Synthesis 2014; 46(01): 101-109
DOI: 10.1055/s-0033-1340052
paper
© Georg Thieme Verlag Stuttgart · New York

Iron(II) Chloride–1,1′-Binaphthyl-2,2′-diamine (FeCl2–BINAM) Complex Catalyzed Domino Synthesis of Bisindolylmethanes from Indoles and Primary Alcohols

Sindhura Badigenchala
Department of Chemistry, Indian Institute of Technology Madras, Chennai, Tamil Nadu-600 036, India   Fax: +91(44)22574202   eMail: gsekar@iitm.ac.in
,
Dhandapani Ganapathy
Department of Chemistry, Indian Institute of Technology Madras, Chennai, Tamil Nadu-600 036, India   Fax: +91(44)22574202   eMail: gsekar@iitm.ac.in
,
Ankita Das
Department of Chemistry, Indian Institute of Technology Madras, Chennai, Tamil Nadu-600 036, India   Fax: +91(44)22574202   eMail: gsekar@iitm.ac.in
,
Rahul Singh
Department of Chemistry, Indian Institute of Technology Madras, Chennai, Tamil Nadu-600 036, India   Fax: +91(44)22574202   eMail: gsekar@iitm.ac.in
,
Govindasamy Sekar*
Department of Chemistry, Indian Institute of Technology Madras, Chennai, Tamil Nadu-600 036, India   Fax: +91(44)22574202   eMail: gsekar@iitm.ac.in
› Institutsangaben
Weitere Informationen

Publikationsverlauf

Received: 22. August 2013

Accepted after revision: 01. Oktober 2013

Publikationsdatum:
04. November 2013 (online)

 


Abstract

Biologically important bisindolylmethanes are synthesized in a domino fashion by using an iron(II) chloride–(±)-1,1′-binaphthyl-2,2′-diamine [FeCl2–(±)-BINAM] complex as the catalyst. This method proceeds via oxidation of a primary alcohol into the corresponding aldehyde followed by nucleophilic addition of an indole in the presence of the catalyst. A reaction intermediate is synthesized separately and converted into the bisindolylmethane product under the same reaction conditions as support for the proposed mechanism.


#

1,1-Bisindolylmethanes and their derivatives are known to have a broad spectrum of biological and pharmacological activities.[1] They are active against human breast cancer cells and are found to activate a specific estrogen receptor.[2] These compounds show growth inhibitory activity toward lung cancer cells,[3] inhibit bladder cancer growth[4] and have antimicrobial,[5] antifungal,[6] antibacterial[7] and antitumor activities.[8] In addition, 1,1-bisindolylmethane derivatives are used as human dietary supplements.[9] The oxidized forms of 1,1-bisindolylmethanes have been reported as chromogenic-sensing molecules.[10] As a result of their potential value in pharmaceuticals and materials, the synthesis of this class of compounds has attracted significant interest from synthetic chemists.[11] 1,1-Bisindolylmethanes have been isolated from metabolites of terrestrial and marine origin,[12] and various protocols have been adopted for their synthesis.[13] Most of the common methods involve the addition of indoles to aldehydes or ketones in the presence of a Lewis acid,[14] a Bronsted acid,[15] transition metals,[16] rare earth catalysts[17] or zeolites.[18] However, many of these methods suffer from the disadvantages of using stoichiometric amounts of acids, expensive metal catalysts and easily oxidizable aldehyde precursors. There have been very few reports in the literature on the synthesis of 1,1-bisindolylmethanes from alcohols.[19] Yokoyama et al. reported the synthesis of 1,1-bisindolylmethanes from benzyl alcohol using a palladium catalyst.[20] Although, this method worked well, the protocol utilized costly palladium as the catalyst and was limited to benzylic alcohols as substrates. Hence, there is a need for an efficient, economic and ecofriendly catalyst for the synthesis of 1,1-bisindolylmethanes starting from primary alcohols.

Iron is an attractive alternative catalyst because of its abundance, low price and environmentally benign character.[21] Unlike other metals, iron is involved as a key element in various biological systems, particularly in oxidations. Due to its ability to undergo facile changes in oxidation state and because of its distinct Lewis acid character, iron catalysts enable a broad range of synthetic transformations such as oxidation, cross-coupling, alkylation and addition reactions.[22] In continuation of our research on environmentally friendly iron-catalyzed reactions,[23] herein, we report an efficient iron(II) chloride­–(±)-1,1′-binaphthyl-2,2′-diamine [FeCl2–(±)-BINAM­] complex catalyzed synthesis of 1,1-bisindolylmethanes from primary alcohols and indoles in a domino fashion.[24]

In our preliminary studies, the synthesis was carried out starting from ethanol via a domino alcohol oxidation in the presence of the FeCl2–BINAM complex as the catalyst and dicumyl peroxide (DCP), followed by condensation of the resulting aldehyde with indole (1) in ethanol, at 120 °C. To our surprise, the bisindolyl product was formed in 68% isolated yield after eight hours (Scheme [1]). It is noteworthy that the reaction did not proceed without the iron catalyst.

Zoom Image
Scheme 1 Synthesis of bisindolylmethane 2 from indole
Zoom Image
Figure 1 Ligand screening for the domino synthesis of 2

In order to improve the reaction efficiency, several BINAM­-derived and other ligands were screened, but none of them provided a better yield compared to BINAM (L1) (Figure [1]). When the reaction was carried out with iron(II) chloride, but without a ligand, only a 27% yield of the product was obtained.

To optimize the conditions in terms of the yield, we screened several other metal salts in combination with BINAM­ (L1) as the ligand in this domino reaction (Table 1). Although copper, cobalt and zinc salts catalyzed the reaction, none of them provided better yields than iron(II) chloride. It was found that when a higher oxidation state iron catalyst (Fe3+) was used, no product formation was observed. The best result was obtained with iron(II) chloride (5 mol%), which gave bisindolylmethane 2 in 75% yield after six hours. The results are summarized in Table 1.

Table 1 Screening of Metal Salts for the Synthesis of 2

Entry

Metal salt

Time (h)

Yield (%)a

1

FeCl2

6

68

2

Fe(OAc)2

16

40

3

FeSO4

16

15

4

FeCl3

24

0

5

FeBr3

20

0

6

Fe(ClO4)2

15

0

7

CuCl2

12

14

8

Cu(OAc)2

12

42

9

Cu(OTf)2

12

45

10

CuI

24

25

11

Co(OAc)2

12

20

12

Ni(OAc)2

24

20

13

Zn(OAc)2

16

38

14

FeCl2

8

50b

15

FeCl2

6

75 c

16

FeCl2

8

69d

17

FeCl2

12

70e

a Yield of isolated product.

b FeCl2 (2.5 mol%).

c FeCl2 (5 mol%) and L1 (10 mol%).

d FeCl2 (10 mol%) and L1 (10 mol%).

e FeCl2 (10 mol%) and L1 (20 mol%).

Next, different types of oxidizing agents were examined. Oxidants including hydrogen peroxide (H2O2), tert-butyl hydroperoxide (t-BuOOH) and benzoyl peroxide were less effective for the formation of product 2 when compared with dicumyl peroxide (DCP). The reaction was also attempted with 2,2,6,6-tetramethylpiperidine 1-oxyl (TEMPO) and molecular oxygen as the oxidant, however, there was no product formation. The number of equivalents of dicumyl peroxide used was important with 3.5 equivalents giving the best result.

Since the temperature plays a major role in catalyst efficiency, the reaction was examined at different temperatures. When the temperature was lowered to 80 °C, the yield of product 2 (76%) remained almost the same (Table 2, entry 9). However, when the temperature was reduced to 60 °C, the yield decreased to 62% (Table 2, entry 10).

Table 2 Screening of Oxidants for the Synthesis of 2

Entry

Oxidant

Temp (°C)

Time (h)

Yield (%)a

1

DCP

120

6

75

2

H2O2

120

16

0

3

t-BuOOH

120

16

0

4

(PhCO2)2

120

18

trace

5

TEMPO

120

40

0b

6

DCP

120

48

0c

7

DCP

120

18

55d

8

DCP

100

10

72

9

DCP

80

6

76

10

DCP

60

12

62

11

DCP

r.t.

24

0

a Yield of isolated product.

b O2 was used as a co-oxidant.

c TEMPO (1.0 equiv) was added.

d DCP (2.0 equiv).

From the optimization studies the best catalytic system was found to be: iron(II) chloride (5 mol%), 1,1′-binaphthyl-2,2′-diamine (L1) (10 mol%), dicumyl peroxide (3.5 equiv), 80 °C. The substrate scope of this methodology was evaluated using the optimized reaction conditions and the results are summarized in Table 3. Notably, indole reacted with ethanol to give the important natural product, vibrindole A (2) (Table 3, entry 1). For substituted indoles, it was found that the presence of an electron-releasing group on the nitrogen atom resulted in a good yield of the corresponding product (Table 3, entry 2). However, electron-withdrawing groups on the indole nitrogen atom, such as tosyl, completely inhibited the reaction (Table 3, entry 8), whilst an electron-withdrawing group on the benzene ring reduced the yield (Table 3, entry 5).

Increasing the length of the aliphatic chain of the alcohol led to reduced yields; when the alkyl chain was more than three carbon atoms long, the reaction did not take place. In the case of benzyl alcohol the reaction required a longer time than aliphatic alcohols (Table 3, entries 11–14).

Table 3 Iron-Catalyzed Domino Synthesis of Bisindolylmethanes from Indoles and Alcohols

Entry

Indole

Alcohol

Product

Time (h)

Yield (%)

1

2

6

76a

2

3

6

80

3

4

24

62

4

5

12

70

5

6

36

63

6

7

15

63

7

8

48

54

8

9

24

0

9

10

10

85

10

11

36

62

11

12

36

60b,c

12

13

48

62b,c

13

14

48

58b,c

14

15

36

59b,c

a Yield of isolated product.

b tert-Butyl hydroperoxide was used as the oxidant.

c tert-Butyl alcohol (2 mL) was used as the solvent.

The reaction was completely suppressed by adding one equivalent of TEMPO (with respect to FeCl2), a radical trapping agent, to the reaction mixture. These results indicate that a radical intermediate is most likely involved in the initial steps of the domino transformation. This explains the observed fact that aliphatic alcohols are more reactive than benzyl alcohol, since benzylic radicals are stabilized by resonance effects. When a secondary alcohol was subjected to the optimized conditions there was no reaction at all. These observations prove that the first step involves oxidation of the alcohol into an aldehyde. More importantly, there was no product formation at all when the strongly electron-deficient indole, N-tosylindole was used. The second step might involve nucleophilic attack of indole, which is directed by the lone pair of electrons on the nitrogen of the indole.

Zoom Image
Scheme 2 A plausible mechanism for the domino synthesis of bisindolylmethanes

Based on these observations, a plausible mechanism for the domino synthesis of bisindolylmethanes, using 2 as an example, is suggested (Scheme [2]). Two catalytic cycles are proposed in the mechanism. In the first cycle, the primary alcohol 16 is oxidized by iron(II) chloride and dicumyl peroxide to give the corresponding aldehyde 19 through the radical intermediates 17 and 18. In the second cycle, nucleophilic addition of indole (1) to the iron(II) chloride activated aldehyde 19 (which is formed in first the cycle) affords secondary alcohol 20a (see Scheme [3]). The secondary alcohol is further activated by iron(II) chloride and undergoes a second addition of indole to give the bisindolylmethane 2. We anticipated that the reaction proceeded through the secondary alcohol 20a of the intermediate 20, and we thus carried out a control experiment to understand the mechanism. The intermediate 20a was synthesized by the reduction of 3-acetylindole using sodium borohydride,[25] and then subjected to our standard conditions for the preparation of bisindolylmethanes. As expected, the reaction took place smoothly and afforded an 80% isolated yield of compound 2 (Scheme [3]). The instability of the intermediate 20a explains the adverse effect of high temperature on this reaction.[26]

Zoom Image
Scheme 3 Synthesis of bisindolylmethane 2 from indole and the secondary alcohol 20a

In summary, an efficient, cost-effective, and environmentally friendly iron-catalyzed domino synthesis of bisindolylmethanes and their derivatives from indoles and primary alcohols has been reported. A plausible mechanism has been proposed for this domino process. In support of the mechanism, one of the postulated reaction intermediates was independently synthesized and converted into the corresponding bisindolylmethane under the same reaction conditions.

All reactions were carried out in screw-cap pressure tubes under N2. All the solvents used for the reactions were obtained from Merck, India and were dried according to standard procedures. EtOH was purchased from Changshu Yangyuan Chemical, China, and dried over 4 Å molecular sieves. Reactions were monitored by thin-layer chromatography (TLC) using Merck silica gel 60 F254 precoated plates (0.25 mm), and samples were made visual by UV fluorescence. Silica gel (particle size: 100–200 mesh) was purchased from SRL India and was used for column chromatography using appropriate mixtures of hexanes–EtOAc as the eluent. FeCl2 was obtained from Sigma-Aldrich Company. Other chemicals were purchased: indole from Spectrochem Pvt. Ltd., Mumbai, India (AR), dicumyl peroxide from Acros Organics, and 1,1′-binaphthyl-2,2′-diamine (BINAM) ligand L1 was purchased from GERCHEM chemicals, Hyderabad, India. Reaction temperatures were controlled using a Varivolt temperature modulator. Melting points were obtained using a Toshniwal melting point apparatus and are uncorrected. FTIR spectra were recorded on a Nicolet 6700 spectrometer and absorptions are reported in wavenumbers (cm–1). 1H and 13C NMR spectra were recorded on Bruker 400 or 500 MHz instruments. 1H NMR spectra are reported relative to Me4Si (δ 0.0) or residual CHCl3 (δ 7.26). 13C NMR are reported relative to CDCl3 (δ 77.16). High-resolution mass spectra (HRMS) were recorded on Q-Tof Micro mass spectrometer.


#

3,3′-(Ethane-1,1-diyl)bis(1H-indole) (2);[27] Typical Procedure

An oven-dried, screw-cap pressure tube containing a magnetic stir bar was charged with FeCl2 (3.2 mg, 0.025 mmol), 1,1′-binaphthyl-2,2′-diamine (BINAM) (14.2 mg, 0.05 mmol), dicumyl peroxide (DCP) (473.2 mg, 1.75 mmol) and indole (1) (58.6 mg, 0.5 mmol). The pressure tube was evacuated and back-filled with N2. Anhydrous EtOH (2 mL) was added and the mixture was stirred at 80 °C for 6 h. After the complete disappearance of indole (the progress of the reaction was monitored by TLC), the mixture was allowed to cool to r.t. and the EtOH was evaporated under reduced pressure using a rotary evaporator. Next, H2O (15 mL) was added, and the product was extracted with EtOAc (3 × 10 mL) and dried over anhydrous Na2SO4. The solvent was evaporated and the residue purified by column chromatography on silica gel (EtOAc–hexane, 12:88) to afford pure product 2.

Yield: 49.8 mg (76%); light yellow solid; mp 148 °C (Lit.[27] 158–160 °C); Rf = 0.39 (20% EtOAc in hexane).

IR (neat): 3413, 3053, 2969, 2871, 1456, 1417, 1340, 1221, 1092, 1012, 746, 585 cm–1.

1H NMR (400 MHz, CDCl3): δ = 1.83 (d, J = 7.2 Hz, 3 H), 4.70 (q, J = 7.2 Hz, 1 H), 6.89 (d, J = 2.4 Hz, 2 H), 7.08 (td, J = 8, 0.8 Hz, 2 H), 7.20 (td, J = 8.0, 0.8 Hz, 2 H), 7.34 (d, J = 8 Hz, 2 H), 7.61 (d, J = 8.4 Hz, 2 H), 7.80 (br s, 2 H).

13C NMR (100 MHz, CDCl3): δ = 21.9, 28.3, 111.2, 119.1, 119.9, 121.3, 121.8, 121.9, 127.0, 136.8.

HRMS (ESI, +): m/z [M + Na]+ calcd for C18H16N2Na: 283.1211; found: 283.1220.


#

3,3′-(Ethane-1,1-diyl)bis(1-methyl-1H-indole) (3)

Yield: 58.4 mg (80%); colorless oil; Rf = 0.77 (20% EtOAc in hexane­).

IR (neat): 2925, 1612, 1469, 739 cm–1.

1H NMR (500 MHz, CDCl3): δ = 1.71 (d, J = 7.0 Hz, 3 H), 3.63 (s, 6 H), 4.59 (q, J = 7.1 Hz, 1 H), 6.71 (s, 2 H), 6.96 (t, J = 7.0 Hz, 2 H), 7.12 (td, J = 7.0, 1.1 Hz, 2 H), 7.20 (t, J = 8.2 Hz, 2 H), 7.51 (d, J = 8 Hz, 2 H).

13C NMR (125 MHz, CDCl3): δ = 22.5, 28.3, 32.9, 109.4, 118.7, 120.1, 120.6, 121.6, 126.3, 127.6, 137.6.

HRMS (ESI, +): m/z [M + H]+ calcd for C20H21N2: 289.1705; found: 289.1714.


#

3,3′-(Ethane-1,1-diyl)bis(1-ethyl-1H-indole) (4)

Yield: 49.7 mg (62%); pale brown solid; mp 101 °C; Rf = 0.54 (5% EtOAc in hexane).

IR (KBr): 2929, 2869, 1607, 1547, 1462, 1394, 1334, 931, 819 cm–1.

1H NMR (500 MHz, CDCl3): δ = 1.33 (t, J = 7 Hz, 6 H), 1.72 (d, J = 7 Hz, 3 H), 4.02 (dq, J = 7.0, 1.5 Hz, 4 H), 4.59 (q, J = 7.5 Hz, 1 H), 6.77 (s, 2 H), 6.95 (t, J = 7 Hz, 2 H), 7.10 (t, J = 7.5 Hz, 2 H), 7.23 (d, J = 8.5 Hz, 2 H), 7.50 (d, J = 8 Hz, 2 H).

13C NMR (125 MHz, CDCl3): δ = 15.5, 22.2, 28.2, 40.8, 109.2, 118.3, 120.0, 121.1, 124.3, 125.6, 127.9, 136.3.

HRMS (ESI, +): m/z [M + K]+ calcd for C22H24N2K: 355.1577; found: 355.1589.


#

3,3′-(Propane-1,1-diyl)bis(5-methoxy-1H-indole) (5)

Yield: 58.2 mg (70%); brown solid; mp 121 °C; Rf = 0.41 (30% EtOAc in hexane).

IR (KBr): 3409, 2924, 2853, 1617, 1452, 1309, 1089, 878, 800, 760 cm–1.

1H NMR (400 MHz, CDCl3): δ = 1.02 (t, J = 7.2 Hz, 3 H), 2.23 (quin, J = 7.2 Hz, 2 H), 3.77 (s, 6 H), 4.27 (t, J = 7.2 Hz, 1 H), 6.81 (dd, J = 8, 2 Hz, 2 H), 6.99 (d, J = 1.6 Hz, 2 H), 7.03 (d, J = 2 Hz, 2 H), 7.22 (d, J = 8 Hz, 2 H), 7.81 (br s, 2 H).

13C NMR (100 MHz, CDCl3): δ = 13.3, 28.2, 55.7, 56.0, 102.1, 111.7, 118.8, 122.4, 128.3, 128.4, 144.6, 158.0.


#

3,3′-(Ethane-1,1-diyl)bis(5-bromo-1-methyl-1H-indole) (6)[28]

Yield: 65.8 mg (63%); brown solid; mp 112 °C; Rf = 0.41 (10% EtOAc in hexane).

IR (neat): 3114, 2968, 2821, 1611, 1535, 1420, 1366, 907, 866, 790 cm–1.

1H NMR (500 MHz, CDCl3): δ = 1.66 (d, J = 7.0 Hz, 3 H), 3.62 (s, 6 H), 4.44 (q, J = 7.0 Hz, 1 H), 6.69 (s, 2 H), 7.07 (d, J = 8.5 Hz, 2 H), 7.19 (dd, J = 8.5, 2.0 Hz, 2 H), 7.57 (d, J = 2 Hz, 2 H).

13C NMR (125 MHz, CDCl3): δ = 22.1, 28.0, 32.9, 110.8, 112.2, 119.6, 122.3, 124.4, 127.3, 128.9, 136.2.


#

3,3′-(Ethane-1,1-diyl)bis(5-methoxy-1H-indole) (7)[28]

Yield: 50.6 mg (63%); brown sticky solid; Rf = 0.43 (30% EtOAc­ in hexane).

IR (KBr): 3412, 2954, 2830, 1619, 1479, 1361, 1212, 805, 739 cm–1.

1H NMR (400 MHz, CDCl3): δ = 1.79 (d, J = 7.2 Hz, 3 H), 3.77 (s, 6 H), 4.57 (q, J = 6.8 Hz, 1 H), 6.83 (d, J = 9.2 Hz, 2 H), 6.92 (s, 2 H), 7.01 (s, 2 H), 7.23 (s, 2 H), 7.81 (s, 2 H).

13C NMR (100 MHz, CDCl3): δ = 21.6, 28.2, 56.0, 101.9, 111.7, 111.8, 121.3, 122.2, 127.4, 132.0, 153.6.


#

3,3′-(Ethane-1,1-diyl)bis[1-(4-methoxyphenyl)-1H-indole] (8)

Yield: 63.7 mg (54%); brown solid; mp 89 °C; Rf = 0.46 (10% EtOAc­ in hexane).

IR (KBr): 3049, 2925, 2838, 1607, 1457, 1370, 1246, 834, 743 cm–1.

1H NMR (500 MHz, CDCl3): δ = 1.80 (d, J = 7.5 Hz, 3 H), 3.77 (s, 6 H), 4.69 (q, J = 7 Hz, 1 H), 6.91 (d, J = 8.5 Hz, 4 H), 7.02 (m, 4 H), 7.11 (t, J = 7 Hz, 2 H), 7.29 (d, J = 9 Hz, 4 H), 7.37 (d, J = 8.5 Hz, 2 H), 7.59 (d, J = 8 Hz, 2 H).

13C NMR (125 MHz, CDCl3): δ = 22.0, 28.2, 55.7, 110.5, 114.7, 119.5, 120.0, 121.9, 122.2, 125.6, 125.9, 128.1, 133.2, 136.9, 158.0.

HRMS (ESI, +): m/z [M + Na]+ calcd for C32H28N2O2Na: 495.2048; found: 495.2068.


#

3,3′-(Propane-1,1-diyl)bis(1H-indole) (10)[29]

Yield: 58.2 mg (85%); colorless oil; Rf = 0.45 (20% EtOAc in hexane­).

1H NMR (500 MHz, CDCl3): δ = 0.93 (t, J = 7.4 Hz, 3 H), 2.16 (quin, J = 7.4 Hz, 2 H), 4.29 (t, J = 7.4 Hz, 1 H), 6.87 (d, J = 2.2 Hz, 2 H), 6.95 (t, J = 7.1 Hz, 2 H), 7.06 (t, J = 7.6 Hz, 2 H), 7.22 (d, J = 8.1 Hz, 2 H), 7.51 (d, J = 8 Hz, 2 H), 7.75 (s, 2 H).

13C NMR (125 MHz, CDCl3): δ = 13.2, 28.8, 36.0, 111.1, 119.1, 119.8, 120.4, 121.5, 121.8, 127.3, 136.7.


#

3,3′-(Propane-1,1-diyl)bis(1-methyl-1H-indole) (11)

Yield: 46.5 mg (62%); dark red solid; Rf = 0.56 (10% EtOAc in hexane­).

IR (KBr): 2923, 1607, 1465, 1370, 1086, 747 cm–1.

1H NMR (500 MHz, CDCl3): δ = 0.93 (t, J = 7.5 Hz, 3 H), 2.14 (q, J = 7.5 Hz, 2 H), 3.64 (s, 6 H), 4.29 (t, J = 7.0 Hz, 1 H), 6.77 (s, 2 H), 6.94–6.97 (m, 2 H), 7.09–7.12 (m, 2 H), 7.18 (s, 2 H), 7.53 (d, J = 8 Hz, 2 H).

13C NMR (125 MHz, CDCl3): δ = 13.3, 29.4, 32.8, 35.9, 109.2, 118.5, 119.1, 119.9, 121.3, 126.4, 127.7, 137.4.

HRMS (ESI, +): m/z [M + Na]+ calcd for C21H22N2Na: 325.1681; found: 325.1696.


#

3,3′-(Phenylmethane-1,1-diyl)bis(1H-indole) (12)[30]

Yield: 48.3 mg (60%); pink solid; mp 139 °C (Lit.[30] 141–142 °C); Rf = 0.49 (20% EtOAc in hexane).

IR (KBr): 3409, 2924, 2853, 1605, 1455, 1198, 744 cm–1.

1H NMR (400 MHz, CDCl3): δ = 5.80 (s, 1 H), 6.55 (d, J = 1.6 Hz, 2 H), 6.92 (t, J = 7.2 Hz, 2 H), 7.06–7.14 (m, 3 H), 7.19 (t, J = 7.2 Hz, 2 H), 7.25–7.27 (m, 4 H), 7.31 (d, J = 8 Hz, 2 H), 7.79 (s, 2 H).

13C NMR (100 MHz, CDCl3): δ = 40.2, 111.0, 119.2, 119.9, 121.9, 123.6, 126.1, 127.1, 128.2, 128.7, 133.7, 136.7, 144.0.


#

3,3′-(Phenylmethane-1,1-diyl)bis(1-methyl-1H-indole) (13)[27]

Yield: 54.2 mg (59%); dark red solid; mp 163 °C; Rf = 0.43 (10% EtOAc in hexane).

IR (KBr): 3020, 2927, 1607, 1547, 1472, 1366, 744, 700 cm–1.

1H NMR (400 MHz, CDCl3): δ = 3.66 (s, 6 H), 5.87 (s, 1 H), 6.52 (d, J = 0.8 Hz, 2 H), 6.96–7.00 (m, 2 H), 7.16–7.21 (m, 3 H), 7.24–7.29 (m, 4 H), 7.33–7.35 (m, 2 H), 7.38 (d, J = 7.6 Hz, 2 H).

13C NMR (100 MHz, CDCl3): δ = 32.8, 40.2, 109.2, 118.4, 118.8, 120.2, 121.5, 126.1, 127.6, 128.3, 128.4, 128.8, 137.5, 144.6.

HRMS (ESI, +): m/z [M + Na]+ calcd for C25H22N2Na: 373.1681; found: 373.1696.


#

3,3′-(Phenylmethane-1,1-diyl)bis(5-methoxy-1H-indole) (14)[28]

Yield: 55.4 mg (58%); red solid; mp 198 °C; Rf = 0.56 (30% EtOAc in hexane).

IR (KBr): 3004, 2934, 1619, 1586, 1484, 1448, 1208, 1028, 799 cm–1.

1H NMR (500 MHz, CDCl3): δ = 3.61 (s, 6 H), 5.69 (s, 1 H), 7.11–7.22 (m, 7 H), 7.27 (d, J = 7.2 Hz, 2 H), 7.39 (t, J = 7.6 Hz, 2 H), 7.76 (s, 2 H), 8.04 (d, J = 7.6 Hz, 2 H).

13C NMR (125 MHz, CDCl3): δ = 40.4, 56.0, 102.2, 111.8, 112.0, 119.5, 124.6, 126.2, 128.3, 128.6, 128.9, 130.3, 132.0, 133.8.

HRMS (ESI, +): m/z [M + Na]+ calcd for C25H22N2O2Na: 405.1579; found: 405.1566.


#

3,3′-(Phenylmethane-1,1-diyl)bis(5-bromo-1H-indole) (15)[31]

Yield: 59.1 mg (59%); dark red solid; mp 223 °C; Rf = 0.55 (30% EtOAc in hexane).

IR (KBr): 3416, 3068, 2927, 2859, 1592, 1558, 1449, 976, 775 cm–1.

1H NMR (500 MHz, CDCl3): δ = 5.75 (s, 1 H), 6.65 (d, J = 1.5 Hz, 2 H), 7.23–7.24 (m, 4 H), 7.26 (s, 2 H), 7.29–7.30 (m, 4 H), 7.47 (d, J = 0.5 Hz, 2 H), 7.99 (s, 2 H).

13C NMR (125 MHz, CDCl3): δ = 40.0, 112.7, 112.8, 119.2, 122.4, 124.9, 125.1, 126.7, 128.6, 128.7, 128.8, 135.5, 143.1.

HRMS (ESI, +): m/z [M + H]+ calcd for C23H17N2Br2: 478.9758; found: 478.9743.


#
#

Acknowledgment

We thank CSIR [Project No. 01/(2378)/10/EMR-II], DST New Delhi­ for financial support. B.S. thanks UGC, New Delhi for a JRF, D.G. thanks CSIR New Delhi for an SRF, and A.D. thanks IIT-M for infrastructure.

Supporting Information

  • References

    • 1a Mahboobi S, Teller S, Pongratz H, Hufsky H, Sellmer A, Botzki A, Uecker A, Beckers T, Baasner S, Schaechtele C, Ueberall F, Kassack MU, Dove S, Boehmer FD. J. Med. Chem. 2002; 45: 1002
    • 1b Samsoniya SA, Lomtatidze ZS, Ovsyannikova NN, Suvorov NN. Khim. Farm. Zh. 1987; 21: 827
    • 1c Shiri M, Zolfigol MA, Kruger HG, Tanbakouchian Z. Chem. Rev. 2010; 110: 2250
    • 2a Safe S, Papineni S, Chintharlapalli S. Cancer Lett. 2008; 269: 326
    • 2b Vanderlaag K, Su Y, Frankel AE, Burghardt RC, Barhoumi R, Chadalapaka G, Jutooru I, Safe S. BMC Cancer 2010; 10: 669
  • 3 Ichite N, Chougule MB, Jackson T, Fulzele SV, Safe S, Singh M. Clin. Cancer Res. 2009; 15: 543
  • 4 Inamoto T, Papineni S, Chintharlapalli S, Cho SD, Safe S, Kamat AM. Mol. Cancer Ther. 2008; 7: 3825
  • 5 Sivaprasad G, Perumal PT, Prabavathy VR, Mathivanan N. Bioorg. Med. Chem. Lett. 2006; 16: 6302
  • 6 Kamal A, Khan MN. A, Srinivasa RK, Srikanth YV. V, Kaleem AS, Pranay KK, Murthy US. N. J. Enzyme Inhib. Med. Chem. 2009; 24: 559
  • 7 Bell R, Carmeli S, Sar N. J. Nat. Prod. 1994; 57: 1587
  • 8 Chen I, McDougal A, Wang F, Safe S. Carcinogenesis 1998; 19: 1631
    • 9a Bonnesen C, Eggleston IM, Hayes JD. Cancer Res. 2001; 61: 6120
    • 9b Carter TH, Liu K, Ralph Jr W, Chen D, Qi M, Fan S, Yuan F, Rosen EM, Auborn KJ. J. Nutr. 2002; 132: 3314
  • 10 He X, Hu S, Liu K, Guo Y, Xu J, Shao S. Org. Lett. 2006; 8: 333
    • 11a Jafarpour M, Rezaeifard A, Golshani T. J. Heterocycl. Chem. 2009; 46: 535
    • 11b Gong H, Xie Z. Youji Huaxue 2012; 32: 1195
    • 12a Osawa T, Namiki M. Tetrahedron Lett. 1983; 24: 4719
    • 12b Porter JK, Bacon CW, Robbins JD, Himmelsbach DS, Higman HC. J. Agric. Food Chem. 1977; 25: 88
  • 13 Deb ML, Bhuyan PJ. Tetrahedron Lett. 2006; 47: 1441
    • 14a Wang YM, Wen Z, Chen XM, Du D.-M, Matsuura T, Meng JB. J. Heterocycl. Chem. 1998; 35: 313
    • 14b Chatterjee A, Manna S, Banerji J, Pascard C, Prange T, Shoolery JN. J. Chem. Soc., Perkin Trans. 1 1980; 553
    • 15a Kokare ND, Sangshetti JN, Shinde DB. Chin. Chem. Lett. 2008; 19: 1186
    • 15b Singh P, Singh D, Samant S. Synth. Commun. 2005; 35: 2133
    • 16a Young PC, Hadfield MS, Arrowsmith L, MacLeod KM, Mudd RJ, Jordan-Hore JA, Lee A.-L. Org. Lett. 2012; 14: 898
    • 16b Yang J, Wang Z, Pan F, Li Y, Bao W. Org. Biomol. Chem. 2010; 8: 2975
    • 16c Xia D, Wang Y, Du Z, Zheng QY, Wang C. Org. Lett. 2012; 14: 588
    • 16d Guo X, Pan S, Liu J, Li Z. J. Org. Chem. 2009; 74: 8848
  • 18 Karthik M, Magesh CJ, Perumal PT, Palanichamy M, Arabindoo B, Murugesan V. Appl. Catal., A 2005; 286: 137
    • 19a Suda K, Takanami T. Chem. Lett. 1994; 10: 1915
    • 19b Khosropour AR, Mohammadpoor-Baltork I, Khodaei MM, Ghanbary P. Z. Naturforsch., B: Chem. Sci. 2007; 62: 537
    • 19c Zanardi A, Corberan R, Mata JA, Peris E. Organometallics 2008; 27: 3570
  • 20 Hikawa H, Yokoyama Y. RSC Adv. 2013; 3: 1061
  • 21 Bolm C, Legros J, Le PJ, Zani L. Chem. Rev. 2004; 104: 6217
    • 22a Gopalaiah K. Chem. Rev. 2013; 113: 3248
    • 22b Chen QY, He YB, Yang ZY. J. Fluorine Chem. 1986; 34: 255
    • 22c Sherry BD, Furstner A. Chem. Commun. 2009; 7116
    • 22d Takeuchi M, Shimakoshi H, Kano K. Organometallics 1994; 13: 1208
    • 22e Zhang L, Peng D, Leng X, Huang Z. Angew. Chem. Int. Ed. 2013; 52: 3676
    • 22f Wei Y, Ding H, Lin S, Liang F. Org. Lett. 2011; 13: 1674
    • 22g Niu T, Huang L, Wu T, Zhang Y. Org. Biomol. Chem. 2011; 9: 273
    • 22h De H J, Abbaspour TK, Maes BU. W. Angew. Chem. Int. Ed. 2012; 51: 2745
    • 22i Enthaler S. ChemCatChem 2011; 3: 1929
    • 22j Liu W, Liu J, Ogawa D, Nishihara Y, Guo X, Li Z. Org. Lett. 2011; 13: 6272
    • 22k Agrawal T, Cook SP. Org. Lett. 2013; 15: 96
    • 22l Kuzmina OM, Steib AK, Flubacher D, Knochel P. Org. Lett. 2012; 14: 4818
    • 22m Singh PP, Aithagani SK, Yadav M, Singh VP, Vishwakarma RA. J. Org. Chem. 2013; 78: 2639
    • 23a Muthupandi P, Alamsetti SK, Sekar G. Chem. Commun. 2009; 3288
    • 23b Muthupandi P, Sekar G. Org. Biomol. Chem. 2012; 10: 5347

      Domino reactions have attracted a great deal of attention because they provide a high degree of molecular complexity from structurally simple molecules. These reactions are typically economic and more environmentally friendly as they avoid the isolation of unstable intermediates and reduce production costs. See:
    • 24a Tietze LF. Chem. Rev. 1996; 96: 115
    • 24b Breinbauer R. Synthesis 2007; 794
    • 24c Mueller TJ. J. Angew. Chem. Int. Ed. 2007; 46: 2977
    • 24d Bogdanowicz-Szwed K, Krasodomska M, Krasodomski W. Wiad. Chem. 1997; 51: 643
    • 24e Pellissier H. Chem. Rev. 2013; 113: 442
    • 24f Parsons PJ, Penkett CS, Shell AJ. Chem. Rev. 1996; 96: 195
    • 24g Tietze LF, Duefert SC, Clerc J, Bischoff M, Maass C, Stalke D. Angew. Chem. Int. Ed. 2013; 52: 3191
  • 25 Secondary alcohol 20a is highly unstable to column chromatographic purification and the sodium borohydride reduction of 3-acetylindole in EtOH at r.t. gave single compound 20a, which was used directly in the control experiment without any further purification.
  • 26 The reaction intermediate might be unstable at 100 °C and may decompose before it reacts. This speculation might explain the low yield at 100 °C. Intermediate 20a might be somewhat stable at 80 °C, the temperature at which it gives the maximum yield. At lower temperature, product formation may be sluggish.
  • 27 Ganguly NC, Mondal P, Barik SK. Green Chem. Lett. Rev. 2012; 5: 73
  • 28 Ramachandiran K, Muralidharan D, Perumal PT. Tetrahedron Lett. 2011; 58: 5
  • 29 Xu HY, Zi Y, Xu XP, Wang SY, Ji SJ. Tetrahedron 2013; 69: 1600
  • 30 Thirupathi P, Kim SS. J. Org. Chem. 2010; 75: 5240
  • 31 Mendes SR, Thurow S, Fortes MP, Penteado F, Lenardão EJ, Alves D, Perin G, Jacob RG. Tetrahedron Lett. 2012; 53: 5402

  • References

    • 1a Mahboobi S, Teller S, Pongratz H, Hufsky H, Sellmer A, Botzki A, Uecker A, Beckers T, Baasner S, Schaechtele C, Ueberall F, Kassack MU, Dove S, Boehmer FD. J. Med. Chem. 2002; 45: 1002
    • 1b Samsoniya SA, Lomtatidze ZS, Ovsyannikova NN, Suvorov NN. Khim. Farm. Zh. 1987; 21: 827
    • 1c Shiri M, Zolfigol MA, Kruger HG, Tanbakouchian Z. Chem. Rev. 2010; 110: 2250
    • 2a Safe S, Papineni S, Chintharlapalli S. Cancer Lett. 2008; 269: 326
    • 2b Vanderlaag K, Su Y, Frankel AE, Burghardt RC, Barhoumi R, Chadalapaka G, Jutooru I, Safe S. BMC Cancer 2010; 10: 669
  • 3 Ichite N, Chougule MB, Jackson T, Fulzele SV, Safe S, Singh M. Clin. Cancer Res. 2009; 15: 543
  • 4 Inamoto T, Papineni S, Chintharlapalli S, Cho SD, Safe S, Kamat AM. Mol. Cancer Ther. 2008; 7: 3825
  • 5 Sivaprasad G, Perumal PT, Prabavathy VR, Mathivanan N. Bioorg. Med. Chem. Lett. 2006; 16: 6302
  • 6 Kamal A, Khan MN. A, Srinivasa RK, Srikanth YV. V, Kaleem AS, Pranay KK, Murthy US. N. J. Enzyme Inhib. Med. Chem. 2009; 24: 559
  • 7 Bell R, Carmeli S, Sar N. J. Nat. Prod. 1994; 57: 1587
  • 8 Chen I, McDougal A, Wang F, Safe S. Carcinogenesis 1998; 19: 1631
    • 9a Bonnesen C, Eggleston IM, Hayes JD. Cancer Res. 2001; 61: 6120
    • 9b Carter TH, Liu K, Ralph Jr W, Chen D, Qi M, Fan S, Yuan F, Rosen EM, Auborn KJ. J. Nutr. 2002; 132: 3314
  • 10 He X, Hu S, Liu K, Guo Y, Xu J, Shao S. Org. Lett. 2006; 8: 333
    • 11a Jafarpour M, Rezaeifard A, Golshani T. J. Heterocycl. Chem. 2009; 46: 535
    • 11b Gong H, Xie Z. Youji Huaxue 2012; 32: 1195
    • 12a Osawa T, Namiki M. Tetrahedron Lett. 1983; 24: 4719
    • 12b Porter JK, Bacon CW, Robbins JD, Himmelsbach DS, Higman HC. J. Agric. Food Chem. 1977; 25: 88
  • 13 Deb ML, Bhuyan PJ. Tetrahedron Lett. 2006; 47: 1441
    • 14a Wang YM, Wen Z, Chen XM, Du D.-M, Matsuura T, Meng JB. J. Heterocycl. Chem. 1998; 35: 313
    • 14b Chatterjee A, Manna S, Banerji J, Pascard C, Prange T, Shoolery JN. J. Chem. Soc., Perkin Trans. 1 1980; 553
    • 15a Kokare ND, Sangshetti JN, Shinde DB. Chin. Chem. Lett. 2008; 19: 1186
    • 15b Singh P, Singh D, Samant S. Synth. Commun. 2005; 35: 2133
    • 16a Young PC, Hadfield MS, Arrowsmith L, MacLeod KM, Mudd RJ, Jordan-Hore JA, Lee A.-L. Org. Lett. 2012; 14: 898
    • 16b Yang J, Wang Z, Pan F, Li Y, Bao W. Org. Biomol. Chem. 2010; 8: 2975
    • 16c Xia D, Wang Y, Du Z, Zheng QY, Wang C. Org. Lett. 2012; 14: 588
    • 16d Guo X, Pan S, Liu J, Li Z. J. Org. Chem. 2009; 74: 8848
  • 18 Karthik M, Magesh CJ, Perumal PT, Palanichamy M, Arabindoo B, Murugesan V. Appl. Catal., A 2005; 286: 137
    • 19a Suda K, Takanami T. Chem. Lett. 1994; 10: 1915
    • 19b Khosropour AR, Mohammadpoor-Baltork I, Khodaei MM, Ghanbary P. Z. Naturforsch., B: Chem. Sci. 2007; 62: 537
    • 19c Zanardi A, Corberan R, Mata JA, Peris E. Organometallics 2008; 27: 3570
  • 20 Hikawa H, Yokoyama Y. RSC Adv. 2013; 3: 1061
  • 21 Bolm C, Legros J, Le PJ, Zani L. Chem. Rev. 2004; 104: 6217
    • 22a Gopalaiah K. Chem. Rev. 2013; 113: 3248
    • 22b Chen QY, He YB, Yang ZY. J. Fluorine Chem. 1986; 34: 255
    • 22c Sherry BD, Furstner A. Chem. Commun. 2009; 7116
    • 22d Takeuchi M, Shimakoshi H, Kano K. Organometallics 1994; 13: 1208
    • 22e Zhang L, Peng D, Leng X, Huang Z. Angew. Chem. Int. Ed. 2013; 52: 3676
    • 22f Wei Y, Ding H, Lin S, Liang F. Org. Lett. 2011; 13: 1674
    • 22g Niu T, Huang L, Wu T, Zhang Y. Org. Biomol. Chem. 2011; 9: 273
    • 22h De H J, Abbaspour TK, Maes BU. W. Angew. Chem. Int. Ed. 2012; 51: 2745
    • 22i Enthaler S. ChemCatChem 2011; 3: 1929
    • 22j Liu W, Liu J, Ogawa D, Nishihara Y, Guo X, Li Z. Org. Lett. 2011; 13: 6272
    • 22k Agrawal T, Cook SP. Org. Lett. 2013; 15: 96
    • 22l Kuzmina OM, Steib AK, Flubacher D, Knochel P. Org. Lett. 2012; 14: 4818
    • 22m Singh PP, Aithagani SK, Yadav M, Singh VP, Vishwakarma RA. J. Org. Chem. 2013; 78: 2639
    • 23a Muthupandi P, Alamsetti SK, Sekar G. Chem. Commun. 2009; 3288
    • 23b Muthupandi P, Sekar G. Org. Biomol. Chem. 2012; 10: 5347

      Domino reactions have attracted a great deal of attention because they provide a high degree of molecular complexity from structurally simple molecules. These reactions are typically economic and more environmentally friendly as they avoid the isolation of unstable intermediates and reduce production costs. See:
    • 24a Tietze LF. Chem. Rev. 1996; 96: 115
    • 24b Breinbauer R. Synthesis 2007; 794
    • 24c Mueller TJ. J. Angew. Chem. Int. Ed. 2007; 46: 2977
    • 24d Bogdanowicz-Szwed K, Krasodomska M, Krasodomski W. Wiad. Chem. 1997; 51: 643
    • 24e Pellissier H. Chem. Rev. 2013; 113: 442
    • 24f Parsons PJ, Penkett CS, Shell AJ. Chem. Rev. 1996; 96: 195
    • 24g Tietze LF, Duefert SC, Clerc J, Bischoff M, Maass C, Stalke D. Angew. Chem. Int. Ed. 2013; 52: 3191
  • 25 Secondary alcohol 20a is highly unstable to column chromatographic purification and the sodium borohydride reduction of 3-acetylindole in EtOH at r.t. gave single compound 20a, which was used directly in the control experiment without any further purification.
  • 26 The reaction intermediate might be unstable at 100 °C and may decompose before it reacts. This speculation might explain the low yield at 100 °C. Intermediate 20a might be somewhat stable at 80 °C, the temperature at which it gives the maximum yield. At lower temperature, product formation may be sluggish.
  • 27 Ganguly NC, Mondal P, Barik SK. Green Chem. Lett. Rev. 2012; 5: 73
  • 28 Ramachandiran K, Muralidharan D, Perumal PT. Tetrahedron Lett. 2011; 58: 5
  • 29 Xu HY, Zi Y, Xu XP, Wang SY, Ji SJ. Tetrahedron 2013; 69: 1600
  • 30 Thirupathi P, Kim SS. J. Org. Chem. 2010; 75: 5240
  • 31 Mendes SR, Thurow S, Fortes MP, Penteado F, Lenardão EJ, Alves D, Perin G, Jacob RG. Tetrahedron Lett. 2012; 53: 5402

Zoom Image
Scheme 1 Synthesis of bisindolylmethane 2 from indole
Zoom Image
Figure 1 Ligand screening for the domino synthesis of 2
Zoom Image
Scheme 2 A plausible mechanism for the domino synthesis of bisindolylmethanes
Zoom Image
Scheme 3 Synthesis of bisindolylmethane 2 from indole and the secondary alcohol 20a