1,166
Views
3
CrossRef citations to date
0
Altmetric
Research Article

Palladium-catalyzed amination of 2-chlorothienopyridone with primary aromatic amines

, & | (Reviewing editor)
Article: 1567894 | Received 03 Jan 2018, Accepted 07 Jan 2019, Published online: 12 Feb 2019

Abstract

A series of ethyl 7-cyclopropyl-2-(arylamino)-3-nitro-4-oxo-4,7-dihydrothieno[2,3-b]pyridine-5-carboxylate (9a-d) were prepared by coupling of ethyl 7-cyclopropyl-2-chloro-3-nitro-4-oxo-4,7-dihydrothieno[2,3-b]pyridine-5-carboxylate (8) with primary aromatic amines via palladium-catalyzed amination using palladium acetate Pd(OAc)2 in the presence of cesium fluoride in good yields. The new compounds were characterized by 1H-NMR, 13C-NMR, mass spectrometry, high resolution mass spectrometry and elemental analysis. 7-cyclopropyl-2-(phenylamino)-3-nitro-4-oxo-4,7-dihydrothieno[2,3-b]pyridine-5-carboxylic acid (10a) showed weak activity against E.aerogenas and S.aureus bacteria.

PUBLIC INTEREST STATEMENT

Antibiotic resistance is one of the biggest threats to global health, which is rising to alarming levels. The current research is the successful addition to the field of medicinal chemistry, in order to synthesize a new generation of antibiotics, which is bioisosters of fluoroquinolones antibiotics.

Competing interests

The authors declare no competing interests.

1. Introduction

The quinolones are synthetic antibiotics, which trace their origin to a concerted effort by scientists to synthesis novel antibacterial agents. After the discovery of nalidixic acid (1) (Figure ) by Lesher in 1962 (Lesher, Froelich, Gruett, Bailey, & Brundage, Citation1962), an antibacterial agent, which is indicated for the treatment of urinary tract infection, the research in this area became more extensive. Fluoroquinolones are direct inhibitors of bacterial DNA synthesis (Dilica & Zhao, Citation1997). Fluoroquinolones inhibit two bacterial enzymes, DNA gyrase and topoisomerase IV, which have a distinct role in DNA replication. They bind to the complex of each of these enzymes with DNA, the resulting complexes, including the drug, block progress of the DNA replication enzyme complex. This results in damage to bacterial DNA and bacterial cell death.

Figure 1. Structures of nalidixic acid, norfloxacine, and ciprofloxacin.

Figure 1. Structures of nalidixic acid, norfloxacine, and ciprofloxacin.

Structure–activity relationship studies established that the following structural requirements are essential for the activity of fluoroquinolone: an amino group at C7, a small group such as methyl, ethyl or cyclopropyl at N1 and carboxylic acid group at C3, fluorine atom at C6 increases antibacterial activity, while piperazine moiety at C7 confers antipseudomonal properties. This research resulted in the discovery of the first fluoroquinolone, norfloxacine (2) (Figure ) (Bryskier, Citation2005; Dalhoff & Schmitz, Citation2003; De Souza, Citation2005; Emami, Shafiee, & Foroumadi, Citation2006; CitationGootz & Brighty, 198; Peterson, Citation2001; Wagman & Wentland, Citation2007), which is characterized by a piperazine moiety at C7 and the most active fluoroquinolone, ciprofloxacin (3) (Figure ) (Itoh et al., Citation1980; Koga, Itoh, Murayama, Suzue, & Irikura, Citation1980). The structure–activity relationship of quinolones has been the subject of extensive reviews (Felmingham et al., Citation1985; Maurer & Grohe, Citation1986; Petersen et al., Citation1996; Wise, Andrews, & Edwards, Citation1983).

The synthesis of aromatic amines has attracted much attention due to their role in many fields of science and industry. They are common motifs in natural products, pharmaceuticals, agrochemicals, dyes and polymers. An aryl-nitrogen linkage is included in molecules like chloroquine (4) (Figure ) (Johnson & Buell, Citation1952), a widely used antimalarial drug and dichlophenac (5) (Figure ) (Skoutakis et al., Citation1988), a 5HT receptor antagonist. Dichlophenac (sodium[(2-[(2,6-dichlorophenyl)amino]phenyl]acetate)), is a potent NSAID (nonsteroidal anti-inflammatory drug), therapeutically used in inflammatory and painful diseases of rheumatic and nonrheumatic origin.

Figure 2. Structures of chloroquine(4) and diclophenac(5).

Figure 2. Structures of chloroquine(4) and diclophenac(5).

Despite the simplicity of arylamine moiety, the synthesis of these molecules is often difficult and challenging. This class of compounds was prepared via classical copper-mediated Ullmann coupling (Scheme ) (Ullmann, Citation1903) and recently developed palladium-catalyzed aryl coupling (Schemes and ) (Hamann & Hartwig, Citation1998; Old, Wolfe, & Buchwald, Citation1998) are the more commonly used methods. The Ullmann coupling often requires high temperature and the use of copper salts in greater than stoichiometric amounts. The reaction is also very sensitive to the substitution on the aryl halide. Due to these limitations, copper salts have been supplanted by palladium catalyst (Hamann & Hartwig, Citation1998; Old et al., Citation1998).

Scheme 1. Ullmann coupling.

Scheme 1. Ullmann coupling.

Scheme 2. Migita palladium-catalyzed amination of aryl halides with aminotin compounds.

Scheme 2. Migita palladium-catalyzed amination of aryl halides with aminotin compounds.

Scheme 3. Buchwald-Hartwig coupling.

Scheme 3. Buchwald-Hartwig coupling.

Palladium-catalyzed coupling of amines with aryl halides has become the most versatile tool for the preparation of aryl amines. This method has been introduced to provide an easy access to diarylamines. Migita and coworkers reported the first palladium-catalyzed amination of aryl halides in 1983 (Scheme ) (Kosugi, Kameyama, & Migita, Citation1983). The treatment of bromobenzene with the aminotin compound in the presence of a palladium catalyst provided N, N-diethylaniline. However, this method requires the use of a stoichiometric quantity of a toxic and moisture sensitive tin reagent.

Buchwald, Guram, & Rennels (Citation1995) and Hartwig and Lounie (Citation1995) independently reported the palladium-catalyzed amination of aryl bromides with secondary amines using NaOtBu or LiHMDS as a base (Scheme ) to afford tertiary arylamines in good to excellent yields. The palladium-catalyzed amination of aryl halides has been the subject of the extensive reviews (Hartwig, Citation1998; Wolfe, Wagaw, Marcoux, & Buchwald, Citation1998; Yang & Buchwald, Citation1999).

Several thienopyridones, namely;4-oxothieno[2,3-b]pyridine-5-carboxylic acid, potential bioisosteres of quinolone antibacterial agents were prepared and bioassayed (Al-Masoudi, Al-Soud, & Al-Masoudi, Citation2000; Al-Trawneh et al., Citation2011; Bacon & Daun, Citation1991; Bompart, Citation1988; Bompart, Giral, Malicorne, & Puygrenier, Citation1987; Bravic, Cotrait, Bompart, & Giral, Citation1989; El-Abadelah et al., Citation1997; El-Abadelah, Sabri, & Al-Ashqar, Citation1997; El-Abadelah et al., Citation1998; Giral, Bompart, & Puygrenier, Citation1985; Hooper, Utsunomiya, & Hartwig, Citation2003; Yamazaki, Matsubara, Morishima, & Suenaga, Citation1983). 2-Chloro-4-oxo-4,7-dihydrothieno[2,3-b]pyridine-5-carboxylic acids (6) (Figure ) has been prepared and was found to possess a good level of activity against gram-negative bacteria, while 7-Cyclopropyl-4,7-dihydro-2-(4-methyl-l-piperazinyl)-3-nitro-4-oxothieno[2,3-b]pyridine-5-carboxylic acid(7) (Figure ) showed reduced activity (El-Abadelah et al., Citation1998).

Figure 3. Structures of some thieno[2,3-b]pyridine-5-carboxylic acids.

Figure 3. Structures of some thieno[2,3-b]pyridine-5-carboxylic acids.

In view of the fact that most potent quinolones are substituted by different amino groups at C7 and since substitution at C2 and C3 positions of 4-oxo-4,7-dihydrothieno[2,3-b]pyridine-5-carboxylic acid has not been sufficiently explored. This work involves the development of palladium-catalyzed amination of ethyl 2-chloro-3-nitro-4-oxo-4,7-dihydrothieno[2,3-b]pyridine-5-carboxylate(8) with primary aromatic amines to produce ethyl 7-cyclopropyl-2-(arylamino)-3-nitro-4-oxo-4,7-dihydrothieno[2,3-b]pyridine-5-carboxylate (9a-d, Scheme ). Herein; their synthesis, characterization and biological activity are described.

2. Results and discussion

2.1. Synthesis

The title synthon (8), ethyl 2-chloro-3-nitro-4-oxo-4,7-dihydrothieno[2,3-b]pyridine-5-carboxylate, required in this work, is prepared from 2,5-dichlorothiophene-3-carboxylic acid, ethyl 3-(N,N-dimethylamino)acrylate and cyclopropyl amine by following the stepwise synthetic procedure reported in literature (Scheme ). The overall yield of 8 starting from 2,5-dichlorothiophene-3-carboxylic acid is 40%. The latter compound, 2,5-dichlorothiophene-3-carboxylic acid was prepared by reaction of 3-acetyl-2,5-dichlorothiophene with Br2/NaOH, following the literature procedure (El-Abadelah et al., Citation1997). Limited quantities of (8) have been prepared by this methodology, after sequential, meticulous and time-consuming steps.

Scheme 4. Synthesis of ethyl 2-chloro-3-nitro-4-oxo-4,7-dihydro thieno[2,3-b]pyridine-5-carboxylate(8).

Scheme 4. Synthesis of ethyl 2-chloro-3-nitro-4-oxo-4,7-dihydro thieno[2,3-b]pyridine-5-carboxylate(8).

Attempt to bring about direct amination of 2-chloro substituent compound(8) by anisidine, in absence of Pd(OAc)2 (DMF at 120°C for 48 h) have, thus far, been unsuccessful, no coupled product (9a) have been isolated, as evidenced from 1H-NMR analysis of the reaction mixture, and the recovery of the starting synthon (8) (70%).

While Ullmann reaction utilize copper for the formation of C-N bonds. Buchwald and Hartwig reported palladium-catalyzed amination of inactivated aryl bromides and aryl halides with low catalyst loading under mild conditions. Although, palladium is more expensive metal than copper, in addition to using sterically hindered, expensive phosphine ligands. Palladium-catalyzed amination of aryl halides has provided easy access to diarylamines. Bulky trialkylphosphines has been used for coupling of primary and secondary anilines (Wolfe, Hiroshi, Sadighi, Jingjun., & Buchwald, Citation2000) with inactivated, sterically hindered aryl chlorides (Scheme ).

Scheme 5. Palladium-catalyzed synthesis of diaryl amines.

Scheme 5. Palladium-catalyzed synthesis of diaryl amines.

However, the application of Buchwald-Hartwig methodology to electron rich five-membered heteroaromatic substances, such as halo furans, thiophenes, pyrroles, indoles and imidazoles has been limited. Hartwig (Hooper et al., Citation2003) reported that bromothiophenes and chlorothiophenes react with diphenylamine and N-methylaniline in the presence of palladium catalyst using PtBu3 ligand (Scheme ).

Scheme 6. Palladium-catalyzed synthesis of 2-(N-methyl-N-phenylamino)thiophene.

Scheme 6. Palladium-catalyzed synthesis of 2-(N-methyl-N-phenylamino)thiophene.

Preparation of the target, ethyl 7-cyclopropyl-2-(arylamino)-3-nitro-4-oxo-4,7-dihydro thieno[2,3-b]pyridine-5-carboxylate (9a-d) were carried out via palladium-catalyzed amination using Pd(OAc)2 as a catalyst, with 14% loading, DMF as solvent and CsF as a base at 80°C for 48 h (Scheme ). The cross-coupling products (9a-d), were isolated in 60–68% yield.

Scheme 7. Palladium-catalyzed coupling of thienopyridone(8) with aniline derivatives.

Scheme 7. Palladium-catalyzed coupling of thienopyridone(8) with aniline derivatives.

It is worth to mention that no cross amination products were isolated using potassium t-butoxide or calcium carbonate as bases. On the other hand, Pd(OAc)2 catalyzed amination of synthon (8) with p-chloroaninline, p-bromoaninline, p-trifluoromethylaninline and 2,4-difluoroaniline produces the cross products (9e-h) in trace amounts, while p-nitroaniline and p-aminobenzenesulfamide failed to give products (9i-j). These results can be explained based on the nucleophilicity of nitrogen atom of substituted aniline, where electron-releasing groups enhance nucleophilicity, while electron-withdrawing reduce the nucleophilicity.

The new compounds (9a-d) were characterized by elemental analyses, IR. MS, HRMS,1H-NMR and 13C-NMR spectral data. These data detailed in the experimental part are consistent with the predicted structures. Thus, the mass spectra display the correct molecular ion peaks for which the measured high resolution (HRMS) data are in good agreement with calculated values.

In conclusion, this is the first example of palladium-catalyzed amination of chlorothienopyridone (8) with anilines, which have reduced nitrogen nucleophilicity, in absence of sterically hindered phosphine ligands. This catalytic activity of palladium acetate may be explained in terms of the presence of keto-ester functionality in ethyl 2-chloro-3-nitro-4-oxo-4,7-dihydrothieno[2,3-b]pyridine-5-carboxylate(8), which act as a chelate ligand to stabilizes catalytic palladium intermediate. These data and results encourage further investigations to explore the coupling of more anilines having electron-releasing groups, as well as, sterically hindered anilines.

2.2. Biology

The antibacterial activities of compounds 9(a-d) and 10a were assayed against E-coli, B.subtilis, M. luteus, S. aureus, E. aerogenas and P. aeruginosa. The initial investigation by agar diffusion test show that ethyl 7-cyclopropyl-2-(arylamino)-3-nitro-4-oxo-4,7-dihydrothieno[2,3-b]pyridine-5-carboxylate (9a-d) have no activity against all tested bacteria; while 7-cyclopropyl-2-(phenylamino)-3-nitro-4-oxo-4,7-dihydrothieno[2,3-b]pyridine-5-carboxylic acid (10a) showed weak activity against Gram-negative, E. aerogenas and Gram-positive S. aureus bacterial (Table ). 2-chloro-7-cyclopropyl-3-nitro-4-oxo-4,7-dihydrothieno[2,3-b]pyridine-5-carboxylic acid (11) showed promising activities against E. aerogenas and S. aureus (Table ) (Felmingham et al., Citation1985). In terms of structure–activity relationship replacing chlorine atom at position 2 by arylamino group results in decreased activity for 3-nitro-4-oxo-4,7-dihydrothieno[2,3-b]pyridine-5-carboxylic acid.

Table 1. Inhibition zone of compound 10a and 11 against E. aerogenas and S. aureus

3. Experimental

3.1. Chemistry

Palladium (II) acetate, caesium fluoride, cyclopropylamine, aniline, p-fluoroaniline, p-toludine and 3-acetyl-2,5-dichlorothiophene was purchased from Aldrich. Sodium hydride, sodium hydroxide, aluminium chloride, magnesium sulfate, triethylamine, dimethylformamide and ethyl 3-(N,N-dimethylamino)acrylate was purchased from Acros. p-Anisidine was purchased from Fluka . Silica gel for column chromatography was purchased from Macherew-Nagel GmbH & Co. KG (Germany). Carbon disulfide (CS2) was dried over anhydrous phosphorous pentoxide (P2O5) and distilled. Thionyl chloride (SOCl2) was purified by fractional distillation. Benzene and tetrahydrofuran were dried over sodium metal and distilled under a nitrogen atmosphere. Melting points were determined on a scientific melting point apparatus in open-capillary tubes. 1H-NMR and 13C-NMR spectra were recorded on 400 MHz spectrometer (Bruker Avance III 400 MHZ) (Al-Yarmouk University/Jordan) with TMS as the internal standard for solutions in CDCl3, chemical shifts are expressed in δ units. Electron impact mass spectra (EIMS) were obtained using MASPEC system msw/A017(Al-Yarmouk University/Jordan). High-resolution mass spectra (HRMS) were measured by electrospray ionization (ESI) technique on Bruker APEX-IV instrument (University of Jordan/Jordan), the samples were dissolved in chloroform, diluted in spray solution (methanol/chloroform) and infused using a syringe pump with a flow rate of 2 µL/min. Elemental analysis (CHNS) were determined by EA.96-mth (Al-Bayt University/Jordan). IR spectra were recorded on a MATTSON 500 FTIR spectrophotometer (Mutah University/Jordan).

3.2. Determination of inhibition zones (agar diffusion test)

The test microorganisms were cultured in a nutrient broth for 16 h. Bacterial test plates were prepared in nutrient agar medium in a density of 106 bacterial cell/ml. Aliquots of 0.5 and 0.3 mg/ml of freshly prepared solutions of compounds 10a and 11 dissolved in DMSO were placed in a blank disk of 6 mm in diameter. Plates were then incubated at 37° C for 24 h. The zone of inhibition was determined as the diameter of the zone of inhibition around the disk.

3.2.1. Ethyl 2-chloro-7-cyclopropyl-3-nitro-4-oxo-4,7-dihydrothieno[2,3-b] pyridine-5-carboxylate (8)

This compound was prepared from 3-acetyl-2,5-dichlorothiophene, ethyl 3-(N,N-dimethylamino) acrylate and cyclopropylamine by following literature procedures (El-Abadelah et al., Citation1997)

3.2.2. General procedure for the preparation of ethyl 7-cyclopropyl-2-(N-arylamino)-3-nitro-4-oxo-4,7-dihydrothieno[2,3-b]pyridine-5-carboxylates (9a-d)

The particular aniline (~1.5 mmol), Pd(OAc)2 (0.03 g, 0.14 mmol, 15%) and CsF (0.26 g, 1.96 mmol), were added to a stirred solution of ethyl 2-chloro-7-cyclopropyl-3-nitro-4-oxo-4,7-dihydrothieno[2,3-b]pyridine-5-carboxylate (8) (0.20 g, 0.58 mmol) in DMF (5 ml). The reaction mixture was heated at 80–90°C for 48 h, the resulting solution was then cooled to room temperature, water (10 ml) was added, yellow to brown solid precipitate was obtained that dissolved in CHCl3 and filtered to remove the insoluble material. The chloroform solution was then dried over MgSO4 and evaporated under reduced pressure to give a brown to yellow residue, crude product was purified by column chromatography using silica gel and eluting with chloroform, ethyl acetate or acetonitrile.

3.2.2.1. Ethyl 7-cyclopropyl-2-(phenylamino)-3-nitro-4-oxo-4,7-dihydrothieno[2,3-b] pyridine-5-carboxylate (9a)

This compound was prepared according to general procedure from ethyl 2-chloro-7-cyclopropyl-3-nitro-4-oxo-4,7-dihydrothieno[2,3-b]pyridine-5-carboxylate(8) (0.20 g, 0.58 mmol) and aniline (0.10 g, 1.07 mmol) to give a brown solid, reaction temperature 80–90°C; reaction time 24 h; ratio of the eluting mixture 1:2, v/v [CHCl3:CH3CN]. Yield 0.16 g (68%), m.p 237–240°C.Anal:calcd for C19H17N3O5S (399.42): C,57.13; H,4.29; N,10.52; S, 8.03. Found: C,57.62; H,4.43; N,10.58; S,8.02. IR: νmax(KBr)/cm−1 3457, 1723, 1622, 1593, 1561, 1487, 1323, 1300, 1250, 1171, 1117, 1038, 812, 766. 1H-NMR: (400 MHz, CDCl3): δ 1.31(m, 2H) and 1.35(m, 2H)(H-2′/H-3′), 1.42(t,3J = 7.2Hz, 3H, CH3CH2O-), 3.56(m, 1H, H-1′), 4.43(q,3J = 7.2 Hz, 2H, OCH2CH3), 6.96(d, 3J = 7.2 Hz, 2H/H-2′′,H-6′′), 7.17(t, 3J = 7.6 Hz, 1H, H-4′′), 7.33(dd, 3J = 7.2 Hz, 4J = 3.2 Hz 2H, H-3′′/H-5′′), 8.45(s, 1H, H-6), 11.56 (s, 1H, H-N). 13C-NMR: (100 MHz, CDCl3): δ 7.9(C-2′, C-3′), 14.4(CH3CH2O), 36.6(C-1′), 61.7(-OCH2CH3), 116.7(C-5), 119.6(C-4′′), 120.1(C-2′′,C-6′′), 120.3(C-3a), 125.1(C-1′′), 128.3(C-3′′,C-5′′), 138.7 (C-2), 140.9(C-3), 147.0(C-6), 153.1(C-7a), 163.9(CO2Et), 173.6 (C-4). HRMS (ESI): found 400.09617 ([M+H]+) C19H18N3O5S requires 400.09672; found 422.07811 ([M+Na]+) C19H17N3NaO5S requires 422.0787; m/z (EI): 399.5 (M+,13).

3.2.2.2. Ethyl 7-cyclopropyl-2-(4-methylphenylamino)-3-nitro-4-oxo-4,7-dihydrothieno[2,3-b]pyridine-5-carboxylate (9b)

This compound was prepared according to general procedure from ethyl 2-chloro-7-cyclopropyl-3-nitro-4-oxo-4,7-dihydrothieno[2,3-b]pyridine-5-carboxylate(8) (0.20 g,0.58 mmol) and p-toludine (0.09 g, 0.84 mmol) to give a brown solid, reaction temperature: 80–90°C; reaction time: 24 h; ratio of the eluting mixture: 1:3,v/v[CHCl3:CH3COOC2H5]. Yield = 0.14 g (58%), m.p 198–200°C. Anal: calcd for C20H19N3O5S (413.45): C,58.10; H,4.63; N,10.16; S, 7.76. Found: C,58.06; H,4.90; N,9.88; S,7.48 . IR: νmax(KBr)/cm−1 3447, 1730, 1699, 1632, 1557, 1491, 1458, 1343, 1317, 1250, 1161, 1125, 814, 741 . 1H-NMR: (400 MHz, CDCl3): δ 1.31(m, 2H) and 1.34(m, 2H)(H-2′/H-3′), 1.42(t, 3J = 8.0 Hz, 3H, CH3CH2O-), 3.55 (m, 1H, H-1′), 2.34 (s, 3H, C(4′′)-CH3), 4.43(q, 3J = 8.0 Hz, 2H, OCH2CH3), 6.85(d,3J = 8.0 Hz, 2H, H-2′′/H-6′′), 7.12(d, 3J = 8.0 Hz, 2H, H-3′′/H-5′′), 8.42(s, 1H, H-6), 11.54(s, 1H, H-N). 13C-NMR: (100MHz, CDCl3): δ 7.9(C-2′,C-3′), 14.4(CH3CH2O), 21.1(C(4′′)-CH3), 36.6(C-1′), 61.7(-OCH2CH3), 116.6(C-5), 119.1(C-3a), 120.1(C-2′′,C-6′′), 120.3(C-1′′), 128.8(C-3′′,C-5′′), 134.8(C-4′′), 136.2(C-2), 141.1(C-3), 146.9(C-6), 153.2(C-7a), 163.9(-CO2Et), 173.5(C-4). HRMS: (ESI): found 436.09376 ([M+Na]+) C20H19N3NaO5S requires 436.0943; m/z (EI): 414 (M+,68).

3.2.2.3. Ethyl 7-cyclopropyl–2-(4-methoxphenylamino)-3-nitro-4-oxo-4,7-dihydrothieno- [2,3-b] pyridine-5-carboxylate (9c)

This compound was prepared according to general procedure from ethyl 2-chloro −7-cyclopropyl-3-nitro-4-oxo-4,7-dihydrothieno[2,3-b]pyridine-5-carboxylate (8) (0.20 g, 0.58 mmol) and p-anisidine (0.16 g, 1.30 mmol) to give a brown solid, reaction temperature: 80–90°C; reaction time: 24 h; ratio of the eluting mixture 1:1;v/v[CHCl3:CH3CN]. Yield 0.17g (68%), m.p:192–195°C. Anal: calcd, for C20H19N3O6S (429.4463): C,55.94; H,4.46; N,9.78; S,7.47. found:C,55.82; H,4.57; N,10.16; S,7.38. IR: νmax(KBr)/cm−1 3447, 1732, 1699, 1632, 1562, 1491, 1346, 1314, 1248, 1121, 1032, 808. 1H-NMR: (400 MHz, CDCl3); δ = 1.33(m,2H) and 1.37(m,2H), (H2-2′/H2-3′), 1.42(t, 3J = 7.2Hz, 3H, CH3CH2O-), 3.56(m, 1H, H-1′), 3.82(s, 3H, C(4′′)-OCH3), 4.43(q,3J = 7.2 Hz, 2H, OCH2CH3), 6.86(d, 3J = 9.2Hz, 2H/H-2′′, H-6′′), 6.92(d, 3J = 9.2 Hz, 2H, H-3′′/H-5′′), 8.44(s, 1H, H-6), 11.57 (s, 1H, H-N). 13C-NMR: (100 MHz, CDCl3): δ 7.9(C-2′, C-3′), 14.4(CH3CH2O), 36.6(C-1′), 55.4(C(4′′)-OCH3), 61.7(-OCH2CH3), 113.5(C-2′′,C-6′′), 116.7(C-5), 118.6(C-3a), 120.3(C-1′′), 121.7(C-3′′,C-5′′), 132.0(C-2), 141.4(C-3), 146.9(C-6), 153.2(C-7a), 157.0(C-4′′), 164.0(-CO2Et), 173.5(C-4). HRMS (ESI): found 430.10673 ([M+H]+), C20H20N3O6S requires 430.1073; found 452.08868 ([M+Na]+), C20H19N3NaO6S requires 452.0892;m/z (EI):429 (M+,35).

3.2.2.4. Ethyl 7-cyclopropyl-2-(4-fluorophenylamine)-3-nitro-4-oxo-4,7-dihydrothieno-[2,3-b]pyridine-5-carboxylate (9d)

This compound was prepared according to general procedure from ethyl 2-chloro-7-cyclopropyl-3-nitro-4-oxo-4,7-dihydrothieno[2,3-b]pyridine-5-carboxylate(8) (0.20 g, 0.58 mmol) and p-fluoroaniline (0.10 g,0.90 mmol) to give a yellow solid (reaction temperature 80–90°C); reaction time 24 h; ratio of the eluting mixture 1:2,v/v[CHCl3:CH3CN]. Yield 0.12 g (50%), m.p 220-225°C. Anal: calcd for C19H16FN3O5S (417.41): C,54.67; H,3.86; N,10.07; S, 7.68. Found: C,53.79; H,3.79; N,9.75; S,7.38 . IR: νmax(KBr)/cm−1 3457, 1723, 1622, 1593, 1561, 1487, 1323, 1300, 1249, 1171, 1117, 1038, 812, 766. 1H-NMR:(400 MHz,CDCl3): δ 1.33(m,2H) and 1.36(m,2H) (H2-2′/H2-3′), 1.42(t,3J = 7.2 Hz, 3H,CH3CH2O-), 3.57(m, 1H, H-1′), 4.44(q,3J = 7.2 Hz, 2H, OCH2CH3), 6.95(dd, 3J = 6.8 Hz, 4JH-F = 2.4 Hz, 2H/H-2′′,H-6′′), 7.03(dd, 3J = 6.8 Hz, 3JH-F = 8.8 Hz, 2H, H-3′′/H-5′′), 8.46 (s, 1H, H-6), 11.54 (s, 1H, H-N). 13C-NMR: (100 MHz, CDCl3): δ 7.9(C-2′,C-3′), 14.4 (CH3CH2O), 36.6(C-1′), 61.8(-OCH2CH3), 115.1(d, 2JC-F = 23Hz, C-3′′,C-5′′), 116.7(C-5), 120.2 (C-3a), 121.9(d, 3JC-F = 8 Hz C-2′′,C-6′′), 128.0(C-1′′), 135.0(C-2), 141.3(C-3), 147.0(C-6), 153.2(C-7a), 159.9(d, 1JC-F = 244 Hz, C-4′′), 163.9(-CO2Et), 173.6(C-4). HRMS (ESI): found 418.08675 ([M+H]+) C19H17FN3O5S requires 418.0873; found 440.06869 ([M+Na]+) C19H16FN3NaO5S requires 440.0692;m/z (EI):417(M+,26).

3.2.2.5. Ethyl 7-cyclopropyl-2-(phenylamino)-3-nitro-4-oxo-4,7-dihydrothieno[2,3-b] pyridine-5-carboxylic acid (10a)

This compound was prepared by hydrolysis of ethyl 7-cyclopropyl-2-(phenylamino)-3-nitro-4-oxo-4,7-dihydrothieno[2,3-b]pyridine-5-carboxylate (9a) (0.04 g, 0.1 mmol) in 10% HCl (8ml) and few drops of ethanol, reaction temperature: 80–85°C under reflux condition; reaction time: 24 h. Yield 0.03 g (81%), m.p: 285–287°C.1H-NMR:(400 MHz, CDCl3): δ = 1.26(m, 2H) and 1.36(m, 2H)(H2-2′/H2-3′), 3.54(m, 1H, H-1′), 7.16(s, 3H, H-OCH3), 7.34(d,3J = 7.6 Hz, 2H, H-3′′/H-5′′), 7.09(d, 3J = 7.09 Hz, 2H, H-2′′/H-6′′), 8.92(s, 1H, H-6), 10.72(s, 1H, H-N). 13C-NMR: (100 MHz, CDCl3): δ 7.10(C-2′, C-3′), 37.60(C-1′), 118.9(C-2′′,C-6′′), 119.6(C-3a), 113.2(C-6), 118.3(C-5), 128.3(C-3′′,C-5′′), 148.4(C-3), 154.5(C-7a), 164.3(CO2-), 124.6(C-4′′) and 175.4(C-4). m/z (EI):371 (M+,49).

Cover image

Source:

Additional information

Funding

This work was supported by the Mutah Univercity/ Jordan [120/14/833].

Notes on contributors

Samir A. Al-Taweel

Samir A. Al-Taweel, the corresponding author, is working as Professor in Mutah University, Karak, Jordan. He has published more than 35 research papers in peer-reviewed international journals. He has guided ten MS students. The main focus of our research group is to develop simple and efficient synthetic strategies for the synthesis of new bioactive organic thiophene-based heterocyclic compounds. Recently, our research group interested in synthesizing dye-sensitizers, for the solar cell applications.

Salah A. Al-Trawneh

Salah A. Al-Trawneh is the co-author he is the associated professor in organic chemistry, the main interest of our research group is to develop simple and efficient synthetic protocol for the synthesis of new bioactive heterocyclic organic compounds.

Wal'A M. Al-Trawneh

Wal'A M. Al-Trawneh is a Master student conducting the synthesis and purification of the coupling products.

References

  • Al-Masoudi, N. A., Al-Soud, Y. A., & Al-Masoudi, I. A. (2000). Thienopyridone nucleosides: Synthesis; antibacterial and antiviral activity of 7-(glycosyl and D-manno-hextitol-1-yl)-2-chloro-4-oxothieno[2, 3-b] pyridine-5-carboxylic acids. Sulfur Letters, 24(1), 13–11.
  • Al-Trawneh, S. A., El-Abadelah, M. M., Zahra, J. A., Al-Taweel, S. A., Zani, F., Incerti, M., … Vicini, P. (2011). Synthesis and biological evaluation of tetracyclic thienopyridones as antibacterial and antitumor agents. Bioorganic & medicinal chemistry, 19, 2541–2548. doi:10.1016/j.bmc.2011.03.018
  • Bacon, E. R., & Daun, S. J. (1991). Synthesis of 7‐ethyl‐4,7‐dihydro‐4‐oxo‐2‐(4‐pyridinyl)thieno[2,3‐b]pyridine‐5‐carboxylic acid. Journal of Heterocyclic Chemistry, 28(8), 1953–1955. doi:10.1002/jhet.v28:8
  • Bompart, J. (1988). Synthèse et activité anti-bactérienned’acides dihydro-4,7 éthyl-7 oxo-4 thiéno[2,3-b]pyridine carboxyliques-5 II. European Journal of Medicinal Chemistry, 23(5), 457–464. doi:10.1016/0223-5234(88)90143-2
  • Bompart, J., Giral, L., Malicorne, G., & Puygrenier, M. (1987). Synthèse et activité anti-bactérienned’acides dihydro-4,7 éthyl-7 oxo-4 thiéno(2,3-b)pyridine carboxyliques-5. European Journal of Medicinal Chemistry, 22(2), 139–145. doi:10.1016/0223-5234(87)90009-2
  • Bravic, G., Cotrait, M., Bompart, J., & Giral, L. (1989). Structure d’un nouvelantibactérien: L’acide chloro-2 dihydro-4,7 éthyl-7 oxo-4 thiéno[2,3-b]pyridine carboxylique-5. Acta Crystallographica Section C Crystal Structure Communications, 45(12), 1983–1986. doi:10.1107/S0108270189004038
  • Bryskier, A. (2005). Fluoroquinolones. In A. Bryskier (Ed.), Antimicrobial agents: Antibacterials and antifungals (pp. 668–788). Washington: ASM Press.
  • Buchwald, S. L., Guram, A. S., & Rennels, R. A. (1995). A simple catalytic method for the conversion of aryl bromides to arylamines. Angewandte Chemie International Editon in English, 34(12), 1348–1350. doi:10.1002/anie.199513481
  • Dalhoff, A., & Schmitz, F. (2003). In vitro antibacterial activity and pharmacodynamics of new quinolones. European Journal of Clinical Microbiology & Infectious Disease, 22, 203.
  • De Souza, M. V. N. (2005). New fluoroquinolones: A class of potent antibiotics. Mini-Reviews in Medicinal Chemistry, 5(11), 1009–1017.
  • Dilica, K., & Zhao, X. (1997). DNA gyrase, topoisomerase IV, and the 4-quinolones. Microbiology and Molecular Biology Reviews, 61(3), 377–392.
  • El-Abadelah, M. M., Nazer, M. Z., Sabri, S. S., Khalil, S. M., Voelter, W., & Geiger, M. Z. (1997). Synthesis and chiroptical properties of some n-(2-chloro-7-cyclopropyl4,7-dihydro-4-oxo-thieno[2,3-b]pyridine-5-carbonyl) L-a-amino esters. Zeitschrift für Naturforschung B, 52b, 419–426. doi:10.1515/znb-1997-0321
  • El-Abadelah, M. M., Sabri, S. S., & Al-Ashqar, H. A. (1997). Thienopyridoneantibacterials. Part I. A synthesis of some 7-alkyl-2-chloro-1,4-dihydro-4-oxothieno[2,3-b]pyridine-5-carboxylic acids. Heterocycles, 2(45), 255–264. doi:10.3987/COM-96-7613
  • El-Abadelah, M. M., Sabri, S. S., Al-Ashqar, H. A., Mion, P., Bompart, J., & Calas, M. (1998). Thienopyridinone antibacterials. Part II. Synthesis and antibacterial activity of some 2-Chloro-7-Cyclopropyl-4,7-Dihydro-4-Oxothieno[2,3-b]Pyridine-5-Carboxylic acids. Phosphorus, Sulfur, and Silicon and the Related Elements, 134/135(1), 21–29. doi:10.1080/10426509808545450
  • Emami, S., Shafiee, A., & Foroumadi, A. (2006). Structural features of new quinolones and relationship to antibacterial activity against Gram-positive bacteria. Mini-Reviews in Medicinal Chemistry, 6(4), 375–386.
  • Felmingham, D., O’Hare, M. D., Robbins, M. J., Wall, R. A., Williams, A. H., Cremer, A. W., … Gruneberg, R. N. (1985). Comparative in vitro studies with 4-quinolone antimicrobials. Drugs under Experimental and Clinical Research, 11(5), 317329.
  • Giral, L., Bompart, J., & Puygrenier, M. (1985). M. Eur. Pat. 1985, EP 161, 235. (Chem. Abstr. 1986, 104, 88515w).
  • Gootz, T. D., & Brighty, K. E. (1998). Chemistry and mechanism of action of the quinolone antibacterials. In V. T. Andriole (Ed.), The quinolones (2nd ed., pp. 29–80). San Diego: Academic Press.
  • Hamann, B. C., & Hartwig, J. F. (1998). Sterically hindered chelating alkyl phosphines provide large rate accelerations in palladium-catalyzed amination of aryl iodides, bromides, and chlorides, and the first amination of aryl tosylates. Journal of American Chemical Society, 120(29), 7369–7370. doi:10.1021/ja981318i
  • Hartwig, J. F. (1998). Transition metal catalyzed synthesis of arylamines and aryl ethers from aryl halides and triflates: Scope and mechanism. Angewandte Chemie International Edition, 37, 2046–2067. doi:10.1002/(ISSN)1521-3773
  • Hartwig, J. F., & Lounie, J. (1995). Palladium-catalyzed synthesis of arylamines from aryl halides. Mechanistic studies lead to coupling in the absence of tin reagents. Tetrahedron Letters, 36(21), 3609–3612.
  • Hooper, M. W., Utsunomiya, M., & Hartwig, J. F. (2003). Scope and mechanism of palladium-catalyzed amination of five-membered heterocylic halides. Journal of Organic Chemistry, 68, 2861–2873. doi:10.1021/jo0266339
  • Itoh, A., Hirai, K., Inoue, M., Koga, H., Suzue, S., Irikura, T., & Mitsuhashi, S. (1980). In vitro antibacterial activity of AM-715, a new nalidixic acid analog. Antimicrobial Agents and Chemotherapy, 17(2), 103–108.
  • Johnson, W. S., & Buell, B. G. (1952). A new synthesis of chloroquine. Journal of American Chemical Society, 74(18), 4513–4516. doi:10.1021/ja01138a014
  • Koga, H., Itoh, A., Murayama, S., Suzue, S., & Irikura, T. (1980). Structure-activity relationships of antibacterial 6,7- and 7,8-disubstituted 1-alkyl-1,4-dihydro-4-oxoquinoline-3-carboxylic acids. Journal of Medicinal Chemistry, 23(12), 1358–1363.
  • Kosugi, M., Kameyama, M., & Migita, T. (1983). Cheminform abstract: Palladium‐catalyzed aromatic amination of aryl bromides with N,N-diethylaminotributyltin. Chemistry Letters, 19, 927–928. doi:10.1246/cl.1983.927
  • Lesher, G. Y., Froelich, E. J., Gruett, M. D., Bailey, J. H., & Brundage, R. P. (1962). 1,8-Naphthyridine derivatives. A new class of chemotherapeutic agents. Journal of Medicinal Chemistry, 5, 1063–1065. doi:10.1021/jm01240a021
  • Maurer, F., & Grohe, K. (1986).2,4-Dichloro-5-fluorobenzoic acid. Ger. Offen. 3,435, 392. (Chem. Abstr. 1986, 105, 97158e).
  • Old, D. W., Wolfe, J. P., & Buchwald, S. L. (1998). A highly active catalyst for palladium-catalyzed cross-coupling reactions: Room-temperature suzuki couplings and amination of unactivated aryl chlorides. Journal of American Chemical Society, 120(37), 9722–9723.
  • Petersen, U., Bartel, S., Bremm, K.-D., Himmler, T., Krebs, A., & Schenke, T. (1996). The synthesis and biological properties of 6-fluoroquinolonecarboxylic acids. The Bulletin des Sociétés Chimiques Belges, 105, 683–699.
  • Peterson, L. R. (2001). Quinolone molecular structure-activity relationships: What we have learned about improving antimicrobial activity. Clinical Infectious Diseases, 33, S180. doi:10.1086/cid.2001.33.issue-s3
  • Skoutakis, V. A., Carter, C. A., Mickle, T. R., Smith, V. H., Arkin, C. R., Alisantros, J., & Petty, D. A. (1988). Review of diclofenac and evaluation of its place in therapy as a nonsteroidal antiinflammatory agent. Drug Intelligence & Clinical Pharmacy, 22(11), 850–859. doi:10.1177/106002808802201102
  • Ullmann, F. (1903). UebereineneueBildungsweise von Diphenylaminderivaten. Berichte der Deutschen Chemischen Gesellschaft, 36(2), 2382–2384. doi:10.1002/(ISSN)1099-0682
  • Wagman, A. S., & Wentland, M. P. (2007). Quinolone antibacterial agents. In J. B. Taylor & D. J. Triggle (Eds.), Comprehensive medicinal chemistry II (Vol. 7, pp. 567–596). Oxford: Elsevier Ltd.
  • Wise, R., Andrews, J. M., & Edwards, L. (1983). In vitro activity of bay 09867, a new quinoline derivative, compared with those of other antimicrobial agents. Antimicrobial Agents and Chemotherapy, 23(4), 559–564. doi:10.1128/AAC.23.4.559
  • Wolfe, J. P., Hiroshi, T., Sadighi, J. P., Jingjun, Y., & Buchwald, S. L. (2000). Simple, efficient catalyst system for the palladium-catalyzed amination of aryl chlorides, bromides, and triflates. Journal of Organic Chemistry, 65, 1158–1174.
  • Wolfe, J. P., Wagaw, S., Marcoux, J.-F., & Buchwald, S. L. (1998). Rational development of practical catalysts for aromatic carbon−nitrogen bond formation. Accounts of Chemical Research, 31(12), 805–818. doi:10.1021/ar9600650
  • Yamazaki, T., Matsubara, Y., Morishima, K., & Suenaga, I. (1983). Takeda Kenkyushoho 1983, 42, 297-307. (Chem. Abstr., 100, 203171n).
  • Yang, B. H., & Buchwald, S. L. (1999). Palladium-catalyzed amination of aryl halides and sulfonates. Journal of Organometallic Chemistry, 576(1–2), 125–146. doi:10.1016/S0022-328X(98)01054-7