electrophilic amination of 1,3-dicarbonyl compounds. The results of this study are disclosed herein.

First, the search for the appropriate catalyst to carry out this reaction was tackled using ethyl 2-oxocyclopentane­carboxylate (1a) and di-tert-butylazodicarboxylate (2a) as model substrates (Table [1]) and different trans-cyclohexanediamine benzimidazole derivatives IVIII (Table [1], entries 1–8). The more basic catalysts IIV afforded the corresponding amination product 3aa in high conversions and enantioselectivities (Table [1], entries 1–4), reaching up to 90% ee in the case of dimethylamino derivative II (Table [1], entry 2). The presence of less basic nitrogen in the catalysts, as is the case of V and VI, resulted in a dramatic drop of enantioselection (Table [1], entries 5 and 6). Next, bis(2-aminobenzoimidazole) derivatives VII and VIII were also evaluated, but poorer results were observed in both conversion and enantioselectivity (Table [1], entries 7 and 8). Finally, for the sake of comparison, Takemoto’s thiourea catalyst IX and the bisthiourea X were also evaluated but moderate conversion and enantioselectivity and low conversion were observed, respectively (Table [1], entries 9 and 10)

Once the organocatalyst screening revealed that benz­imidazole II provided the best results, further optimization of reaction conditions was performed (Table [2]). Firstly, different solvents were tested (Table [2], entries 1–7) obtaining the best results in terms of both conversion and enantioselectivity when toluene, diethyl ether and hexane were employed (Table [2], entries 1, 3 and 6). With these solvents the temperature influence was evaluated. Thus, at 0 °C, the same results were observed (Table [2], entries 8–10) and lowering the temperature to –20 °C resulted in lower conversions with enantioselectivity remaining the same. At this point, and since the influence of the temperature was negligible, we decided to continue the optimization at room temperature, using diethyl ether as solvent for solubility reasons. For substrate efficiency, we carried out the reaction using 1.05 equivalents of 2a and the same results were observed (Table [2], entry 11). Next, the effect of concentration of 1a was studied, and exactly the same results were obtained using 0.2 M and 0.05 M reaction solutions (Table [2], entries 12 and 13); therefore we chose the latter as optimal concentration. Finally, we tried to reduce the amount of catalyst and we observed that not only 5 mol% (Table [2], entry 14), but also as low as 1 mol% of catalyst loading was enough to promote the reaction with full conversion and excellent enantioselectivity (Table [2], entry 15).

Table 2 Optimization of Reaction Parametersa

Entry

Solvent

Temp (°C)

Conv. (%)b

ee (%)c

1

toluene

25

>95

90

2

CH2Cl2

25

45

90

3

Et2O

25

>95

92

4

THF

25

90

86

5

TBME

25

95

87

6

hexane

25

>95

92

7

MeOH

25

90

52

8

toluene

0

>95

90

9

Et2O

0

>95

92

10

hexane

0

95

92

11d

Et2O

25

>95

92

12d,e

Et2O

25

>95

92

13d,f

Et2O

25

>95

92

14d,e,g

Et2O

25

>95

92

15d,e,h

Et2O

25

>95

92

a Reaction conditions: 1a (0.10 mmol), 2a (0.15 mmol), II (10 mol%) in solvent (1.0 mL).

b Determined by 1H NMR analysis from the reaction crude.

c Determined by chiral HPLC (Daicel Chiralpak IA, see Supporting Information for details).

d Amount of 2a used was 0.105 mmol (1.05 equiv).

e Volume of Et2O ([1a] = 0.2 M) used was 0.5 mL.

f Volume of Et2O ([1a] = 0.05 M) used was 2 mL.

g Conditions: 5 mol% of II, 0.5 mL of Et2O.

h Conditions:1 mol% of II, 0.5 mL of Et2O.

Then, with the optimal reaction conditions established (Table [2], entry 15), we decided to study the influence of the diazocarboxylate structure (Scheme [1]). Thus, β-keto ester 1a was allowed to react with different alkyl diazocarboxylates 2bd but in all cases the results turned out to be worse than that in the case of 2a.

Zoom Image
Scheme 1 Study of different diazocarboxylates

With the reaction parameters optimized we next explored substrate scope (Table [3]).[10] First, cyclic β-keto esters were examined. As previously noted 1a yielded the desired product in high yields and with 92% ee (Table [3], entry 1). Surprisingly, when the six-membered analogue was submitted to the optimal reaction conditions it failed completely, even when higher catalyst loadings were investigated (Table [3], entry 2). The use of benzocondensed substrate 1c rendered the amination product 3ca in high yield and moderate enantioselectivity (Table [3], entry 3). In contrast, high optical purity along with high yield were obtained with keto ester 1d (Table [3], entry 4). Cyclic β-amido ester was also examined but a disappointingly low enantio­selectivity was obtained despite several reaction conditions tested (Table [3], entry 5).

Next, the more reactive cyclic 1,3-diketones were considered. The five-membered ring diketone 1f was firstly tested obtaining good yield and moderate enantioselectivity (Table [3], entry 6). In this case, the yield was slightly increased by using 5 mol% catalyst loading. As already observed in the case of keto esters, the six-membered 1,3-diketone 1g afforded low conversions, regardless of the reaction conditions tested (Table [3], entry 7). The corresponding benzocondensed analogues 1h and 1i were also evaluated and, in both cases, gave high yields although moderate enantioselectivities for the corresponding amination products were achieved (Table [3], entry 7). In both cases a slight increase of the optical purity was observed by lowering the temperature. Finally, compounds 1j and 1k gave rise to the corresponding amination products, 3ja and 3ka respectively, in good yields but with poor ee values.

Table 3 Substrate Scopea

Entry

1

3

Yield (%)b

ee (%)c

1

1a

3aa

99

92

2

1b

3ba

<10

n.d.

3

1c

3ca

97

45

4

1d

3da

98

88

5

1e

3ea

87

20

6

1f

3fa

75 (82)d

50

7

1g

3ga

<10

n.d.

8

1h

3ha

88 (70)e

27 (35)e

9

1i

3ia

70 (68)f

48 (54)f

10

1j

3ja

66

26

11

1k

3ka

89

25

a Unless otherwise stated, the reaction conditions were: 1a (0.20 mmol), 2a (0.21 mmol), II (1 mol%) in Et2O (1 mL), 25 °C.

b Isolated yield after column chromatography.

c Determined by chiral HPLC (see Supporting Information for details).

d The reaction was carried out using 5 mol% of II.

e The reaction was carried out at –20 °C.

f The reaction was carried out at –50 °C.

Different linear β-keto ester and 1,3-diketones were also evaluated but, despite our efforts, racemic mixtures were obtained in all the cases.

Regarding the reaction mechanism, and based on previous computational and experimental studies carried out by our research group employing identical catalysts for the asymmetric conjugate addition of 1,3-dicarbonyl compounds onto nitroalkenes,[9a] we propose the catalytic cycle depicted in Scheme [2] in which benzimidazole II can act as a bifunctional organocatalyst. Thus, II can act initially as a base, forming the corresponding 1,3-dicarbonyl compound enolate, that can coordinate through hydrogen-bonding to the catalyst, as depicted in intermediate A. Then, the protonated dimethylamino moiety can activate the diazocarboxylate and hence facilitate the enantioselective attack of the enolate (intermediate B), releasing the corresponding amination product and regenerating the organocatalyst II.

Zoom Image
Scheme 2 Proposed catalytic cycle

It is worthy of note that the S-configured amination product seems to be obtained when (R,R)-II is employed. This assumption was taken from a specific rotation comparison between product 3aa and the values reported in literature.[11]

In conclusion, we have demonstrated that chiral trans-cyclohexanediamine benzimidazole derivative II is a suitable and effective organocatalyst for the asymmetric electrophilic amination of cyclic 1,3-dicarbonyl compounds. The corresponding amination products are obtained in the majority of the cases with high yields and moderate to high enantioselectivities using just 1 mol% of catalyst loading. In addition, a bifunctional role of the catalyst is assumed due to a presumably dual hydrogen-bond activation of both the 1,3-dicarbonyl compound and the diazocarboxylate.


#

Acknowledgment

Financial support from the University of Alicante (VIGROB-173, GRE12-03, UAUSTI13-01, UAUSTI13-02) is gratefully acknowledged.

Supporting Information

  • References

    • 1a Broggini G, Borsini E, Piarulli U In Science of Synthesis, Cross-Coupling and Heck-Type Reactions. Vol. 3. Molander GA, Wolfe JP, Larhed M. Thieme; Stuttgart: 2013: 521-583
    • 1b Christoffers J, Mann A. Angew. Chem. Int. Ed. 2001; 40: 4591
    • 1c Corey EJ, Guzman-Perez A. Angew. Chem. Int. Ed. 1998; 37: 389
  • 2 Ciganek E In Organic Reactions. Vol. 72. Denmark SE. Wiley; New Jersey: 2008: 1-366

    • For selected reviews about electrophilic amination, see:
    • 3a Erdik E. Tetrahedron 2004; 60: 8747
    • 3b Greck C, Drouillat B, Thomassigny C. Eur. J. Org. 2004; 1377
    • 3c Guillena G, Ramón DJ. Tetrahedron: Asymmetry 2006; 17: 1465
    • 3d Vilaivan T, Bhanthumnavin W. Molecules 2010; 15: 917
    • 3e Vallribera A, Sebastian RM, Shafir A. Curr. Org. Chem. 2011; 15: 1539
    • 3f Russo A, De Fusco C, Lattanzi A. RSC. Adv. 2012; 2: 385
    • 3g Chauhan P, Chimni SS. Tetrahedron: Asymmetry 2013; 24: 343
  • 4 Marigo M, Juhl K, Jørgensen KA. Angew. Chem. Int. Ed. 2003; 42: 1367

    • For selected examples of metal-catalyzed electrophilic amination of 1,3-dicarbonyl compounds with diazocarboxylates, see:
    • 5a Foltz C, Stecker B, Marconi G, Bellemin-Laponaz S, Wadepohl H, Gade LH. Chem. Commun. 2005; 5115
    • 5b Kang YK, Kim DY. Tetrahedron Lett. 2006; 47: 4565
    • 5c Mashiko T, Kumagai N, Shibasaki M. J. Am. Chem. Soc. 2009; 131: 14990
    • 5d Mang JY, Kwon DG, Kim DY. Bull. Korean Chem. Soc. 2009; 30: 249
    • 5e Ghosh S, Nandakumar MV, Krautscheid H, Schneider C. Tetrahedron Lett. 2010; 51: 1860
    • 5f Torres M, Maisse-François A, Bellemin-Laponaz S. ChemCatChem 2013; 5: 3078

      For selected examples of electrophilic amination of 1,3-dicarbonyl compounds with diazocarboxylates using chiral amines as organocatalysts, see:
    • 6a Saaby S, Bella M, Jørgensen KA. J. Am. Chem. Soc. 2004; 126: 8120
    • 6b Pihko PM, Pohjakallio A. Synlett 2004; 2115
    • 6c Liu X, Li H, Deng L. Org. Lett. 2005; 7: 169
    • 6d Santacruz L, Niembro S, Santillana A, Shafir A, Vallribera A. New J. Chem. 2014; 38: 636

      For selected examples of electrophilic amination of 1,3-dicarbonyl compounds with diazocarboxylates using phase-transfer catalysts, see:
    • 7a He R, Wang X, Hashimoto T, Maruoka K. Angew. Chem. Int. Ed. 2008; 47: 9466
    • 7b Lan Q, Wang X, He R, Ding C, Maruoka K. Tetrahedron Lett. 2009; 50: 3280
    • 7c He R, Maruoka K. Synthesis 2009; 2289

      For selected examples of electrophilic amination of 1,3-dicarbonyl compounds with diazocarboxylates using hydrogen-bond catalysis, see:
    • 8a Xu X, Yabuta T, Yuan P, Takemoto Y. Synlett 2006; 137
    • 8b Terada M, Nakano M, Ube H. J. Am. Chem. Soc. 2006; 128: 16044
    • 8c Jung SH, Kim DY. Tetrahedron Lett. 2008; 49: 5527
    • 8d Konishi H, Lam TY, Malerich JP, Rawal VH. Org. Lett. 2010; 12: 2028
    • 8e Han X, Zhong F, Lu Y. Adv. Synth. Catal. 2010; 352: 2778
    • 8f Zhang Z.-H, Dong X.-Q, Tao H.-Y, Wang C.-J. ARKIVOC 2011; (ii): 137
    • 8g Inokuma T, Furukawa M, Uno T, Suzuki Y, Yoshida K, Yano Y, Matsuzaki K, Takemoto Y. Chem. Eur. J. 2011; 17: 10470
    • 8h Murai K, Fukushima S, Nakamura A, Shimura M, Fujioka H. Tetrahedron 2011; 67: 4862

      For recent publications, see:
    • 9a Almasi D, Alonso DA, Gómez-Bengoa E, Nájera C. J. Org. Chem. 2009; 74: 6163
    • 9b Gómez-Torres E, Alonso DA, Gómez-Bengoa E, Nájera C. Org. Lett. 2011; 13: 6106
    • 9c Gómez-Torres E, Alonso DA, Gómez-Bengoa E, Nájera C. Eur. J. Org. Chem. 2013; 1434
    • 9d Trillo P, Baeza A, Nájera C. Synthesis 2014; in press; DOI: 10.1055/s-0034-1378618
  • 10 General Procedure for the Asymmetric Amination of Cyclic 1,3-Dicarbonyl Compounds: In a tube, open to the atmosphere, in a thermostated bath (25 °C) the requisite 1,3-dicarbonyl compound (0.2 mmol) was added to a solution of organocatalyst II (0.002 mmol, 1 mol%) in Et2O (1 mL). After 5 min, di-tert-butylazodicarboxylate (2a; 0.21 mmol, 1.05 equiv) was added in one portion and the reaction was then allowed to react for 10 h. After this time, H2O (5 mL) and EtOAc were added, and then the aqueous layer was re-extracted with EtOAc (2 × 5 mL). The combined organic phases were dried (MgSO4), filtered and the solvent was evaporated under reduced pressure. Finally, the crude product was purified by flash chromatography using hexanes–EtOAc mixtures as eluent.Physical and spectroscopic data given below are for compound 3da and may be taken as representative. For further details; see Supporting Information. Di-tert-butyl 1-[2-(Ethoxycarbonyl)-1-oxo-1,2,3,4-tetrahydronaphthalen-2-yl]hydrazine-1,2-dicarboxylate (3da) 8b Slightly yellow viscous oil (88 mg, 98% yield, 91% ee); [α]D 28 +23.3 (c = 2.0, CHCl3). 1H NMR (300 MHz): δ = 1.33 (br m, 21 H), 2.67 (m, 1 H), 2.95 (m, 2 H), 3.44 (br m, 1 H), 4.31 (q, J = 7.0 Hz, 2 H), 6.23 (m, 1 H), 7.25 (m, 2 H), 7.46 (m, 1 H), 7.95 (dd, J = 28.0, 7.2 Hz, 1 H). 13C NMR (75 MHz): δ = 14.1, 25.6, 27.7, 28.0, 31.1, 60.3, 61.9, 80.8, 82.7, 126.4, 127.7, 128.5, 131.7, 133.4, 154.4, 155.5, 169.5, 191.0. MS (IE): m/z (%) = 348 (6.5) [M+ – Boc], 292 (47), 219 (100), 175 (86), 158 (30). Chiral HPLC analysis: Chiralcel IA column, hexane–i-PrOH (85:15), flow rate = 1 mL/min, λ = 254 nm, retention times: tR = 8.0, 11.5 min.
  • 11 (S)-3aa [α]D 29 +3.8 (c = 1, CHCl3, 92% ee). Reported value in ref. 5b for R-enantiomer: [α]D 32 –3.47 (c = 1.09, CHCl3, 97% ee). See Supporting Information for further details.

  • References

    • 1a Broggini G, Borsini E, Piarulli U In Science of Synthesis, Cross-Coupling and Heck-Type Reactions. Vol. 3. Molander GA, Wolfe JP, Larhed M. Thieme; Stuttgart: 2013: 521-583
    • 1b Christoffers J, Mann A. Angew. Chem. Int. Ed. 2001; 40: 4591
    • 1c Corey EJ, Guzman-Perez A. Angew. Chem. Int. Ed. 1998; 37: 389
  • 2 Ciganek E In Organic Reactions. Vol. 72. Denmark SE. Wiley; New Jersey: 2008: 1-366

    • For selected reviews about electrophilic amination, see:
    • 3a Erdik E. Tetrahedron 2004; 60: 8747
    • 3b Greck C, Drouillat B, Thomassigny C. Eur. J. Org. 2004; 1377
    • 3c Guillena G, Ramón DJ. Tetrahedron: Asymmetry 2006; 17: 1465
    • 3d Vilaivan T, Bhanthumnavin W. Molecules 2010; 15: 917
    • 3e Vallribera A, Sebastian RM, Shafir A. Curr. Org. Chem. 2011; 15: 1539
    • 3f Russo A, De Fusco C, Lattanzi A. RSC. Adv. 2012; 2: 385
    • 3g Chauhan P, Chimni SS. Tetrahedron: Asymmetry 2013; 24: 343
  • 4 Marigo M, Juhl K, Jørgensen KA. Angew. Chem. Int. Ed. 2003; 42: 1367

    • For selected examples of metal-catalyzed electrophilic amination of 1,3-dicarbonyl compounds with diazocarboxylates, see:
    • 5a Foltz C, Stecker B, Marconi G, Bellemin-Laponaz S, Wadepohl H, Gade LH. Chem. Commun. 2005; 5115
    • 5b Kang YK, Kim DY. Tetrahedron Lett. 2006; 47: 4565
    • 5c Mashiko T, Kumagai N, Shibasaki M. J. Am. Chem. Soc. 2009; 131: 14990
    • 5d Mang JY, Kwon DG, Kim DY. Bull. Korean Chem. Soc. 2009; 30: 249
    • 5e Ghosh S, Nandakumar MV, Krautscheid H, Schneider C. Tetrahedron Lett. 2010; 51: 1860
    • 5f Torres M, Maisse-François A, Bellemin-Laponaz S. ChemCatChem 2013; 5: 3078

      For selected examples of electrophilic amination of 1,3-dicarbonyl compounds with diazocarboxylates using chiral amines as organocatalysts, see:
    • 6a Saaby S, Bella M, Jørgensen KA. J. Am. Chem. Soc. 2004; 126: 8120
    • 6b Pihko PM, Pohjakallio A. Synlett 2004; 2115
    • 6c Liu X, Li H, Deng L. Org. Lett. 2005; 7: 169
    • 6d Santacruz L, Niembro S, Santillana A, Shafir A, Vallribera A. New J. Chem. 2014; 38: 636

      For selected examples of electrophilic amination of 1,3-dicarbonyl compounds with diazocarboxylates using phase-transfer catalysts, see:
    • 7a He R, Wang X, Hashimoto T, Maruoka K. Angew. Chem. Int. Ed. 2008; 47: 9466
    • 7b Lan Q, Wang X, He R, Ding C, Maruoka K. Tetrahedron Lett. 2009; 50: 3280
    • 7c He R, Maruoka K. Synthesis 2009; 2289

      For selected examples of electrophilic amination of 1,3-dicarbonyl compounds with diazocarboxylates using hydrogen-bond catalysis, see:
    • 8a Xu X, Yabuta T, Yuan P, Takemoto Y. Synlett 2006; 137
    • 8b Terada M, Nakano M, Ube H. J. Am. Chem. Soc. 2006; 128: 16044
    • 8c Jung SH, Kim DY. Tetrahedron Lett. 2008; 49: 5527
    • 8d Konishi H, Lam TY, Malerich JP, Rawal VH. Org. Lett. 2010; 12: 2028
    • 8e Han X, Zhong F, Lu Y. Adv. Synth. Catal. 2010; 352: 2778
    • 8f Zhang Z.-H, Dong X.-Q, Tao H.-Y, Wang C.-J. ARKIVOC 2011; (ii): 137
    • 8g Inokuma T, Furukawa M, Uno T, Suzuki Y, Yoshida K, Yano Y, Matsuzaki K, Takemoto Y. Chem. Eur. J. 2011; 17: 10470
    • 8h Murai K, Fukushima S, Nakamura A, Shimura M, Fujioka H. Tetrahedron 2011; 67: 4862

      For recent publications, see:
    • 9a Almasi D, Alonso DA, Gómez-Bengoa E, Nájera C. J. Org. Chem. 2009; 74: 6163
    • 9b Gómez-Torres E, Alonso DA, Gómez-Bengoa E, Nájera C. Org. Lett. 2011; 13: 6106
    • 9c Gómez-Torres E, Alonso DA, Gómez-Bengoa E, Nájera C. Eur. J. Org. Chem. 2013; 1434
    • 9d Trillo P, Baeza A, Nájera C. Synthesis 2014; in press; DOI: 10.1055/s-0034-1378618
  • 10 General Procedure for the Asymmetric Amination of Cyclic 1,3-Dicarbonyl Compounds: In a tube, open to the atmosphere, in a thermostated bath (25 °C) the requisite 1,3-dicarbonyl compound (0.2 mmol) was added to a solution of organocatalyst II (0.002 mmol, 1 mol%) in Et2O (1 mL). After 5 min, di-tert-butylazodicarboxylate (2a; 0.21 mmol, 1.05 equiv) was added in one portion and the reaction was then allowed to react for 10 h. After this time, H2O (5 mL) and EtOAc were added, and then the aqueous layer was re-extracted with EtOAc (2 × 5 mL). The combined organic phases were dried (MgSO4), filtered and the solvent was evaporated under reduced pressure. Finally, the crude product was purified by flash chromatography using hexanes–EtOAc mixtures as eluent.Physical and spectroscopic data given below are for compound 3da and may be taken as representative. For further details; see Supporting Information. Di-tert-butyl 1-[2-(Ethoxycarbonyl)-1-oxo-1,2,3,4-tetrahydronaphthalen-2-yl]hydrazine-1,2-dicarboxylate (3da) 8b Slightly yellow viscous oil (88 mg, 98% yield, 91% ee); [α]D 28 +23.3 (c = 2.0, CHCl3). 1H NMR (300 MHz): δ = 1.33 (br m, 21 H), 2.67 (m, 1 H), 2.95 (m, 2 H), 3.44 (br m, 1 H), 4.31 (q, J = 7.0 Hz, 2 H), 6.23 (m, 1 H), 7.25 (m, 2 H), 7.46 (m, 1 H), 7.95 (dd, J = 28.0, 7.2 Hz, 1 H). 13C NMR (75 MHz): δ = 14.1, 25.6, 27.7, 28.0, 31.1, 60.3, 61.9, 80.8, 82.7, 126.4, 127.7, 128.5, 131.7, 133.4, 154.4, 155.5, 169.5, 191.0. MS (IE): m/z (%) = 348 (6.5) [M+ – Boc], 292 (47), 219 (100), 175 (86), 158 (30). Chiral HPLC analysis: Chiralcel IA column, hexane–i-PrOH (85:15), flow rate = 1 mL/min, λ = 254 nm, retention times: tR = 8.0, 11.5 min.
  • 11 (S)-3aa [α]D 29 +3.8 (c = 1, CHCl3, 92% ee). Reported value in ref. 5b for R-enantiomer: [α]D 32 –3.47 (c = 1.09, CHCl3, 97% ee). See Supporting Information for further details.

Zoom Image
Scheme 1 Study of different diazocarboxylates
Zoom Image
Scheme 2 Proposed catalytic cycle