Next Article in Journal
Fabrication and Characterization of Chitosan–Pea Protein Isolate Nanoparticles
Next Article in Special Issue
Catalyst-Free Formal Conjugate Addition/Aldol or Mannich Multicomponent Reactions of Mixed Aliphatic Organozinc Reagents, π-Electrophiles and Michael Acceptors
Previous Article in Journal
Cryogel-Templated Fabrication of n-Al/PVDF Superhydrophobic Energetic Films with Exceptional Underwater Ignition Performance
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Triflamidation of Allyl-Containing Substances:Unusual Dehydrobromination vs. Intramolecular Heterocyclization

A.E. Favorsky Irkutsk Institute of Chemistry, Siberian Division of the Russian Academy of Sciences, 1 Favorsky Street, 664033 Irkutsk, Russia
*
Author to whom correspondence should be addressed.
Molecules 2022, 27(20), 6910; https://doi.org/10.3390/molecules27206910
Submission received: 15 September 2022 / Revised: 4 October 2022 / Accepted: 12 October 2022 / Published: 14 October 2022
(This article belongs to the Special Issue Multicomponent Reactions in Organic Synthesis)

Abstract

:
Allyl halides with triflamide under oxidative conditions form halogen-substituted amidines. Allyl cyanide reacts with triflamide in acetonitrile or THF solutions in the presence of NBS to give the products of bromotriflamidation with a solvent interception, whereas in CH2Cl2 two regioisomers of the bromotriflamidation product without a solvent interception were obtained. The formed products undergo base-induced dehydrobromination to give linear isomers with the new C=C bond conjugated either with the nitrile group or the amidine moiety or alkoxy group. Under the same conditions, the reaction of allyl alcohol with triflamide gives rise to amidine, which was prepared earlier by the reaction of diallyl formal with triflamide. Unlike their iodo-substituted analogs, bromo-substituted amidines successfully transform into imidazolidines under the action of potassium carbonate.

1. Introduction

Oxidative sulfonamidation of unsaturated compounds is a convenient method for the formation of the C–N bond and an expedient route to the synthesis of various linear and cyclic compounds capable of further functionalization. The course of the reaction and the structure of products strongly depend on the reagent, oxidant, and reaction conditions [1,2,3,4].
Allylic substrates differ from their vinylic analogs in the possibility of migration of the double bond upon nucleophilic or electrophilic attack of the terminal olefinic carbon atom, which is impossible in vinylic substrates. In the literature, there are not so many examples of the reactions of oxidative sulfonamidation with the participation of allyl-containing substrates. Thus, in the presence of mild oxidant Cu(OAc)2 and Cs2CO3 as a base, N-arylsulfonyl-ortho-allylanilines undergo oxidative cyclization to afford the products with four fused rings [5]. Homoallylic aromatic sulfonamides ArSO2NHCH(R)CH2CH=CH2 are intramolecularly oxidized by PhI(OAc)2 in the presence of KBr with cyclization to 4-bromopyrrolidines to give a mixture of the cis and trans isomers in high yield [6]. N-Bromosuccininide (NBS) induced enantioselective cyclization of allyl-N-tosylcarbamates catalyzed by a complex of Sc(OTf)3 with chiral phosphine was reported [7]; the yield of the target products, substituted oxazolidinones, reached 71–90%. The latter was easily recyclized to the oxymethyl-substituted aziridines (Scheme 1).
Various 6-halomethyl-substituted 1-tosylpiperazin-2-ones were obtained by NBS-induced intramolecular cyclization of N-allyl-N-benzyl-2-(tosylamido)acetamide [8], PdCl2(MeCN)2-catalyzed cyclization of tosylglycine-N-allylamides with CuCl2 in THF [9], or by the combined action of N-chlorosuccinimide (NCS) and PdCl2(PhCN)2 [10]. The replacement of PdCl2(MeCN)2 by PdCl2(PhCN)2 allowed to increase the yield to 90% as compared to 65% in [9] (Scheme 2).
Intramolecular bromoamination of O-allyl-N-hydroxytosylamides via 5-endo-tet-cyclization with bromoacetamide proceeds trans-diastereoselectively leading to isoxazolidines in good yield and opening a way to aminoalcohols and aziridines as useful building blocks [11] (Scheme 3).
Earlier, we studied a lot of unsaturated substrates in the reactions of oxidative sulfonamidation as summarized in review [12] but only a few of them contained electron-withdrawing groups. Among them were divinyl sulfone and divinyl sulfoxide, which reacted with triflamide in the system t-BuOCl/NaI via iodotriflamidation with subsequent cyclization into 2,6-diiodo-4-(triflyl)thiomorpholine 1,1-dioxide [13], mono- and diallyltriflamides, which under the same conditions reacted with carboxamides and sulfonamides via halogenation of the double bond [14,15] and/or iodosulfonamidation and cyclization to 3,7-diiodo-1,5-bis(triflyl)-1,5-diazocane and 3,7,9-tris(triflyl)-3,7,9- triazabicyclo[3.3.1]nonane [15], and mono- and diallyl ethers and allyl acetate, which on cooling to −30 °C gave the products of triflamidation or cyclization [16].

2. Results and Discussion

With this in mind, in the present work, we have studied the reactions of allyl halides, allyl alcohol, allylamine, acrylonitrile, and allyl cyanide with triflamide under oxidative conditions in different solvents.
The reaction of triflamide (1) with allyl chloride (2) and allyl bromide (3) in the presence of N-bromosuccinimide (NBS) and acetonitrile at room temperature affords the products of halogenation with a solvent interception, N-(2-bromo-3-halopropyl)-N’- (trifluoromethylsulfonyl)acetamidamides (4, 5) (Scheme 4).
Analytically pure compounds were isolated by column chromatography. The structure of compounds 4 and 5 was proved by NMR and IR spectroscopy, as well as elemental analysis data. In particular, the IR spectrum of 4 contains absorption bands at 3334 (νNH), 1556 (νC=N), and 663 cm–1C–Br). The 1H NMR spectrum shows a broadened singlet of the NH group, a triplet of triplets of the CHBr proton, and a singlet at 2.5 ppm, typical for the methyl group in the amidine fragment. The 13C NMR spectrum displays the signal of the azomethine group C=N and a quartet of the CF3 group. Note, that no products of the addition of the triflamide residue to the double bond were observed.
The reaction of allyl iodide 6 with triflamide under the same conditions gave amidine 5 identical to that obtained in the reaction of allyl bromide 3 (Scheme 4). The product does not contain iodine, which is apparently indicative of its substitution in the intermediate bromoiodo derivative 7 by bromine from NBS (Scheme 5).
A possible explanation of the formation of the dibromo-substituted amidine 5 from allyl iodide 6 is given in Scheme 6, suggesting the bromine/iodine exchange in the intermediate 7.
Replacing NBS with N-iodosuccinimide (NIS) in the reaction of triflamide with allyl halides 2 and 3, N-(2-iodo-3-halopropyl)-N’-(trifluoromethylsulfonyl)acetamidamides 8 and 9 were obtained (Scheme 7). The low yields in the reaction using NIS can be due to the lower Lewis acidity of the generated iodine cation than that of the bromine cation.
The structure of products 8 and 9 was proved by NMR and IR spectroscopy, as well as elemental analysis data. The IR spectra of both products show two νNH absorption bands at 3326 and 3231 cm–1 and the bands at 1577, 1553 cm–1C=N). In the 1H NMR spectrum of 8, a broad singlet of the NH group and a doublet of doublets of the CHI proton appears. The CHI signals in 8 strongly differ from that in 9 in the position and the character of splitting (ddd in 8 and a multiplet in 9).
Surprisingly, no reaction occurred between allyl iodide 6 and triflamide in the presence of NIS: the reagents were recovered unchanged.
No products could be isolated from the NBS-induced reaction of allyl amine 10 with triflamide because of the strong polymerization of the reaction mixture. In contrast, allyl alcohol 11 afforded a low yield of amidine 12, which was obtained earlier from the NBS-induced reaction of diallylformal with triflamide (Scheme 8) [16].
With acrylonitrile, neither in the system t-BuOCl/NaI nor in the presence of NBS, at room temperature or on cooling, any products were isolated, apparently, due to strong polymerization under oxidative conditions (see, e.g., [17]). As distinct from that, the NBS-induced reaction of triflamide with allyl cyanide 13 in acetonitrile gave the product of bromotriflamidation with solvent interception 14 similar to the reactions of other substrates under analogous conditions [16,18]. The yield of N-(2-bromo-3-cyanopropyl)- N’-(trifluoromethylsulfonyl)ethaneimidamide 14 isolated by column chromatography was 60%. Its structure was proved by the methods of IR, NMR spectroscopy, and HRMS. In particular, the IR spectrum of amidine 14 shows absorption bands νNH (3324), νC≡N (2259), and νNHC=N (1560 cm–1), its 1H NMR spectrum displays a broad NH singlet, the signals of diastereotopic CH2N protons and a singlet of the methyl group at the azomethine bond. The 13C NMR spectrum contains the C=N and C≡N signals, the CF3 quartet, and the corresponding signal appears in the 19F NMR spectrum. The use of larger amounts of the reagents allowed to isolate the minor product, N-(2-bromo-3-cyanopropyl)triflamide 15 having no acetonitrile moiety (Scheme 9). Its structure was also proved by NMR and IR spectroscopy. The ratio of compounds 14:15, from 1H NMR spectroscopy, was ~4:1 (Scheme 9).
By replacing acetonitrile with THF as a solvent, we hoped to synthesize amino esters, as was previously successfully completed in our works [18,19]. However, with allyl chloride, instead, the product of bromination, 1,2-dibromo-3-chloropropane 16, was isolated in a low yield (Scheme 10) indicating that triflamide is not involved in the reaction.
The reason for this behavior is that triflamide practically does not react with unsaturated substrates in solvents of low basicity [20].
Carrying out the reaction of allyl cyanide 13 in Scheme 9 in THF instead of MeCN also led to the solvent interception product, N-[4-(2-bromo-3-cyanopropoxy)butyl]- triflamide 17 formed via the THF ring opening and its addition as an O-nucleophile (Scheme 11).
Excluding the possibility of the formation of amidine 14 by replacing acetonitrile with methylene chloride, we obtained two regioisomers of the product of bromotriflamidation 18 and 19, isolated them as individual compounds and proved their structure and composition by IR, NMR spectroscopy and elemental analysis. 3-Bromo-4-hydroxybutanenitrile 20 was also obtained in a comparable yield (Scheme 12). The prevalence of bromination over bromotriflamidation is probably due to the low solubility of triflamide in methylene chloride.
For comparison, the reaction of allyl cyanide 13 with tosylamide was examined under the same conditions. However, no products of sulfonamidation were obtained, but only dibromide 19 and unreacted tosylamide were recovered.
Amidines 4 and 5 were examined in the reaction with K2CO3 in acetonitrile. As a result of intramolecular cyclization, substituted 4,5-dihydro-1H-imidazoles 21, 22 were obtained in quantitative yield. However, upon prolonged exposure to humid air, the bromo-substituted imidazoline 22 hydrolyzed to linear adduct 23 (Scheme 13):
The structure of imidazolines 21, 22 was proved by IR and NMR spectroscopy, as well as elemental analysis data. The presence of two NH signals in the 1H NMR spectrum, as well as the presence of signals for CH2NH, CHNH and C=O groups in the 13C spectrum, indicates the formation of adduct 23.
Amidines 8 and 9 having two halogen atoms could give the products of cyclization with different ring sizes, but neither of them was formed; no reaction with K2CO3 occurred.
Amidines similar to 14 containing bromine at the β-position to the amidine nitrogen atom readily undergo base-induced intramolecular cyclization to afford 5-substituted 2-methyl-1-triflyl-4,5-dihydro-1H-imidazolines in up to quantitative yield [16,19,20,21]. With this in mind, we examined the reaction of amidine 14 with potassium carbonate and triethylamine in acetonitrile and found that dehydrobromination did occur but, instead of the expected 5-cyanomethylimidazoline 24, N-[3-cyanoprop-2-en-1-yl)]-N’- (triflyl)ethaneimidamide 25 was unexpectedly formed. Even more surprising was the formation of the isomeric N-[3-cyanoprop-1-en-1-yl)]-N’-(triflyl)ethanemidamide 26 in carrying out the two-step reaction of triflamide, alkene, NBS and K2CO3 using one pot procedure, which was also shown to lead to imidazolines [20], (Scheme 14). Replacement of triethylamine or K2CO3 by sterically hindered 2,4,6-trimethylpyridine (2,4,6-collidine) does not change the course of the reaction under the same conditions, leading to the formation of amidine 25 as the only isomer in 76% yield.
The structure of isomers 25 and 26 was deduced from their 1H NMR spectra, in particular, from the multiplicity pattern of the high-field signal of the methylene group. In isomer 25, the signal of –CH2N– group appears as a triplet of doublets at 4.22 ppm due to splitting on the NH and =CH protons with almost equal constants of ~6 Hz, and subsplitting with small constant of 1.5 Hz on the CHC≡N proton. In accordance with this, the CHC≡N signal at 5.65 ppm is detected as a doublet of triplets with coupling constants of 11.2 and 1.5 Hz, and the CH=CHCH2 signal at 6.49 as a doublet of triplets with the J values of 11.2 and 6.2 Hz. The structure of 25 is unequivocally proved by the 2D 1H–1H COSY NMR spectrum, which contains cross-peaks between the CH2 and NH signals, as well as between the CH2 and the signals of the adjacent (more intense) and remote (less intense) vinylic protons (Supplementary Materials Figure S30). The C=C bond is polarized towards the cyano group, Δδ = 0.84 ppm. In contrast, in isomer 26, the signal of the –CH2N– group at 3.28 ppm appears as a doublet of doublets coupled only with the adjacent and remote vinylic protons with J = 7.3 and 1.2 Hz, respectively. Both compounds have trans-configuration about the double bond. Polarization of the C=C bond in 26 (Δδ = 1.90 ppm) is much larger than in 25, in compliance with the oppositely directed effects of the CN and –CH2N groups in 25, and the unidirectional effect of the NCCH2 and NH groups in 26.
Molecules 27 06910 i001
Earlier, the products of oxidative sulfonamidation with THF interception have been shown to undergo base-induced intramolecular heterocyclization to the corresponding 1,4-oxazocanes [19]. However, as in Scheme 15, the reaction of compound 17 with potassium carbonate, instead of cyclization, occurred as dehydrobromination to the isomeric linear products, N-(4-((3-cyanoallyl)oxy)butyl)triflamide 27 and N-(4-((3-cyanoprop-1-en-1-yl)oxy)butyl)triflamide 28 in the ratio of 1:2 (Scheme 15) and the total yield of 80%.
The formation of two regioisomers 27 and 28 by dehydrobromination of ether 17 as distinct from the reaction of amidine 14 (Scheme 14) can be due to better conjugation of the C=C bond with the oxygen atom than with the amine nitrogen atom in amidine 26 because of very strong conjugation of the latter in the amidine fragment [22]. The structure of regioisomers 27 and 28 was proved by their 1H NMR spectra as described above for regioisomers 25 and 26.
The proposed pathways for the formation of products 14, 15, and 18 are presented in Scheme 16. The process could start with the reaction of TfNH2 and NBS leading to the reactive species TfNHBr, which acts as a source of electrophilic Br+. The latter adds to the double bond of the substrate to give bromonium cation. The further course of the reaction is determined by the reaction medium. In acetonitrile, having higher basicity than triflamide (780 [23] vs. 740 kJ/mol [24]), the molecule of MeCN is captured by the cation with further addition of the triflamide anion to give amidine 14. A competitive attack of triflamide anion gives rise to a small amount of adduct 15 (Scheme 16). In CH2Cl2, in the absence of an alternative nucleophile, only the isomeric bromamines 15 and 18 are formed via the attack of TfNH¯ on the terminal and internal carbon, respectively, in the intermediate bromonium ion. The formation of dibromide 19 and bromoalcohol 20 (Scheme 12) can proceed either by the replacement of the triflamide residue in 15 by bromine or hydroxy group or via the ring opening in the bromonium ion by the terminal attack with these groups.
The most challenging question is why the reaction of dehydrobromination of compound 14 in Scheme 15 results in the formation of isomeric linear products 25 and 26, being drastically different from all earlier studied reactions of similar β-bromoamidines with bases leading to cyclization to imidazolines. The formation of imidazolines in all our previous works is not surprising because of the higher energy of the bonds of different types (C–C, C–N, and C–H in imidazolines vs. C=C and N–H in linear products of dehydrogenation). In the search for a rationale for the specific behavior of compound 14, we assumed that there could be two reasons for the formation of linear products 25 and 26: (i) conjugation of the formed C=C bond with the nitrile group in 25 or with the NH group in 26, and (ii) the presence of acidic NH proton in the amidine motif of 25 and 26, capable of associating with the basic sites of the second molecule. For this, we performed high-level MP2/6-311++G(d,p) calculations including frequency analysis of molecules 25, 26, their dimers, and the isomeric imidazoline 24 shown in Scheme 15. The relative energies and free energies are given in Table 1. Remarkably, isomers 25 and 26 form different types of associates: while for 26 it is a 12-membered cyclic dimer with two N–H∙∙∙O=S hydrogen bonds, for the similar dimer of compound 25 the geometry optimization results in its transformation to the eight-membered dimer with two N–H∙∙∙N hydrogen bonds (Figure 1).
The analysis of the data of Table 1 allowed us to explain two apparent inconsistencies with the experiment. First, the ΔE and ΔG differences of 3–4 kcal/mol between the monomers 25 and 26 seem to contradict the formation of both isomers. However, for the dimers, the corresponding differences in ΔE become equal. In spite of different types of H-bonding, the entropy losses upon the formation of dimers in Figure 1 and the ΔG values are also equal. Apparently, the lowering of the energy of 25-dimer is due to higher basicity of the azomethine nitrogen caused by strong conjugation in the NH–C=N tryad, whereas in 26-dimer this effect is reduced by the rivalry with that in the NH–C=C fragment. Second, while the monomers of amidines 25 and 26 are far less favorable than imidazoline 24, the dimers are much closer in energy and free energy to this heterocycle. Calculations of higher associates at the used very high level of theory are practically impossible, but the presence of acidic NH protons in monomeric molecules 25 and 26 allows them to be formed. This will certainly further increase the stability of the associates and make it highly probable the reversal of the relative stability with respect to imidazoline 24.

3. Materials and Methods

3.1. General Details

All starting materials have been described in the literature. All products were identified using IR, 1H, 13C, and 19F NMR spectroscopy. IR spectra were taken on a Bruker Vertex 70 spectrophotometer in KBr. 1H, 13C, and 19F NMR spectra were recorded in CDCl3 or CD3CN on Bruker DPX 400 spectrometer at working frequencies 400 (1H), 100 (13C), and 376 (19F) MHz. All shifts are reported in parts per million (ppm) relative to residual CHCl3 peak (7.27 and 77.1 ppm, 1H and 13C), and CFCl3 (19F). All coupling constants (J) are reported in hertz (Hz). Abbreviations are: s, singlet; d, doublet; t, triplet; q, quartet; brs, broad singlet. High-resolution mass spectra were measured on an Agilent 1200 HPLC chromatograph (Palo Alto, CA, USA) with Agilent 6210 mass spectrometer (Santa Clara, CA, USA) (HR-TOF-MS, ESI + ionization in acetonitrile with 0.1% HFBA). Elemental compositions were determined by accurate mass measurement with standard deviation. Melting points were measured on a Boetius apparatus. Flash chromatography was performed using silica gel, 60 Å, 300 mesh. TLC analysis was carried out on aluminum plates coated with silica gel 60 F254, 0.2 mm thickness. The plates were visualized using a 254 nm UV lamp.

Theoretical Calculations

All structures were optimized without restrictions at the MP2/6-311++G(d,p) level of theory. Frequency calculations were performed on the optimized geometry at the same level of theory. All calculations were performed by the use of Gaussian09 program suite [25].

3.2. Synthesis

3.2.1. Reactions of Allyl Halides with Triflamide in the Presence NBS + MeCN

To solution of 1 g (6.7 mmol) of triflamide and 6.7 mmol of allyl halide 1, 2 in 30 mL of acetonitrile added 1.19 g (6.7 mmol) of NBS and reaction mixture was stirred in the dark for 24 h. Solvent was removed in vacuum, then the succinimide was precipitated with diethyl ether, filtered off, and ether removed in a vacuum. Analytically pure samples of substances were separated by column chromatography (0.063–0.2 mm, Acros Organics, Waltham, MA, USA). From the hexane–ether = 1:1 eluate, not reacted triflamide and dibromides were isolated, and from the diethyl ether:hexane = 4:1 or diethyl ether eluates amidines 4, 5 were obtained.
N-(2-Bromo-3-chloropropyl)-N’-(trifluoromethylsulfonyl)acetamidamide, 4. Yield 0.7 g, 43.2%. Oil. 1H NMR (400 MHz, CDCl3) δ 6.85 (s, 1H, NH), 4.35 (ddd, J = 12.5, 8.3, 4.1 Hz, 1H, CHBr), 4.15 (ddd, J = 14.5, 5.9, 4.1 Hz, 1H, CHAHNH), 3.93 (dd, J = 11.8, 4.1 Hz, 1H, CHHBNH), 3.76 (dd, J = 11.9, 8.3 Hz, 1H, CHAHCl), 3.68 (ddd, J = 14.5, 8.3, 5.9 Hz, 1H, CH2Cl), 2.53 (s, 3H, CH3). 13C NMR (100 MHz, CDCl3) δ 169.5 (C=NTf), 121.4 (q, J = 319.4 Hz, CF3), 48.3 (CHBr), 46.6 (CH2NH), 45.6 (CH2Cl), 22.1 (CH3). 19F NMR (376 MHz, CDCl3) δ −78.99. IR (thin): 3334 (NH), 3137, 2928, 1719, 1566 (C=N), 1430, 1323, 1213, 1198 (CF3), 1136, 1086, 1055, 929, 775, 745, 663 (C–Br), 601, 542, 474. HRMS (ESI): m/z: [M+H]+ calcd for C6H9BrClF3N2O2S: 343,92087; found [M+H]+: 344.92869.
N-(2,3-Dibromopropyl)-N’-(trifluoromethylsulfonyl)acetamidamide,5. Yeild 1.1 g, 60.1%. Oil. 1H NMR (400 MHz, CDCl3) δ 7.14 (s, 1H, NH), 4.38 (tt, J = 8.6, 4.1 Hz, 1H, CHBr), 4.20 (ddd, J = 14.6, 6.0, 3.9 Hz, 1H, CH2NH), 3.83 (dd, J = 10.9, 4.3 Hz, 1H, CH2NH), 3.65 (m, 2H, CH2Br), 2.52 (s, 3H, CH3). 13C NMR (100 MHz, CDCl3) δ169.7 (C=NTf), 121.4 (q, J = 319.6 Гц, CF3), 47.7 (CHBr), 47.4 (CH2NH), 32.8 (CH2Br), 21.9 (CH3). IR (thin): 3335, 3227 (NH), 3136, 2943, 1774, 1721, 1580, 1562 (C=N), 1428, 1373, 1323, 1278, 1215, 1197 (CF3), 1135, 1081, 1053, 914, 833, 775, 746, 665, 602 (C–Br), 548, 476. 19F NMR (376 MHz, CDCl3) δ −78.86. Anal. calcd. for (C6H9Br2F3N2O2S): C, 18.48; H, 2.33; F, 14.61; N, 7.18; S, 8.22. Found: C, 18.88; H, 2.60; F, 15.00; N, 7.53; S, 8.56.

3.2.2. Reactions of Allyl Halides with Triflamide in the Presence NIS + MeCN

To the solution of 1 g (6.7 mmol) of triflamide and 6.7 mmol of allyl halide 1, 2 in 30 mL of acetonitrile added 1.53 g (6.7 mmol) of NIS and reaction mixture was stirred in the dark for 24 h. Solvent was removed in vacuum, then the succinimide was precipitated with diethyl ether, mixture were cooled and succinimide was filtered off, ether removed in a vacuum. Analytically pure samples of substances were separated by column chromatography (0.063–0.2 mm, Acros Organics, Waltham, MA, USA). From the hexane–ether = 1:1 eluate, not reacted triflamide and dibromides were isolated, and from the diethyl ether:hexane = 4:1 or diethyl ether eluates amidines 8, 9 were obtained.
N-(2-Iodo-3-chloropropyl)-N’-(trifluoromethylsulfonyl)acetamidamide, 8. Yield 0.65 g, 41.4%. Oil. 1H NMR (400 MHz, CDCl3) δ 6.96 (s, 1H, NH), 4.44 (ddd, J = 12.9, 8.6, 4.4 Hz, 1H, CHI), 4.05 (dd, J = 9.9, 4.7 Hz, 1H, CHAHNH), 4.00 (dd, J = 11.7, 4.4 Hz, 1H, CHHBNH), 3.81 (dd, J = 11.7, 9.9 Hz, 1H, CHAHCl), 3.71 (ddd, J = 14.4, 8.6, 6.2 Hz, 1H, CHHBCl), 2.52 (s, 3H, CH3). 13C NMR (100 MHz, CDCl3) δ 169.4 (C=NTf), 121.3 (q, J = 319.59 Гц, CF3), 48.0 (CH2NH), 47.5 (CH2Cl), 26.1 (CHI), 22.1 (CH3). 19F NMR (376 MHz, CDCl3) δ −78.91. IR (thin): 3326, 3231 (NH), 3080, 2928, 2859, 1723, 1664, 1577, 1553 (C=N), 1428, 1371, 1322, 1215, 1197 (CF3), 1140, 1082, 1050, 939, 846, 774, 745, 708, 662 (C–I), 636, 600, 528, 475. Anal. calcd. for (C6H9ClF3IN2O2S): C, 18.36; H, 2.31; F, 14.52; N, 7.14; S, 8.17; Found: C, 18.50; H, 2.71; F, 15.07; N, 7.53; S, 9.03.
N-(3-Bromo-2-iodopropyl)-N’-(trifluoromethylsulfonyl)acetamidamide, 9. Yield: 0.41 g, 27.5%. Oil. 1H NMR (400 MHz, CDCl3) δ 7.43 (br.s, 1H, NH), 4.60–4.43 (m, 1H, CHI), 4.14–3.94 (m, 2H, CH2NH), 3.78–3.65 (m, 2H, CH2Br), 2.53 (s, 3H, CH3). 13C NMR (100 MHz, CDCl3) δ 169.5 (C=NTf), 121.3 (q, J = 319.4 Hz, CF3), 48.9 (CH2NH), 35.1 (CH2Br), 25.8 (CHI), 22.1 (CH3). 19F NMR (376 MHz, CDCl3) δ −78.81. IR (thin): 3325, 3241 (NH), 3093, 3036, 2927, 2852, 1727, 1650, 1586, 1555 (C=N), 1428, 1375 (SO2), 1321, 1268, 1211, 1195 (CF3), 1139, 1081, 1049, 1007, 970, 945, 912, 811, 775, 740, 662, 639, 604, 580, 527, 475. Anal. calcd. for (C6H9BrF3IN2O2S) C, 16.49; H, 2.08; F, 13.04; N, 6.41; S, 7.34. Found: C, 16.50; H, 2.19; F, 12.73; N, 6.18; S, 7.59.

3.2.3. Reaction of Allyl Alcohol with Triflamide in the NBS + MeCN System

To a solution of 1.00 g (6.7 mmol) of triflamide and 0.39 g (6.7 mmol) of allyl alcohol 4 in 25 mL of acetonitrile was added 1.19 g (6.7 mmol) of NBS, and the reaction mixture was kept in the dark for 24 h. The solvent was removed under reduced pressure, the residue was dissolved in 20 mL of diethyl ether, cooled and the formed succinimide was filtered off. The filtrate was evaporated in vacuum, the residue (1.79 g) was placed on a silica gel column (0.063–0.2 mm, Acros Organics, Waltham, MA, USA) and eluted with ether:hexane = 1:1 mixture, isolating unreacted triflamide (~0.4 g), then with ether, obtaining N-(2-bromo-3-hydroxypropyl)-N’-(trifluoromethylsulfonyl)acetamidamide 12 as a colorless oil.
N-(2-Bromo-3-hydroxypropyl)-N’-(trifluoromethylsulfonyl)acetamidamide, 12. Yield 0.34 g, 26%. The product was obtained earlier and described in [16].

3.2.4. Reaction of Allyl Cyanide with Triflamide in the System NBS+MeCN

To a solution of 1.00 g (6.7 mmol) of triflamide and 0.45 g (6.7 mmol) of allyl cyanide in 40 mL of acetonitrile was added 1.19 g (6.7 mmol) of NBS, and the reaction mixture kept in the dark for 24 h. The solvent was removed under reduced pressure, the residue dissolved in 40 mL of diethyl ether, kept in a refrigerator and the formed succinimide filtered off. The filtrate was evaporated in a vacuum, the residue (~1.81 g) was placed on a silica gel column (0.063–0.2 mm, Acros Organics, Waltham, MA, USA) and eluted with ether:hexane (1:1) to give unreacted triflamide (0.2 g), then with ether to afford 1.08 g N-(2-bromo-3-cyanopropyl)-N’-(triflyl)acetimidamide 14 as a yellow oil.
N-(2-Bromo-3-cyanopropyl)-N’-(trifluoromethylsulfonyl)acetimidamide, 14. Oil. Yield 60%. 1H NMR (400 MHz, CDCl3) δ 7.21 (br t, J = 5.6 Hz, 1H, NH), 4.36 (ddd, J = 11.1, 7.1, 5.6 Hz, 1H, CHBr), 3.90 (ddd, J = 14.4, 5.6, 5.6 Hz, 1H, CHAN), 3.80 (ddd, J = 14.4, 7.1, 5.6 Гц, 1H, CHBN), 3.04 (dd, J = 5.6, 2.6 Гц, 2H, CH2CN), 2.54 (s, 3H, CH3). 13C NMR (100 MHz, CDCl3) δ 170.2 (C=NTf), 121.4 (q, J = 320.0 Hz, CF3), 115.9 (C≡N), 48.0 (CH2N), 41.8 (CHBr), 25.8 (CH2CN), 21.9 (CH3). 19F NMR (376 MHz, CDCl3) δ −78.88. IR (thin): 3324 (NH), 3135, 3025, 2952, 2933, 2259 (C≡N), 1711, 1560 (NHC=N), 1430, 1325, 1195, 1139, 1049, 747, 659 (C–Br), 600, 475. HRMS (ESI): m/z: [M+H]+ calcd for C7H9BrF3N3O2S+: 335.962919; found: 335.962880.

3.2.5. Reaction of Allyl Cyanide with Triflamide in the System NBS + CH2Cl2

To a solution of 1.00 g (6.7 mmol) of triflamide and 0.45 g (6.7 mmol) of allyl cyanide in 40 mL of CH2Cl2 was added 1.19 g (6.7 mmol) of NBS. The reaction was carried out for 24 h in the dark. Then, the solvent was removed under reduced pressure, the residue was dissolved in 40 mL of diethyl ether, placed in a refrigerator for 1 h, and the formed succinimide was filtered off. The ether fraction was evaporated in vacuum, the residue (~2.21 g) was placed on a silica gel column (0.063–0.2 mm, Acros Organics, Waltham, MA, USA) and eluted with hexane to give 3,4-dibromobutanenitrile 19 (0.40 g, 26%), followed by ether:hexane = 1:1, isolating unreacted triflamide (0.6 g), then with ether:hexane (4:1) to afford 3-bromo-4-hydroxybutanenitrile 20 (0.20 g, 18%), and hexane:chloroform:ether (1:2:2) to obtained N-(2-bromo-3-cyanopropyl)triflamide 15 (0.15 g, 19%) and N-(1-bromo-3-cyanoprop-2-yl)triflamide 18 (0.10 g, 13%).
N-(2-Bromo-3-cyanopropyl)trifluoromethanesulfonamide, 15. Yield 19%. Oil. 1H NMR (400 MHz, CDCl3) δ 6.03 (t, J = 5.5 Hz, 1H, NH), 4.18–4.09 (m, 1H, CHBr), 3.67–3.57 (m, 2H, CH2N), 2.92 (t, J = 6.0 Hz, 2H, CH2CN). 13C NMR (100 MHz, CDCl3) δ 119.3 (q, J = 320.1 Hz, CF3), 115.3 (C≡N), 49.0 (CH2NH), 43.1 (CHBr), 23.4 (CH2CN). 19F NMR (376 MHz, CDCl3) δ −77.23. IR (thin): 3199 (NH), 2923, 2259 (C≡N), 1723, 1615, 1454, 1440, 1380 (SO2), 1230 (CF3), 1197, 1145, 1098, 1067, 1045, 974, 924, 891, 829, 763, 677, 610, 587, 518. Anal. calcd. for (C5H6BrF3N2O2S): C, 20.35; H, 2.05; N, 9.49; Br, 27.08; S, 10.87. Found: C, 20.13; H, 2.10; N, 9.98; Br, 27.18; S 10.12.
N-(1-Bromo-3-cyanopropan-2-yl)trifluoromethanesulfonamide, 18. Yield 13%. Oil. 1H NMR (400 MHz, CDCl3) δ 6.17 (d, J = 7.9 Hz, 1H, NH), 4.23 (quint, J = 6.3 Hz, 1H, CHNH), 3.71 (t, J = 6.0 Hz, 2H, CH2Br), 3.09 (d, J = 5.7 Hz, 2H, CH2CN). 13C NMR (100 MHz, CDCl3) δ 119.5 (q, J = 321.0 Hz, CF3), 115.8 (C≡N), 51.8 (CHN), 33.5 (CH2Br), 25.1 (CH2CN). 19F NMR (376 MHz, CDCl3) δ −76.8.
3,4-Dibromobutanenitrile, 19. Yield 26%. Oil. 1H NMR (400 MHz, CDCl3) δ 4.18 (ddd, J = 11.8, 6.2, 5.2 Hz, 1H, CHBr), 3.94 (dd, J = 11.8, 5.2 Hz, 1H, CHAHBr), 3.86 (dd, J = 11.8, 6.2 Hz, 1H, CHHBBr), 3.11 (dd, J = 17.2, 5.2 Hz, 1H, CHAHCN), 3.05 (dd, J = 17.2, 6.2 Hz, 1H, CHHBCN). 13C NMR (100 MHz, CDCl3) δ 116.6 (C≡N), 65.4 (CHBr), 46.1 (CH2Br), 24.2 (CH2C≡N). IR (thin): 3420, 2957, 2928, 2883, 2257 (C≡N), 2066, 1773, 1723, 1649, 1636, 1625, 1577, 1562, 1546, 1457, 1413, 1379, 1343, 1289, 1198, 1149, 1088, 1060, 1028, 976, 942, 916, 876, 846, 724, 645 (C–Br), 607, 539. Anal. calcd. for C4H5Br2N: C, 21.17; H, 2.22; Br, 70.43; N, 6.17. Found: C, 21.10; H, 2.11.
3-Bromo-4-hydroxybutanenitrile,20. Yield 18%. Oil. 1H NMR (400 MHz, CDCl3) δ 4.17 (ddd, J = 12.1, 6.0, 5.5 Hz, 1H, CHBr), 3.93 (dd, J = 12.1, 5.5 Hz, 1H, CHAHBr), 3.85 (dd, J = 12.1, 6.0 Hz, 1H, CHHBBr), 3.10 (dd, J = 17.3, 5.5 Hz, 1H, CHAHCN), 3.04 (dd, J = 17.3, 6.6 Hz, 1H, CHHBCN), 2.59 (br.s, 1H, OH). 13C NMR (100 MHz, CDCl3) δ 116.7 (C≡N), 65.3 (CH2OH), 46.0 (CHBr), 24.2 (CH2C≡N). IR (thin): 3430, 2962, 2928, 2256, 1783, 1725, 1613, 1543, 1453, 1413, 1380, 1350, 1285, 1228, 1198, 1148, 1087, 1058, 1029, 976, 944, 917, 863, 822, 727, 672, 645, 608, 580, 537, 512. Anal. calcd. for C4H6BrNO: C, 29.20; H, 3.40; Br, 48.00; N, 9.51. Found: C, 29.29; H, 3.69; Br, 48.72; N, 8.54.

3.2.6. Reaction of Allyl Chloride with Triflamide in the NBS + THF System

To a solution of 1.00 g (6.7 mmol) of triflamide and 0.51 g (6.7 mmol) of allyl chloride 1 in 30 mL of tetrahydrofuran was added 1.19 g (6.7 mmol) of NBS, and the reaction mixture was kept in the dark for 24 h. The solvent was removed under reduced pressure, the residue was dissolved in 20 mL of diethyl ether, mixture was cooled and succinimide was filtered off. The filtrate was evaporated in vacuo, the residue (~2.20 g) was placed on a silica gel column (0.063–0.2 mm, Acros Organics, Waltham, MA, USA) and eluted with ether:hexane = 1:1 mixture, isolating unreacted triflamide, then with ether, obtaining 0.30 g of 1,2- dibromo-3-chloropropane 16 as a yellow oil. Product 16 was obtained and described earlier [26].

3.2.7. Reaction of Allyl Cyanide with Triflamide in the System NBS + THF

To a solution of 1.00 g (6.7 mmol) of triflamide and 0.45 g (6.7 mmol) of allyl cyanide in 40 mL of THF 1.19 g (6.7 mmol) of NBS was added. The reaction was carried out for 24 h in the dark. The solvent was removed under reduced pressure, the residue dissolved in 40 mL of diethyl ether, placed in a refrigerator for 1 h, and the formed succinimide was filtered off. The ether fraction was evaporated in vacuum, and the residue (~2.21 g) was placed on a silica gel column (0.063–0.2 mm, Acros Organics, Waltham, MA, USA) and eluted with ether:hexane (1:1), isolating unreacted triflamide (0.4 g), then with ether:hexane (4:1) to give N-(4-(2-bromo-3-cyanopropoxy)butyl)triflamide 17 (1.16 g, 79%).
N-(4-(2-Bromo-3-cyanopropoxy)butyl)trifluoromethanesulfonamide,17. (45%). Oil. 1H NMR (400 MHz, CD3CN) δ 5.86 (s, 1H, NH), 4.24–4.07 (m, 1H, CHBr), 3.85–3.72 (m, 1H, CHBrCHAHO), 3.72–3.61 (m, 1H, CHBrCHHBO), 3.61–3.37 (m, 4H), 3.36–3.22 (m, 1H), 3.08–3.00 (m, 1H), 1.79–1.55 (m, 4H). 13C NMR (100 MHz, CDCl3) δ 119.7 (q, J = 321.3 Hz, CF3), 116.6 (C≡N), 73.0 (CHBrCH2O), 70.9 (CH2CH2O), 44.0 (CHBr), 41.8 (CH2NH), 27.1 (CHBr), 26.2 (CH2CH2), 26.63 (CH2CH2), 24.6 (CH2C≡N). 19F NMR (376 MHz, CDCl3) δ −77.21. IR (thin): 3304, 3221 (NH), 2946, 2876, 2258 (C≡N), 1652, 1452, 1439, 1420, 1373 (SO2), 1284, 1230 (CF3), 1192, 1148, 1078, 990, 920, 877, 812, 742, 609, 579, 511. Anal. calcd. for C9H14BrF3N2O3S: C, 29.44; H, 3.84; N, 7.63; Br, 21.76; found: C, 29.90; H, 3.52; N, 7.42; Br, 21.90.

3.2.8. Reaction of N-(2-bromo-3-chloropropyl)-N’-(trifluoromethylsulfonyl)acetamidamide 4 with K2CO3 in MeCN

To a solution of amidine 4 0.2 g (0.6 mmol) in acetonitrile (10 mL) was added a 2-fold excess of potassium carbonate 0.17 g (1.2 mmol) and stirred for 4 h. The precipitate in the form of salt was filtered off, the acetonitrile fraction was distilled off under reduced pressure, obtaining 0.13 g of 5-(chloromethyl)-2-methyl-1-((trifluoromethyl)sulfonyl)-4,5- dihydro-1H-imidazole 21 as a colorless oil.
5-(Chloromethyl)-2-methyl-1-(trifluoromethylsulfonyl)-4,5-dihydro-1H-imidazole,21. Yield 0.13 g, 81.3%. 1H NMR (400 MHz, CDCl3) δ 4.64-4.49 (m, 1H, CHN), 4.07 (d.d.d, J = 16.0, 9.3, 1.9 Hz, 1H, CHAHN), 3.95 (dd, J = 16.0, 1.8 Hz, 1H, CHHBN), 3.73 (dd, J = 11.5, 6.1 Hz, 1H, CHAHCl), 3.67 (dd, J = 11.5, 3.1 Hz, 1H, CHHBCl), 2.29 (br. t, J = 1.6 Hz, 3H, CH3).). 13C NMR (100 MHz, CDCl3) δ 153.5 (C=N), 118.64 (q, J = 322.64 Гц, CF3), 61.0 (CHN), 57.5 (CH2N), 46.1 (CH2Cl), 16.3 (CH3). 19F NMR (376 MHz, CDCl3) δ −74.834. IR (thin): 2947, 2879, 2856, 2622, 1723, 1676, 1438, 1404, 1388 (SO2), 1352, 1312, 1283, 1239, 1207 (CF3), 1157, 1101, 1075, 1058, 1023, 1003, 939, 898, 872, 784, 767, 736, 681, 666, 621, 592, 536, 505. Anal. calcd. For (C6H8ClF3N2O2S): C, 27.23; H, 3.05; F, 21.54; N, 10.59; S, 12.11. Found: C, 27.42; H, 3.15; F, 21.67; N, 10.70; S, 12.23.

3.2.9. Reaction of N-(2,3-dibromopropyl)-N’-((trifluoromethyl)sulfonyl)acetamidamide 5 with K2CO3 in MeCN

To a solution of amidine 5 0.16 g (0.4 mmol) in acetonitrile (10 mL) was added a 2-fold excess of potassium carbonate 0.11 g (0.8 mmol) and stirred for 4 h. The precipitate in the form of salt was filtered off, the acetonitrile fraction was distilled off under reduced pressure, obtaining 5-(bromomethyl)-2-methyl-1-((trifluoromethyl)sulfonyl)-4,5-dihydro- 1H-imidazole 22 as a colorless oil.
5-(Bromomethyl)-2-methyl-1-(trifluoromethylsulfonyl)-4,5-dihydro-1H-imidazole, 22. Yield 0.11 g, 91.7%. Oil. 1H NMR (400 MHz, CDCl3) δ 4.59–4.49 (m, 1H, CHN), 4.07 (d.d.d, J = 16.0, 9.4, 2.1 Hz, 1H, CHAHN), 3.90 (d.d.d, J = 16.0, 3.3, 2.1 Hz, 1H, CHHBN), 3.57-3.52 (m, 2H, CH2Br), 2.27 (br. t, J = 1.6 Hz, 3H, CH3). 13C NMR (100 MHz, CDCl3) δ 153.4 (C=N), 121.79 (q, J = 324.2 Гц, CF3), 60.7 (CHN), 58.5 (CH2N), 34.7 (CH2Br), 16.4 (CH3). 19F NMR (376 MHz, CDCl3) δ -74.77. IR (thin): 2945, 2878, 2606, 1675, 1437, 1404, 1388 (SO2), 1347, 1305, 1238, 1206 (CF3), 1156, 1099, 1073, 1054, 1012, 992, 936, 889, 862, 771, 692, 672, 659,641, 613, 590, 574, 535, 477, 416. Anal. calcd. For C6H8BrF3N2O2S: C, 23.31; H, 2.61; F, 18.44; N, 9.06; S, 10.37. Found: C, 23.42; H, 2.71; F, 18.55; N, 9.13; S, 10.42.

3.2.10. Hydrolysis of 5-(bromomethyl)-2-methyl-1-(trifluoromethylsulfonyl)-4,5-dihydro- 1H-imidazole 22

Compound 17 0.11 mg (0.36 mmol) was subjected to hydrolysis with the formation of N-(3-bromo-2-((trifluoromethyl)sulfonamido)propyl)acetamide 23.
N-(3-Bromo-2-((trifluoromethyl)sulfonamido)propyl)acetamide, 23. Yield 0.10 g, 83.3%. 1H NMR (400 MHz, CDCl3) δ 7.50 (d, J = 7.3 Hz, 1H, CHNH), 6.33 (t, J = 4.8 Hz 1H, CH2NH), 3.96–3.87 (m, 1H, CHNH), 3.69–3.60 (m, 2H, CH2NH), 3.44 (dd, J = 10.9, 7.9 Hz, 2H, CH2Br), 2.05 (s, 3H). 13C NMR (100 MHz, CDCl3) δ 173.6 (C=O), 121.67 (q, J = 320.94 Гц, CF3), 56.0 (CHN), 42.5 (CH2N), 32.7 (CH2Br), 22.9 (CH3). 19F NMR (376 MHz, CDCl3) δ −77.15. IR (thin): 3352 (NH), 3119 (NH), 2957, 2922, 2853, 2255, 1656 (C=O), 1561, 1547, 1431, 1377 (SO2), 1324, 1199 (CF3), 1141, 1052, 984, 909, 735, 651, 614, 476. Anal. calcd. For (C6H10BrF3N2O3S): C, 22.03; H, 3.08; F, 17.42; N, 8.56; S, 9.80. Found: C, 22.19; H, 3.22; F, 17.54; N, 8.63; S, 9.91.

3.2.11. Reaction of N-(2-bromo-3-cyanopropyl)-N’-(triflyl)acetimidamide 14 with Base in Acetonitrile

To a solution of amidine 14 (0.27 g, 0.8 mmol) in acetonitrile (10 mL), 2-fold excess of a base (potassium carbonate or triethylamine) was added and stirred for 2 h. The formed salt was filtered off, the acetonitrile fraction was distilled off in a vacuum, affording N-(3-cyanoallyl)-N’-((trifluoromethyl)sulfonyl)acetimidamide 25 (0.19 g, 93%),
N-(3-Cyanoallyl)-N’-(trifluoromethyl)sulfonyl)acetimidamide,25. Yield 93%. Oil. 1H NMR (400 MHz, CD3CN) δ 7.92 (br s, 1H, NH), 6.49 (dt, J = 11.2, 6.0 Hz, 1H, =CHCH2), 5.63 (dt, J = 11.2, 1.4 Hz, 1H, =CHC≡N), 4.20 (td, J = 6.0, 1.4 Hz, 2H, CH2N), 2.40 (s, 3H, CH3). 13C NMR (100 MHz, CDCl3) δ 171.0 (C=NTf), 148.4 (=CHNH); 121.09 (q, J = 319.2 Hz, CF3), 116.0 (C≡N), 102.5 (=CHCH2), 43.4 (CH2); 21.8 (CH3). 19F NMR (376 MHz, CDCl3) δ −79.03. IR (thin): 3323, 3133 (NH), 3082, 2944, 2259 (C≡N), 2228, 1772, 1715, 1661, 1588, 1561, 1427, 1384 (SO2), 1354, 1329, 1279, 1216 (CF3), 1195, 1141, 1102, 1081, 1068, 1042, 975, 920, 901, 871, 843, 776, 739, 604, 584, 537, 475, 435. HRMS (ESI): m/z: [M+H]+ calcd for C7H8F3N3O2S+: 256.036757; found: 256.036460.

3.2.12. Reaction of Allyl Cyanide with Triflamide in the System NBS + MeCN + K2CO3

To a solution of 1.00 g (6.7 mmol) of triflamide and 0.45 g (6.7 mmol) of allyl cyanide in 40 mL of CH3CN was added 1.19 g (6.7 mmol) of NBS. The reaction was carried out for 24 h in the dark. Then, 1.85 g (13.4 mmol) of K2CO3 was added and stirred for another 3 h. The precipitate was filtered off, the solvent removed under reduced pressure, the black residue (~2.43 g) was placed on a silica gel column (0.063-0.2 mm, Acros Organics, Waltham, MA, USA) and eluted with ether:hexane (4:1) giving N-(3-cyanoprop-1-en-1-yl)-N’-(trifluoromethyl- sulfonyl)acetimidamide 26 (1.30 g, 75%).
N-(3-Cyanoprop-1-en-1-yl)-N’-(trifluoromethylsulfonyl)acetimidamide,26. Yield 75%. Oil. 1H NMR (400 MHz, CDCl3) δ 8.66 (br. s, 1H, NH), 6.94 (t, J = 8.8 Hz, 1H, =CHCN), 5.06 (dd, J = 16.0, 7.5 Hz, 1H, =CH), 3.28 (dd, J = 7.5, 1.3 Hz, 2H, CH2), 2.56 (s, 3H, CH3). 13C NMR (100 MHz, CDCl3) δ 166.9 (C=NTf), 125.4 (=CHCH2), 119.2 (q, J = 318.8 Hz, CF3); 117.3 (NC), 104.5 (=CHCN); 29.6 (CH2NH); 21.6 (CH3). 19F NMR (376 MHz, CDCl3) δ −79.01. IR (thin): 3330, 3120 (NH), 3079, 2958, 2929, 2860, 2256 (C≡N), 2230, 1774, 1711, 1680, 1653, 1576, 1541, 1431, 1381 (SO2), 1326, 1267, 1215 (CF3), 1194, 1140, 1049, 950, 907, 838, 785, 753, 683, 643, 615, 581, 534, 497.

3.2.13. Reaction of N-(4-(2-bromo-3-cyanopropoxy)butyl)triflamide 17 with a Base in Acetonitrile

To a solution (0.20 g, 0.05 mmol) of N-(4-(2-bromo-3-cyanopropoxy)butyl)triflamide 10 in acetonitrile (10 mL) was added a 2-fold excess of potassium carbonate (0.01 g, 0.1 mmol) and stirred for 4 h. The precipitated salt was filtered off, the acetonitrile fraction was distilled off in a vacuum to afford N-(4-((3-cyanoprop-1-en-1-yl)oxy)butyl)triflamide 27 and N-(4-(3-cyanolyl)oxy)butyl)triflamide 28 in the ratio of 1:2.
N-(4-(3-Cyanoprop-1-en-1-yl)oxy)butyl)trifluoromethanesulfonamide, 27; N-(4-((3-cyanoallyl)oxy)butyl)trifluoromethanesulfonamide, 28. Oil. 1H NMR (400 MHz, CD3CN) δ 6.74 (comp. 11; dt, J = 16.2, 3.8 Hz, 1H, =CHO), 6.57 (comp. 12; dt, J = 11.3, 5.6 Hz, 1H, =CHCH2O), 5.87 (comp. 12; br s, 1H, NH), 5.64 (comp. 11; d, J = 16.2, 1H, CH2CH=CHO), 5.51 (comp. 12; d, J = 11.3, 1H, NCCH=), 5.45 (comp. 11; br s, 1H, NH), 4.38–4.26 (comp. 12; m, 2H, OCH2CH2), 4.18–4.09 (comp. 11; m, 2H, OCH2CH2), 3.64–3.25 (11 + 12, m), 1.90–1.60 (11 + 12, m). 13C NMR (100 MHz, CDCl3) δ 150.2, 149.6 (OCH=); 118.6, 118.2 (C≡N); 101.1, 100.2 (=CH); 70.9, 70.7, 69.4, 69.1 (OCH2); 44.3, 44.2 (CH2NH); 27.57, 27.50, 26.67, 26.50. 19F NMR (376 MHz, CDCl3) δ −77.18. Anal. calcd. for C9H13F3N2O3S: C, 37.76; H, 4.58; F, 19.91; N, 9.79; S, 11.20; found: C, 39.02; H, 4.95; F, 21.71; N, 9.02; S, 11.90.

4. Conclusions

Substituted amidines were obtained for the first time from allyl halides. Amidines prepared from the reaction of triflamide, allyl halide, NBS, and acetonitrile were successfully converted to the corresponding imidazolidines in good yields. Allyl cyanide reacts with triflamide in the presence of NBS to give, depending on the solvent, different products of oxidative triflamidation. In methylene chloride, two regioisomers of the product of halosulfonamidation are formed, whereas in acetonitrile and THF the main products are those with a solvent interception. The amidine, obtained from the reaction in acetonitrile, behaves differently from all other earlier studied β-bromoamidines, which, when treated with a base, underwent cyclization to imidazolines in quantitative yield. In contrast, N-(2-bromo-3-cyanopropyl)-N’-(triflyl)ethaneimidamide undergoes dehydrobromination with the formation of isomeric linear products, N-[(E)-3-cyanopropen-1-yl)]-N’-(triflyl)ethaneimidamides with the new C=C bond in the α- or β-position to the cyano group. In the same manner, N-[4-(2-bromo-3-cyanopropoxy)butyl]triflamide obtained as the solvent interception product from the reaction in THF, was dehydrobrominated to the equimolar isomeric mixture of linear products with the new C=C conjugated with either the cyano group or the oxygen atom. No cyclization occurred to 1,4-oxazocanes as in all other earlier studied similar products. High-level calculations allowed us to explain the observed unusual course of dehydrobromination and the formation of different regioisomers.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/molecules27206910/s1, Supporting Information: experimental details, NMR spectra, HRMS data for new products.

Author Contributions

Conceptualization, B.A.S., M.Y.M., and A.S.G.; methodology, M.Y.M. and B.A.S.; investigation, A.S.G., I.A.G.; writing—original draft preparation, A.S.G.; writing—review and editing, V.V.A., M.Y.M. and B.A.S.; calculations, B.A.S.; supervision, M.Y.M. and B.A.S.; funding acquisition, M.Y.M. and B.A.S. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the Russian Science Foundation (project 22-13-00036).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data presented in this study are available on request from the corresponding author.

Acknowledgments

All spectral measurements were performed on the equipment of the Baikal Center for Collective Use, SB RAS. We thank A.V. Kuzmin for HRMS analysis performed on the Research Facilities for Physical and Chemical Ultramicroanalysis, Limnological Institute, SB RAS.

Conflicts of Interest

The authors declare no conflict of interest.

Sample Availability

Samples of the compounds are available from the authors.

References

  1. Yamasaki, A.; Terauchi, H.; Takemura, S. Reaction of N-Haloamide. XXVII. Reaction of N, N-Dihaloamides with Dienes. Chem. Pharmaceut. Bull. 1976, 24, 2841–2849. [Google Scholar] [CrossRef] [Green Version]
  2. Wei, H.-X.; Kim, S.H.; Li, G. Electrophilic Diamination of Alkenes by Using FeCl3−PPh3 Complex as the Catalyst. J. Org. Chem. 2002, 67, 4777–4781. [Google Scholar] [CrossRef] [PubMed]
  3. Booker-Milburn, K.I.; Guly, D.J.; Cox, B.; Procopiou, P.A. Ritter-Type Reactions of N-Chlorosaccharin:  A Method for the Electrophilic Diamination of Alkenes. Org. Lett. 2003, 5, 3313–3315. [Google Scholar] [CrossRef]
  4. Zhou, L.; Zhou, J.; Tan, C.K.; Chen, J.; Yeung, Y.-Y. N-Bromosuccinimide Initiated One-Pot Synthesis of Imidazoline. Org. Lett. 2011, 13, 2448–2451. [Google Scholar] [CrossRef]
  5. Sherman, E.S.; Chemler, S.R.; Tan, T.B.; Gerlits, O. Copper(II) Acetate Promoted Oxidative Cyclization of Arylsulfonyl-o-allylanilines. Org. Lett. 2004, 6, 1573–1575. [Google Scholar] [CrossRef]
  6. Fan, R.; Wen, F.; Qin, L.; Pu, D.; Wang, B. PhI(OAc)2 induced intramolecular oxidative bromocyclization of homoallylic sulfonamides with KBr as the bromine source. Tetrahedron Lett. 2007, 48, 7444–7447. [Google Scholar] [CrossRef]
  7. Huang, D.; Liu, X.; Li, L.; Cai, Y.; Liu, W.; Shi, Y. Enantioselective Bromoaminocyclization of Allyl N-Tosylcarbamates Catalyzed by a Chiral Phosphine–Sc(OTf)3 Complex. J. Am. Chem. Soc. 2013, 135, 8101–8104. [Google Scholar] [CrossRef]
  8. Cheng, Y.A.; Yu, W.Z.; Yeung, Y.-Y. Carbamate-Catalyzed Enantioselective Bromolactamization. Angew. Chem. Int. Ed. 2015, 54, 12102–12106. [Google Scholar] [CrossRef]
  9. Broggini, G.; Beccalli, E.M.; Borelli, T.; Brusa, F.; Gazzola, S.; Mazza, A. Intra-Intermolecular Palladium-Catalyzed Domino Reactions of Glycine Allylamides for the Synthesis of Diversely Functionalized Piperazinones. Eur. J. Org. Chem. 2015, 2015, 4261–4268. [Google Scholar] [CrossRef]
  10. Manick, A.-D.; Berhal, F.; Prestat, G. Synthesis of Six- and Seven-Membered Chloromethyl-Substituted Heterocycles via Palladium-Catalyzed Amino- and Oxychlorination. Synthesis 2016, 48, 3719–3729. [Google Scholar]
  11. Egart, B.; Lentz, D.; Czekelius, C. Diastereoselective Bromocyclization of O-Allyl-N-tosyl-hydroxylamines. J. Org. Chem. 2013, 78, 2490–2499. [Google Scholar] [CrossRef] [PubMed]
  12. Moskalik, M.Y.; Astakhova, V.V.; Shainyan, B.A. Oxidative sulfamidation as a route to N-heterocycles and unsaturated sulfonamides. Pure Appl. Chem. 2020, 92, 123–149. [Google Scholar] [CrossRef]
  13. Moskalik, M.Y.; Astakhova, V.V.; Shainyan, B.A. Reaction of trifluoromethanesulfonamide with heterodienes under oxidation conditions. Russ. J. Org. Chem. 2013, 49, 1567–1571. [Google Scholar] [CrossRef]
  14. Astakhova, V.V.; Moskalik, M.Y.; Ganin, A.S.; Shainyan, B.A. Reactions of N-Allyl- and N,N-Diallyltrifluoromethanesulfonamides with Carboxylic Acid Amides under Oxidizing Conditions. Russ. J. Org. Chem. 2018, 54, 855–860. [Google Scholar] [CrossRef]
  15. Shainyan, B.A.; Astakhova, V.V.; Ganin, A.S.; Moskalik, M.Y.; Sterkhova, I.V. Oxidative addition/cycloaddition of arenesulfonamides and triflamide to N-allyltriflamide and N,N-diallyltriflamide. RSC Adv. 2017, 7, 38951–38955. [Google Scholar] [CrossRef] [Green Version]
  16. Ganin, A.S.; Moskalik, M.Y.; Astakhova, V.V.; Sterkhova, I.V.; Shainyan, B.A. Heterocyclization and solvent interception upon oxidative triflamidation of allyl ethers, amines and silanes. Tetrahedron 2020, 76, 131374. [Google Scholar] [CrossRef]
  17. Cetiner, S.; Karakas, H.; Ciobanu, R.; Olariu, M.; Kaya, N.U.; Unsal, C.; Kalaoglu, F.; Sarac, A.S. Polymerization of pyrrole derivatives on polyacrylonitrile matrix, FTIR–ATR and dielectric spectroscopic characterization of composite thin films. Synth. Met. 2010, 160, 1189–1196. [Google Scholar] [CrossRef]
  18. Moskalik, M.Y.; Shainyan, B.A.; Ushakov, I.A.; Sterkhova, I.V.; Astakhova, V.V. Oxidant effect, skeletal rearrangements and solvent interception in oxidative triflamidation of norbornene and 2,5-norbornadiene. Tetrahedron 2020, 76, 131018. [Google Scholar] [CrossRef]
  19. Moskalik, M.Y.; Astakhova, V.V.; Shainyan, B.A. Divergent reactivity of divinylsilanes toward sulfonamides in different oxidative systems. RSC Adv. 2020, 10, 40514–40528. [Google Scholar] [CrossRef]
  20. Astakhova, V.V.; Moskalik, M.Y.; Shainyan, B.A. Solvent interception, heterocyclization and desilylation upon NBS-induced sulfamidation of trimethyl(vinyl)silane. Org. Biomol. Chem. 2019, 17, 7927–7937. [Google Scholar] [CrossRef]
  21. Moskalik, M.Y.; Garagan, I.A.; Astakhova, V.V.; Sterkhova, I.V.; Shainyan, B.A. Solvent-dependent oxidative triflamidation of alkenes and N(O)-Heterocyclization of the products. Tetrahedron 2021, 88, 132145. [Google Scholar] [CrossRef]
  22. Shainyan, B.A.; Meshcheryakov, V.I.; Sterkhova, I.V. A convenient synthesis and structure of N-trifluoromethylsulfonylamidines. Tetrahedron 2015, 71, 7906–7910. [Google Scholar] [CrossRef]
  23. Williams, T.I.; Denault, J.W.; Cooks, R.G. Proton affinity of deuterated acetonitrile estimated by the kinetic method with full entropy analysis. Int. J. Mass Spectrom. 2001, 210–211, 133–146. [Google Scholar] [CrossRef]
  24. Shainyan, B.A.; Chipanina, N.N.; Oznobikhina, L.P. The basicity of sulfonamides and carboxamides. Theoretical and experimental analysis and effect of fluorinated substituent. J. Phys. Org. Chem. 2012, 25, 738–747. [Google Scholar] [CrossRef]
  25. Frisch, M.J.; Trucks, G.W.; Schlegel, H.B.; Scuseria, G.E.; Robb, M.A.; Cheeseman, J.R.; Scalmani, G.; Barone, V.; Petersson, G.A.; Nakatsuji, H.; et al. Gaussian 09, Revision B.01; Gaussian Inc.: Wallingford, CT, USA, 2010. [Google Scholar]
  26. Babich, H.; Davis, D.L.; Stotzky, G. Dibromochloropropane (DBCP): A review. Sci. Total Environ. 1981, 17, 207–221. [Google Scholar] [CrossRef]
Scheme 1. Sc(OTf)3-catalyzed cyclization of allyl-N-tosylcarbamates.
Scheme 1. Sc(OTf)3-catalyzed cyclization of allyl-N-tosylcarbamates.
Molecules 27 06910 sch001
Scheme 2. NBS-induced intramolecular cyclization of N-allyl-N-benzyl-2-(tosylamido)acetamide.
Scheme 2. NBS-induced intramolecular cyclization of N-allyl-N-benzyl-2-(tosylamido)acetamide.
Molecules 27 06910 sch002
Scheme 3. Intramolecular bromoamination of O-allyl-N-hydroxytosylamide.
Scheme 3. Intramolecular bromoamination of O-allyl-N-hydroxytosylamide.
Molecules 27 06910 sch003
Scheme 4. NBS-induced reaction of triflamide with allyl chloride and bromide in acetonitrile.
Scheme 4. NBS-induced reaction of triflamide with allyl chloride and bromide in acetonitrile.
Molecules 27 06910 sch004
Scheme 5. NBS-induced reaction of triflamide with allyl iodide 6 in acetonitrile.
Scheme 5. NBS-induced reaction of triflamide with allyl iodide 6 in acetonitrile.
Molecules 27 06910 sch005
Scheme 6. Possible mechanism for the formation of dibromo-substituted amidine 5.
Scheme 6. Possible mechanism for the formation of dibromo-substituted amidine 5.
Molecules 27 06910 sch006
Scheme 7. NIS-induced reaction of triflamide with allyl chloride and bromide in acetonitrile.
Scheme 7. NIS-induced reaction of triflamide with allyl chloride and bromide in acetonitrile.
Molecules 27 06910 sch007
Scheme 8. NBS-induced reaction of triflamide with allyl alcohol 11 in acetonitrile.
Scheme 8. NBS-induced reaction of triflamide with allyl alcohol 11 in acetonitrile.
Molecules 27 06910 sch008
Scheme 9. NBS-induced reaction of triflamide with allyl cyanide 13 in acetonitrile.
Scheme 9. NBS-induced reaction of triflamide with allyl cyanide 13 in acetonitrile.
Molecules 27 06910 sch009
Scheme 10. NBS-induced reaction of triflamide with allyl chloride in THF.
Scheme 10. NBS-induced reaction of triflamide with allyl chloride in THF.
Molecules 27 06910 sch010
Scheme 11. NBS-induced reaction of triflamide with allyl cyanide 13 in THF.
Scheme 11. NBS-induced reaction of triflamide with allyl cyanide 13 in THF.
Molecules 27 06910 sch011
Scheme 12. NBS-induced reaction of triflamide with allyl cyanide 1 in CH2Cl2.
Scheme 12. NBS-induced reaction of triflamide with allyl cyanide 1 in CH2Cl2.
Molecules 27 06910 sch012
Scheme 13. Dehydrobromination of amidines 4, 5 and hydrolysis of imidazoline 22.
Scheme 13. Dehydrobromination of amidines 4, 5 and hydrolysis of imidazoline 22.
Molecules 27 06910 sch013
Scheme 14. Reaction of triflamide, alkene, NBS, and K2CO3 (one pot procedure).
Scheme 14. Reaction of triflamide, alkene, NBS, and K2CO3 (one pot procedure).
Molecules 27 06910 sch014
Scheme 15. Dehydrobromination of compound 17 with potassium carbonate.
Scheme 15. Dehydrobromination of compound 17 with potassium carbonate.
Molecules 27 06910 sch015
Scheme 16. Formation of compounds 14, 15, and 18.
Scheme 16. Formation of compounds 14, 15, and 18.
Molecules 27 06910 sch016
Figure 1. Optimized structures of the cyclic dimers of isomers 25 and 26.
Figure 1. Optimized structures of the cyclic dimers of isomers 25 and 26.
Molecules 27 06910 g001
Table 1. MP2/6-311++G(d,p) relative energies ΔE and free energies ΔG (kcal/mol) of amidines 25, 26, their dimers, and imidazoline 24.
Table 1. MP2/6-311++G(d,p) relative energies ΔE and free energies ΔG (kcal/mol) of amidines 25, 26, their dimers, and imidazoline 24.
StructureΔEΔG
Amidine 2520.516.1
Amidine 2616.713.0
½(25-dimer)2.556.8
½(26-dimer)2.566.8
Imidazoline 2400
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Ganin, A.S.; Moskalik, M.Y.; Garagan, I.A.; Astakhova, V.V.; Shainyan, B.A. Triflamidation of Allyl-Containing Substances:Unusual Dehydrobromination vs. Intramolecular Heterocyclization. Molecules 2022, 27, 6910. https://doi.org/10.3390/molecules27206910

AMA Style

Ganin AS, Moskalik MY, Garagan IA, Astakhova VV, Shainyan BA. Triflamidation of Allyl-Containing Substances:Unusual Dehydrobromination vs. Intramolecular Heterocyclization. Molecules. 2022; 27(20):6910. https://doi.org/10.3390/molecules27206910

Chicago/Turabian Style

Ganin, Anton S., Mikhail Yu. Moskalik, Ivan A. Garagan, Vera V. Astakhova, and Bagrat A. Shainyan. 2022. "Triflamidation of Allyl-Containing Substances:Unusual Dehydrobromination vs. Intramolecular Heterocyclization" Molecules 27, no. 20: 6910. https://doi.org/10.3390/molecules27206910

Article Metrics

Back to TopTop