Next Article in Journal
MMPBSA Decomposition of the Binding Energy throughout a Molecular Dynamics Simulation of Amyloid-Beta (Aß10−35) Aggregation
Previous Article in Journal
Iodine-124: A Promising Positron Emitter for Organic PET Chemistry
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Access to Original Vinylic Chlorides in the Quinazoline Series via a Monoelectronic Transfer Reaction Approach

by
Martine Maillard-Boyer
,
Caroline Castera-Ducros
,
Pierre Verhaeghe
,
France Sifredi
,
Pascal Rathelot
and
Patrice Vanelle
*
Laboratoire de Pharmacochimie Radicalaire, Faculté de Pharmacie, Universités d’Aix-Marseille I, II et III - UMR CNRS 6264, Laboratoire Chimie Provence, 27 Boulevard Jean Moulin, 13385 Marseille cedex 05, France
*
Author to whom correspondence should be addressed.
Molecules 2010, 15(4), 2719-2729; https://doi.org/10.3390/molecules15042719
Submission received: 26 January 2010 / Revised: 21 February 2010 / Accepted: 12 April 2010 / Published: 13 April 2010

Abstract

:
A series of new quinazoline derivatives bearing a vinylic chloride group on the 2-position was prepared by using a consecutive SRN1 / ERC1 radical strategy.

Graphical Abstract

1. Introduction

The quinazoline ring is of major pharmaceutical interest, considering both that it is the molecular scaffold of several drug-compounds (Erlotinib, Lapatinib, Prazozin, Alfuzosin…) and, that 4-anilino-substituted-quinazolines are key structures for the development of new selective anticancer chemotherapies, as kinase inhibitors. Focusing on the synthesis of new 2,4-bisubstituted-quinazolines, with a scope of studying their biological properties, our research team also discovered that some 4-anilinoquinazolines have original in vitro antiplasmodial properties [1,2]. Quite recently, we prepared some 4-anilinoquinazolines, substituted at position 6 with a nitro group and at position 2 with a vinylic group, one of them exerting in vitro antiplasmodial properties [3]. To explore more precisely the relations between chemical structure and biological activity, we studied the possibility of synthesizing new 4-aminated-quinazoline derivatives presenting a vinylic chloride substituent at position 2. In order to access to such vinylic chloride group, different strategies were developed among which the one presented by Ryabukhin et al. [4] and the use of trichloromethylated heterocycles, as substrates for the radical SRN1+ERC1 reaction with nitroalkanes, successful approach in isoquinoline, imidazole and quinoline series [5,6,7]. We describe herein, the use of the consecutive SRN1+ERC1 reaction from 2-trichloromethylquinazoline substrates and nitroalkanes, for the preparation of these original vinylic chloride derivatives in a rapid and convenient way.

2. Results and Discussion

Since 2009, it is known that 2-chloromethyl-6-nitroquinazoline derivatives can react with nitroalkanes through a SRN1 mechanism [8]. However, the SRN1 reaction involving a radical anion intermediate which is generally stabilized with the presence of a nitro group properly located on the heterocycle, it was not obvious that an aminated quinazoline substrate would react, because of the electrodonating effect of this group.
First, the key-precursor 1 was prepared from commercial 2-methylquinazolin-4(3H)-one via a microwave-assisted chlorination procedure using phosphorus pentachloride in phosphorus oxychloride [9]. Then, after proceeding to a solvent free SNAr reaction with n-propyl- and n-butylamines, the corresponding intermediates 2 and 3 were nitrated in classical conditions to afford 4 and 5 in good yields (Scheme 1).
Scheme 1. Two step preparation of the radical reaction substrates 4 and 5.
Scheme 1. Two step preparation of the radical reaction substrates 4 and 5.
Molecules 15 02719 g001
Then, compound 4, which is prepared in higher global yield, was chosen as a substrate for studying the SRN1+ERC1 reaction with 2-nitropropane. The first reaction assay, at room temperature (RT) under N2 atmosphere and light irradiation, enabled us to isolate both vinylic chloride 6 and the dimer of 2-nitropropane 7 (Scheme 2).
Scheme 2. General presentation of the SRN1+ERC1 reaction of substrate 4 with 2-nitropropane.
Scheme 2. General presentation of the SRN1+ERC1 reaction of substrate 4 with 2-nitropropane.
Molecules 15 02719 g002
In order to optimize the reaction, several experimental conditions were studied, especially the ones described by Kornblum et al. [10] (lithium salt of 2-nitropropane in DMSO or DMF) and Norris et al. [11] (phase transfer conditions using TBAOH and 2-nitropropane). The results showed that the phase transfer conditions of Norris gave the best reaction yield. The amount of 2-nitropropane required was also studied, indicating that the reaction yield is globally optimal (82%) when using 5 equiv. (Table 1, entries 1–6). Reaction progress was monitored by TLC, and appeared to be complete after 4 hours.
The formation of dimer 7 clearly suggested a consecutive SRN1+ERC1 reaction mechanism. To validate this hypothesis, we studied the influence of several radical reaction inhibitors (Table 1 entries 7–10). The results showed that the reaction obeys to a radical pathway, in particular when comparing run 3 with runs 8 or 9. The synthesis of vinylic chloride 6 in low yield, when using 1 equiv. of p-dinitrobenzene (assay 10), may be due to the insufficient difference in reduction potential between p-dinitrobenzene and substrate 4, partially permitting the monoelectronic reduction of this last by the 2-nitropropane anion, followed by the formation of product 6, through complete SRN1+ERC1 reaction sequence.
Table 1. Reactivity of substrate 4 with 2-nitropropane.
Table 1. Reactivity of substrate 4 with 2-nitropropane.
Run*M+Solvent2-Nitropropane (equiv.)InhibitorYield of 6 (%)
1 NBu4CH2Cl2H2O 2 - 60
2 NBu4CH2Cl2H2O 4 - 75
3 NBu4CH2Cl2H2O 5 - 82
4 NBu4CH2Cl2H2O 6 - 83
5Li DMF 5 - 53
6LiDMSO 5 - 70
7LiDMSO 5 CuCl2 (1 equiv.) 25
8 NBu4CH2Cl2H2O 5 TEMPO (1 equiv.)0
9 NBu4CH2Cl2H2O 5 O2 bubbling0
10 NBu4CH2Cl2H2O 5p-dinitrobenzene 35
*Reaction conditions: N2 atmosphere, 60W white light irradiation, RT. Reactions were monitored by TLC.
From the optimal reaction conditions previously presented, we then extended the reaction to other nitroalkanes, in order to prepare several vinylic chlorides in quinazoline series (Scheme 3). So the same radical reaction was conducted with two other secondary nitroalkanes (nitrocycloalkanes) and three primary nitroalkanes. In the reactions using secondary nitroalkanes, the corresponding vinylic chlorides were obtained in identical yields (60–82%) as with nitroalkane dimers, whereas in the reactions involving primary nitroalkanes, only the corresponding vinylic chlorides were obtained, in significantly lower yields (50–56%), as presented in Table 2. Concerning the vinylic chlorides prepared from primary nitroalkanes, the analysis of 2D Noesy NMR spectra revealed a Z configuration for the vinylic double bond.
Such results appeared identical to the ones obtained with the SRN1+ERC1 reaction operated in 2-trichloromethylquinoline series [7], but demonstrate that this reaction can be compatible with the presence of an electrodonating group on the heterocyclic scaffold.
Scheme 3. Optimal reaction conditions for the synthesis of vinylic chloride derivatives from 4.
Scheme 3. Optimal reaction conditions for the synthesis of vinylic chloride derivatives from 4.
Molecules 15 02719 g003
Table 2. Series of vinylic chloride derivatives prepared from 4, using optimal reaction conditions.
Table 2. Series of vinylic chloride derivatives prepared from 4, using optimal reaction conditions.
NitroalkaneVinylic chlorideDimerYield of vinylic chloride (%)
Molecules 15 02719 i001 Molecules 15 02719 i002 Molecules 15 02719 i00382
Molecules 15 02719 i004 Molecules 15 02719 i005 Molecules 15 02719 i00660
Molecules 15 02719 i007 Molecules 15 02719 i008 Molecules 15 02719 i00966
Molecules 15 02719 i010 Molecules 15 02719 i011Not isolated56
Molecules 15 02719 i012 Molecules 15 02719 i013Not isolated50
Molecules 15 02719 i014 Molecules 15 02719 i015Not isolated52
To enlarge the reaction spectrum, we involved two other aminated substrates in the same reaction with 2-nitropropane, using the previously optimal reaction conditions. Substrate 5, containing an other primary amine substituent, was prepared in parallel with 4 and afforded vinylic chloride 15 in 57 % yield (Scheme 4).
Scheme 4. Extension of the reaction with substrate 5.
Scheme 4. Extension of the reaction with substrate 5.
Molecules 15 02719 g004
Aiming at preparing a vinylic chloride product bearing an anilino substituent at position 4 of quinazoline, we had to prepare an appropriate substrate candidate for the SRN1+ERC1 reaction. Thus, since the nitration reaction could not be done after the introduction of the aniline onto the quinazoline ring, the synthetic strategy from 2-methylquinazolin-4(3H)-one was modified, starting by a nitration step, under classical conditions. After successive chlorination and solvent free SNAr reactions, 18 was obtained (Scheme 5).
Scheme 5. Three step preparation of substrate 18.
Scheme 5. Three step preparation of substrate 18.
Molecules 15 02719 g005
So we applied the optimal reaction conditions defined previously to the reaction of 18 with 2-nitropropane and obtained the expected vinylic chloride 19 in 45% yield (Scheme 6).
Scheme 6. SRN1+ERC1 reaction of 18 with 2-nitropropane.
Scheme 6. SRN1+ERC1 reaction of 18 with 2-nitropropane.
Molecules 15 02719 g006
When 2-methylquinazolin-4(3H)-one is first chlorinated and then nitrated, compound 20 is formed in 72% global yield (Scheme 7).
Scheme 7. Two step preparation of substrate 20.
Scheme 7. Two step preparation of substrate 20.
Molecules 15 02719 g007
Finally, because of its trichloromethyl group, we thought that such nitrated lactam ring 20 could be an interesting substrate for the SRN1+ERC1 reaction. Unfortunately, its low solubility in dichloromethane did not allow the use of the optimal phase transfer reaction conditions that we already presented. So we reacted 20 with the lithium salt of 2-nitropropane in DMSO (Kornblum conditions [10]) during 2 hours and formed the vinylic chloride product 21 in 67 % yield, in addition to the dimer of 2-nitropropane 7, in similar yield (Scheme 8).
Scheme 8. SRN1+ERC1 reaction of 20 with the lithium salt of 2-nitropropane.
Scheme 8. SRN1+ERC1 reaction of 20 with the lithium salt of 2-nitropropane.
Molecules 15 02719 g008

3. Experimental

3.1. General

Melting points were determined with a B-540 Büchi melting point apparatus. 200 MHz 1H-NMR and 50 MHz 13C-NMR spectra were recorded on a Brüker ARX 200 spectrometer in CDCl3 or DMSO-d6 at the Faculté de Pharmacie de Marseille. 1H- and 13C-NMR chemical shifts (δ) are reported in ppm with respect to CDCl3 7.26 ppm (1H), 77.2 ppm (13C) or DMSO-d6 2.50 ppm (1H), 39.5 (13C). 2D NOESY NMR spectra and elemental analyses were carried out at the Spectropole, Faculté des Sciences et Techniques de Saint-Jérôme. Silica gel 60 (Merck, particular size 0.063–0.200 mm, 70–230 mesh ASTM) was used for column chromatography. TLC was performed on 5 cm × 10 cm aluminium plates coated with silica gel 60F-254 (Merck) in appropriate solvent. The following substrates or subproducts were prepared and described previously: 4-chloro-2-trichloromethyl-quinazoline (1) [9,12], N-butyl-2-trichloromethylquinazolin-4-amine (3) [13], 2,3-dimethyl-2,3-dinitrobutane (7) [14], 1-nitro-1-(1-nitrocyclopentyl)cyclopentane (9) [15], 1-nitro-1-(1-nitrocyclo-hexyl)cyclohexane (11) [16], 2-methyl-6-nitroquinazolin-4(3H)one (16) [17], 4-chloro-6-nitro-2-trichloromethylquinazoline (17) [18], and 6-nitro-2-trichloromethylquinazolin-4(3H)-one (20) [18].

3.2. Preparation of N-propyl-2-trichloromethylquinazolin-4-amine (2)

Three equiv. of propylamine were added onto 1 equiv. of 4-chloro-2-trichloromethylquinazoline (1). The solvent-free reaction mixture was stirred at 30 °C for 1 min. The mixture was poured into a solution of NaOH (5%) and extracted with dichloromethane. The organic layer was washed with water three times, dried over anhydrous Na2SO4, concentrated in vacuo and recrystalized from cyclohexane to afford 2 in 95% yield.Beige solid; yield 95%, mp 136 °C. 1H-NMR (CDCl3) δ 1.03 (3H, t, J = 7.3 Hz), 1.69–1.87 (2H, m), 3.66–3.76 (2H, m), 6.06 (1H, br s), 7.49–7.57 (1H, m), 7.73–7.81 (2H, m), 7.97 (1H, d, J = 8.3 Hz). 13C-NMR (50 MHz; CDCl3) δ 11.6 (CH3), 22.5 (CH2), 43.4 (CH2), 98.4 (C), 113.8 (C), 121.2 (CH), 127.4 (CH), 129.0 (CH), 133.2 (CH), 149.1 (C), 160.7 (C), 161.5 (C). Anal. calcd for C12H12Cl3N3: C, 47.32; H, 3.97; N, 13.80. Found: C, 47.90; H, 4.24; N, 13.85.

3.3. General Procedure for the Preparation of Compounds 4 and 5

To concentrated sulfuric acid (25 mL), 1 equiv. of N-alkyl-2-trichloromethylquinazolin-4-amine (2) or 3 was added at 0 °C. Then, 5 equiv. of concentrated nitric acid were added slowly. The reaction mixture was stirred for 5 h at rt. The reaction mixture was then slowly poured into a water-ice mixture and alkalinized with sodium carbonate, making the desired products precipitate. The suspension was filtered and the precipitate was dried under reduced pressure. After purification by chromatography on a silica gel column, eluting with dichloromethane, products 4 or 5 were recrystallized from ethyl acetate and obtained, respectively, in 85 and 75% yields.
6-Nitro-N-propyl-2-trichloromethylquinazolin-4-amine (4): Yellow solid; yield 85%, mp 151 °C. 1H-NMR (CDCl3) δ 1.07 (3H, t, J = 7.3 Hz), 1.76-1.94 (2H, m), 3.74–3.84 (2H, m), 6.58 (1H, br s), 8.07 (1H, d, J = 9.1 Hz), 8.54 (1H, dd, J = 9.1 Hz and J = 2.4 Hz), 8.83 (1H, d, J = 2.4 Hz). 13C-NMR (CDCl3) δ 11.7 (CH3), 22,5 (CH2), 44.2 (CH2), 97.3 (C), 113.1 (C), 118.5 (CH), 127,2 (CH), 130.8 (CH), 145.7 (C), 152.7 (C), 161.6 (C), 164.4 (C). Anal. Calcd for C12H11Cl3N4O2: C, 41.23; H, 3.17; N, 16.03. Found: C, 41.74; H, 3.05; N, 16.45.
N-Butyl-6-nitro-2-trichloromethylquinazolin-4-amine (5): Yellow solid; yield 75%, mp 168 °C. 1H-NMR (CDCl3) δ 1.01 (3H, t, J = 7.3 Hz), 1.40–1.59 (2H, m), 1.75–1.87 (2H, m), 3.82 (2H, q, J = 12.7 Hz), 6.75 (1H, br s), 8.07 (1H, d, J = 9.2 Hz), 8.55 (1H, dd, J = 9.2 and 2.3 Hz), 8.83 (1H, d, J = 2.3 Hz). 13C-NMR (CDCl3) δ 13.8 (CH3), 20.1 (CH2), 31.1 (CH2), 41.9 (CH2), 97.3 (C), 112.9 (C), 118.0 (CH), 126.9 (CH), 131.1 (CH), 145.5 (C), 153.0 (C), 161.3 (C), 164.5 (C). Anal. Calcd for C13H13Cl3N4O2: C, 42.94; H, 3.60; N, 15.41. Found: C, 43.01; H, 3.63; N, 15.25.

3.4. General Procedure for the Synthesis of Vinylic Chlorides 6, 8, 10, 12, 13, 14, 15 and 19

Five equiv. each of both nitroalkane and 40% TBAOH were stirred at room temperature for 30 min under nitrogen atmosphere. Then, 1 equiv. of 4-substituted-6-nitro-2-trichloromethylquinazoline substrate 4, 5 or 18, solubilized in dichloromethane, was added. The mixture was rapidly stirred at rt under nitrogen atmosphere with white light irradiation (60 W) for 4 h. After disappearance of substrate (monitored by TLC), the reaction was stopped with addition of water and the product was extracted with dichloromethane. The organic layer, washed with water several times, was finally dried over anhydrous sodium sulphate and evaporated. The product was purified by chromatography on a silica gel, eluting with appropriate solvent, to give the corresponding vinylic chloride products.
2-(1-Chloro-2-methylprop-1-enyl)-6-nitro-N-propylquinazolin-4-amine (6): The eluent was petroleum ether-ethyl acetate 50:50. Yellow solid; yield 82%, mp 145 °C. 1H-NMR (CDCl3) δ 1.06 (3H, t, J = 7.3 Hz), 1.73–1.86 (2H, m), 2.05 (3H, s), 2.09 (3H, s), 3.66–3.76 (2H, m), 6.29 (1H, br s), 7.94 (1H, d, J = 9.2 Hz), 8.48 (1H, dd, J = 9.2 Hz and 2.4 Hz), 8.76 (1H, d, J = 2.4 Hz). 13C-NMR (CDCl3) δ 11.7 (CH3), 22.4 (CH3), 22.7 (CH2), 23.3 (CH3), 43.7 (CH2), 112.5 (C), 118.4 (CH), 124.5 (C), 126.6 (CH), 130.1 (CH), 137.7 (C), 144.7 (C), 153.4 (C), 160.7 (C), 164.2 (C). Anal. Calcd for C15H17ClN4O2: C, 56.16; H, 5.34; N, 17.47. Found: C, 56.75; H, 5.69; N, 17.05.
2-[Chloro(cyclopentylidene)methyl]-6-nitro-N-propylquinazolin-4-amine (8): The eluent was dichloromethane-ethyl acetate 98:2. Orange solid; yield 60%, mp 160 °C. 1H-NMR (DMSO-d6) δ 0.95 (3H, t, J = 7.3 Hz), 1.59–1.80 (6H, m), 2.59–2.65 (2H, m), 2.92–2.98 (2H, m), 3.49–3.51 (2H, m), 7.76 (1H, d, J = 9.2 Hz), 8.42 (1H, dd, J = 9.2 Hz and 2.3 Hz), 8.98–9.00 (1H, m), 9.31 (1H, d, J = 2.3 Hz). 13C-NMR (DMSO-d6) δ 11.7 (CH3), 21.8 (CH2), 25.1 (CH2), 27.8 (CH2), 35.2 (CH2), 36.6 (CH2), 43.0 (CH2), 112.3 (C), 120.6 (CH), 122.4 (C), 126.5 (CH), 129.2 (CH), 144.0 (C), 153.4 (C), 154.2 (C), 160.2 (C), 161.9 (C). HR MS (+ESI): m/z 347.1269 (M+H+). Calcd for C17H19ClN4O2: 346.1197.
2-[Chloro(cyclohexylidene)methyl]-6-nitro-N-propylquinazolin-4-amine (10): The eluent was petroleum ether-ethyl acetate 50:50. Yellow solid; yield 66%, mp 170 °C. 1H-NMR (CDCl3) δ 1.05 (3H, t, J = 7.3 Hz), 1.61–1.85 (10H, m), 2.57–2.62 (2H, m), 3.66–3.76 (2H, m), 6.37 (1H, br s), 7.93 (1H, d, J = 9.2 Hz), 8.46 (1H, dd, J = 9.2 Hz and 2.4 Hz), 8.77 (1H, d, J = 2.4 Hz). 13C-NMR (CDCl3) δ 11.7 (CH3), 22.7 (CH2), 26.4 (CH2), 27.3 (CH2), 27.9 (CH2), 30.0 (CH2), 32.4 (CH2), 43.7 (CH2), 112.4 (C), 118.4 (CH), 121.5 (C), 126.6 (CH), 130.2 (CH), 143.8 (C), 144.7 (C), 153.6 (C), 160.8 (C), 164.3 (C). HR MS (+ESI): m/z 361.1426 (M+H+). Calcd for C18H21ClN4O2: 360.1353.
(Z)-2-(1-Chloroprop-1-enyl)-6-nitro-N-propylquinazolin-4-amine (12): The eluent was dichloro-methane. Green solid; yield 56%, mp 176 °C. 1H-NMR (CDCl3) δ 1.03–1.10 (3H, m), 1.72–1.90 (2H, m), 2.08 (3H, d, J = 7.0 Hz ), 3.67–3.77 (2H, m), 6.43 (1H, br s), 7.63 (1H, q, J = 7.0 Hz), 7.92–7.98 (1H, m), 8.44 (1H, dd, J = 9.2 Hz and 2.4 Hz)), 8.78 (1H, d, J = 2.4 Hz). 13C-NMR (CDCl3) δ 11.8 (CH3), 15.8 (CH3), 22.5 (CH2), 43.9 (CH2), 112.9 (C), 118.9 (CH), 126.9 (CH), 129.4 (CH), 132.4 (C), 134.5 (CH), 144.6 (C), 152.5 (C), 160.0 (C), 160.5 (C). HR MS (+ESI): m/z 307.0956 (M+H+). Calcd for C14H15ClN4O2: 306.0884.
(Z)-2-(1-Chlorobut-1-enyl)-6-nitro-N-propylquinazolin-4-amine (13): The eluent was dichloro-methane. Yellow solid; yield 50%, mp 163 °C. 1H-NMR (CDCl3) δ 0.98–1.23 (6H, m), 1.77–1.88 (2H, m), 2.46–2.61 (2H, m), 3.68–3.78 (2H, m), 6.40 (1H, br s), 7.55 (1H, t, J = 9.3 Hz), 7.97 (1H, d, J = 9.2 Hz), 8.45 (1H, dd, J = 9.2 Hz and 2.4 Hz), 8.78 (1H, d, J = 2.4 Hz). 13C-NMR (CDCl3) δ 11.2 (CH3), 12.2 (CH3), 22.0 (CH2), 22.9 (CH2), 43.1 (CH2), 112.3 (C), 118.0 (CH), 126.1 (CH), 129.3 (CH), 130.8 (C), 139.8 (CH), 143.8 (C), 152.6 (C), 159.4 (C), 160.3 (C). Anal. Calcd for C15H17ClN4O2: C, 56.16; H, 5.34; N, 17.47. Found: C, 55.82; H, 5.75; N, 17.01.
(Z)-2-(1-Chloropent-1-enyl)-6-nitro-N-propylquinazolin-4-amine (14): The eluent was dichloromethane-ethyl acetate 99:1. Yellow solid; yield 52%, mp 156 °C. 1H-NMR (DMSO-d6) δ 0.92–0.99 (6H, m), 1.49–1.76 (4H, m), 2.36–2.44 (2H, m), 3.50–3.60 (2H, m), 7.46–7.53 (1H, m), 7.81 (1H, d, J = 9.1 Hz), 8.43 (1H, dd, J = 9.1 Hz and 2.0 Hz), 9.01 (1H, br s), 9.33 (1H, d, J = 2.0 Hz). 13C-NMR (DMSO-d6) δ 12.1 (CH3), 14.2 (CH3), 21.6 (CH2), 21.9 (CH2), 31.7 (CH2), 43.2 (CH2), 113.5 (C), 120.9 (CH), 126.9 (CH), 129.7 (CH), 132.6 (C), 137.1 (CH), 144.3 (C), 153.8 (C), 160.2 (C), 160.4 (C). HR MS (+ESI): m/z 335.1269 (M+H+). Calcd for C16H19ClN4O2: 334.1197.
N-Butyl-2-(1-chloro-2-methylprop-1-enyl)-6-nitroquinazolin-4-amine (15): The eluent was petroleum ether-ethyl acetate 50:50. Beige solid; yield 57%, mp 135 °C. 1H-NMR (CDCl3) δ 1.00 (3H, t, J = 7.2 Hz), 1.40–1.58 (2H, m), 1.72–1.84 (2H, m), 2.05 (3H, s), 2.09 (3H, s), 3.75 (2H, m), 6.42 (1H, br s), 7.96 (1H, d, J = 9.2 Hz), 8.49 (1H, dd, J = 9.2 and 2.3 Hz), 8.73 (1H, d, J = 2.3 Hz). 13C-NMR (CDCl3)δ 13.8 (CH3), 20.2 (CH2), 22.2 (CH3), 23.1 (CH3), 31.3 (CH2), 41.6 (CH2), 112.3 (C), 118.2 (CH), 124.4 (C), 126.4 (CH), 130.0 (CH), 137.4 (C), 144.5 (C), 153.3 (C), 160.5 (C), 164.0 (C). Anal. Calcd for C16H19ClN4O2: C, 57.40; H, 5.72; N, 16.73. Found: C, 57.02; H, 5.94; N, 16.43.
2-(1-Chloro-2-methylprop-1-enyl)-6-nitro-N-[3-(trifluoromethyl)]-phenylquinazolin-4-amine (19): The eluent was dichloromethane. Yellow solid, yield 45%, mp 201 °C. 1H-NMR (CDCl3) δ 2.13 (6H, s), 7.45–7.60 (2H, m), 7.98-8.11 (3H, m), 8.40 (1H, br s), 8.58 (1H, dd, J = 9.1 Hz and 2.2 Hz), 9.07 (1H, d, J = 1.7 Hz). 13C-NMR (CDCl3) δ 22.5 (CH3), 23.7 (CH3), 112.3 (C), 118.0 (CH), 118.8 (CH, q, J = 4.0 Hz), 121.6 (CH, q, J = 4.0 Hz), 123.8 (C, q, J = 272.6 Hz), 124.1 (C), 124.7 (CH), 127.0 (CH), 129.7 (CH), 130.6 (CH), 131.6 (C, q, J = 32.6 Hz), 138.0 (C), 140.1 (C), 145.2 (C), 153.3 (C), 158.2 (C), 163.1 (C). Anal. calcd for C19H14ClF3N4O2: C, 53.98; H, 3.34; N, 13.25. Found: C, 54.4; H, 3.52; N, 12.95.

3.5. Preparation of 6-nitro-2-trichloromethyl-N-[(3-trifluoromethyl)]phenylquinazolin-4-amine (18)

Three equiv. of 3-trifluoromethylaniline were added onto 1 equiv. (500 mg) of 4-chloro-6-nitro-2-trichloromethylquinazoline (17). The solvent-free reaction mixture was stirred at 120 °C for 15 min. The mixture was poured into a saturated solution of sodium carbonate and extracted with dichloromethane. The organic layer was washed with water three times, dried over anhydrous Na2SO4 and concentrated in vacuo. After purification by chromatography (silica gel, eluent: dichloromethane) the product was obtained in 74% yield. Yellow solid; yield 74%, mp 234 °C. 1H-NMR (DMSO-d6) δ 7.56 (1H, d, J = 7.7 Hz), 7.67–7.75 (1H, m), 8.14 (1H, d, J = 9.2 Hz), 8.22 (1H, d, J = 8.3 Hz), 8.65–8.71 (2H, m), 9.80 (1H, d, J = 2.4 Hz), 11.05 (1H, br s). 13C-NMR (DMSO-d6) δ 97.2 (C), 113.7 (C), 118.7 (CH, q, J = 4.0 Hz), 120.8 (CH), 120.9 (CH), 124.1 (CF3, q, J = 272.7 Hz), 125.5 (CH), 127.8 (CH), 129.3 (C, q, J = 31.9 Hz), 129.8 (CH), 130.4 (CH), 138.9 (C), 146.0 (C), 152.5 (C), 159.5 (C), 162.7 (C). Anal. calcd for C16H8Cl3F3N4O2: C, 42.55; H, 1.79; N, 12.41. Found: C, 42.50; H, 1.91; N, 11.97.

3.6. Preparation of 2-(1-chloro-2-methylprop-1-enyl)-6-nitroquinazolin-4(3H)-one (21)

To a DMSO solution (15 mL) of the lithium salt of 2-nitropropane (0.76 g, 8.1 mmol, 5 equiv.), a DMSO solution (15 mL) containing 0.5 g (1.6 mmol, 1 equiv.) of 20 was added rapidly. The mixture was stirred at rt for 2 h under N2 atmosphere and light irradiation (60W). After the disappearance of 20 (monitored by TLC), the mixture was poured into cold water, filtered, and extracted with dichloromethane. The combined organic layer were dried over anhydrous Na2SO4 and evaporated. The solid product was washed with petroleum ether and dried under reduced pressure to give 21 in 67% yield. Yellow solid; yield 67%, mp 248 °C. 1H-NMR (CDCl3) δ 2.17 (3H, s), 2.30 (3H, s), 7.84 (1H, d, J = 9.0 Hz), 8.56 (1H, dd, J = 9.0 Hz and 2.4 Hz), 8.76 (1H, d, J = 2.4 Hz), 9.74 (1H, br s). 13C-NMR (CDCl3) δ 23.2 (CH3), 24.3 (CH3), 117.0 (C), 121.3 (C), 123.1 (CH), 128.9 (CH), 129.4 (CH), 146.2 (C), 151.7 (C), 152.3 (C), 160.4 (C), (C-NO2) was not observed in these experimental conditions. Anal. Calcd for C12H10ClN3O3: C, 51.53; H, 3.60; N, 15.02. Found: C, 50.99; H, 3.53; N, 14.94.

4. Conclusions

Quinazoline derivatives are molecules with a high pharmaceutical potential, particularly when targeting kinases. Using the consecutive SRN1+ERC1 radical reaction strategy, a series of original 4-propylaminoquinazolines, bearing a vinylic chloride group at position 2, was prepared in a simple and convenient way from corresponding trichloromethylated substrates. The reaction parameters were studied and optimized by using a phase-transfer procedure involving 5 equiv. of both TBAOH and 2-nitropropane, and the radical mechanism was investigated by employing radical reaction inhibitors such as TEMPO. Finally, the reaction was extended in 4-butylaminoquinazoline, 4-anilinoquinazoline and quinazoline-4(3H)-one series, demonstrating that the SRN1+ERC1 reaction operated onto nitrated quinazolines is compatible with the presence of an alkylamino electrodonating group, so as with a lactam function.

Acknowledgements

This work was supported by the CNRS and the Université de la Méditerranée. The authors thank V. Remusat for recording NMR spectra.
  • Samples Availability: Samples of the compounds are available from the authors.

References

  1. Verhaeghe, P.; Azas, N.; Gasquet, M.; Hutter, S.; Ducros, C.; Laget, M.; Rault, S.; Rathelot, P.; Vanelle, P. Synthesis and antiplasmodial activity of new 4-aryl-2-trichloromethylquinazolines. Bioorg. Med. Chem. Lett. 2008, 18, 396–401. [Google Scholar]
  2. Verhaeghe, P.; Azas, N.; Hutter, S.; Castera-Ducros, C.; Laget, M.; Dumètre, A.; Gasquet, M.; Reboul, J.P.; Rault, S.; Rathelot, P.; Vanelle, P. Synthesis and in vitro antiplasmodial evaluation of 4-anilino-2-trichloromethylquinazolines. Bioorg. Med. Chem. 2009, 17, 4313–4322. [Google Scholar] [CrossRef]
  3. Kabri, Y.; Azas, N.; Dumetre, A.; Hutter, S.; Laget, M.; Verhaeghe, P.; Gellis, A.; Vanelle, P. Original quinazoline derivatives displaying antiplasmodial properties. Eur. J. Med. Chem. 2010, 45, 616–622. [Google Scholar] [CrossRef]
  4. Ryabukhin, S.V.; Plaskon, A.S.; Volochnyuk, D.M.; Tolmachev, A.A. Chlorotrimethylsilane-mediated synthesis of 2-aryl-1-chloro-1-heteroarylalkenes. Synthesis 2007, 20, 3163–3170. [Google Scholar]
  5. Vanelle, P.; Rathelot, P.; Maldonado, J.; Crozet, M.P. Consecutive SRN1 and ERC1 reactions in 5-nitroisoquinolines. Tetrahedron Lett. 1994, 35, 8385–8388. [Google Scholar] [CrossRef]
  6. Vanelle, P.; Benakli, K.; Giraud, L.; Crozet, M.P. Synthesis of tetrasubstituted ethylenic compounds from a gem-trichloro imidazole derivative by electron-transfer reactions. Synlett 1999, 6, 801–803. [Google Scholar]
  7. Verhaeghe, P.; Rathelot, P.; Rault, S.; Vanelle, P. Convenient preparation of original vinylic chlorides with antiparasitic potential in quinoline series. Lett. Org. Chem. 2006, 3, 891–897. [Google Scholar] [CrossRef]
  8. Kabri, Y.; Gellis, A.; Vanelle, P. Microwave-assisted synthesis in aqueous medium of new quinazoline derivatives as anticancer precursors. Green Chem. 2009, 11, 201–208. [Google Scholar] [CrossRef]
  9. Verhaeghe, P.; Rathelot, P.; Gellis, A.; Rault, S.; Vanelle, P. Highly efficient microwave assisted α-trichlorination reaction of α-methylated nitrogen containing heterocycles. Tetrahedron 2006, 62, 8173–8176. [Google Scholar] [CrossRef]
  10. Kerber, R.C.; Urry, G.W.; Kornblum, N. Radical anions as intermediates in substitution reactions. Carbon alkylation of nitroparaffin salts. J. Am. Chem. Soc. 1964, 86, 3904. [Google Scholar] [CrossRef]
  11. Burt, B.; Freeman, D.J.; Gray, P.G.; Norris, R.K.; Randles, D. Radical and ionic reactions of tetrabutylammonium aci-nitronates. Tetrahedron Lett. 1977, 35, 3063–3066. [Google Scholar]
  12. Dehoff, L.H. Nitro- and chloro-derivatives of β-methyl-γ-oxyquinazoline (anhydroacetyl-orthamidobenzamide). J. Prakt. Chem. 1890, 42, 346–360. [Google Scholar] [CrossRef]
  13. Wan, Z.K.; Wacharasindhu, S.; Levins, C.G.; Lin, M.; Tabei, K.; Mansour, T.S. The scope and mechanism of phosphonium-mediated SNAr reactions in heterocyclic amides and ureas. J. Org. Chem. 2007, 72, 10194–10210. [Google Scholar] [CrossRef]
  14. Girdler, D.J.; Norris, R.K. Unusual reaction of benzylidine diacetates. Ionic chain mechanism for the substitution of an acetate group by the 2-nitro-2-propanide ion. Tetrahedron lett. 1975, 7, 431–434. [Google Scholar]
  15. Ono, N.; Miyake, H.; Tamura, R.; Hamamoto, I.; Kaji, A. Free radical chain elimination reaction (ERC1). Conversion of vicinal dinitro compounds or β-nitro sulfones to olefins with tributyltin hybride. Chem. Lett. 1981, 10, 1139–1142. [Google Scholar]
  16. Kornblum, N.; Boyd, S.D.; Pinnick, H.W.; Smith, R.G. New synthesis of olefins. J. Am. Chem. Soc. 1971, 93, 4316–4318. [Google Scholar]
  17. Bogert, M.T.; Geiger, G.A. Researches on quinazolines (thirtieth paper). Study of the bromination and nitration of 4-quinazolines; The corresponding aminoquinazolines, and certain other new 4-quinazolines. J. Am. Chem. Soc. 1912, 34, 524–534. [Google Scholar]
  18. Hepworth, W. Quinazoline derivatives. Great Britain Patent 857362, 1960. [Chem. Abstr. 1961, 55, 76248]. [Google Scholar]

Share and Cite

MDPI and ACS Style

Maillard-Boyer, M.; Castera-Ducros, C.; Verhaeghe, P.; Sifredi, F.; Rathelot, P.; Vanelle, P. Access to Original Vinylic Chlorides in the Quinazoline Series via a Monoelectronic Transfer Reaction Approach. Molecules 2010, 15, 2719-2729. https://doi.org/10.3390/molecules15042719

AMA Style

Maillard-Boyer M, Castera-Ducros C, Verhaeghe P, Sifredi F, Rathelot P, Vanelle P. Access to Original Vinylic Chlorides in the Quinazoline Series via a Monoelectronic Transfer Reaction Approach. Molecules. 2010; 15(4):2719-2729. https://doi.org/10.3390/molecules15042719

Chicago/Turabian Style

Maillard-Boyer, Martine, Caroline Castera-Ducros, Pierre Verhaeghe, France Sifredi, Pascal Rathelot, and Patrice Vanelle. 2010. "Access to Original Vinylic Chlorides in the Quinazoline Series via a Monoelectronic Transfer Reaction Approach" Molecules 15, no. 4: 2719-2729. https://doi.org/10.3390/molecules15042719

Article Metrics

Back to TopTop