Next Article in Journal
Optimized Nb-Based Zeolites as Catalysts for the Synthesis of Succinic Acid and FDCA
Next Article in Special Issue
Palladium-Catalyzed Cross-Coupling of Gem-Bromofluoroalkenes with Alkylboronic Acids for the Synthesis of Alkylated Monofluoroalkenes
Previous Article in Journal
The Budesonide-Hydroxypropyl-β-Cyclodextrin Complex Attenuates ROS Generation, IL-8 Release and Cell Death Induced by Oxidant and Inflammatory Stress. Study on A549 and A-THP-1 Cells
Previous Article in Special Issue
Evaluation of Self-Assembly Pathways to Control Crystallization-Driven Self-Assembly of a Semicrystalline P(VDF-co-HFP)-b-PEG-b-P(VDF-co-HFP) Triblock Copolymer
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Trifluoromethanesulfonamide vs. Non-Fluorinated Sulfonamides in Oxidative Sulfamidation of the C=C Bond: An In Silico Study

by
Anton V. Kuzmin
,
Mikhail Yu. Moskalik
and
Bagrat A. Shainyan
*
Siberian Division of Russian Academy of Sciences, A.E. Favorsky Irkutsk Institute of Chemistry, 664033 Irkutsk, Russia
*
Author to whom correspondence should be addressed.
Molecules 2020, 25(21), 4877; https://doi.org/10.3390/molecules25214877
Submission received: 25 September 2020 / Revised: 19 October 2020 / Accepted: 19 October 2020 / Published: 22 October 2020
(This article belongs to the Special Issue Organofluorine Chemistry)

Abstract

:
A theoretical analysis of the reaction of oxidative sulfamidation of several alkenes was performed in order to explain the various experimental observations and different reactivity of triflamide and non-fluorinated sulfonamides. Transformations occurring in the system alkene–sulfonamide in the presence of oxidative system (ButOCl + NaI) were analyzed at the MP2/DGDZVP//B3LYP/DGDZVP level of theory using the IEF-PCM method for taking into account the solvent acetonitrile (MeCN) effect. As the model substrates, styrene, trimethyl(vinyl)silane, dimethyl(divinyl)silane and diphenyl(divinyl)silane were chosen and mesylamide, triflamide, tosylamide and p-nosylamide were taken as the reagents. ButOI generated from ButOCl and NaI reacts with sulfonamides to give N-iodinated sulfonamides RSO2NHI and RSO2NI2 as active intermediates, the iodinating activity of the latter being notably higher. The analysis allowed to answer such challenging questions as different reactivity of nonfluorinated sulfonamides leading to aziridination and of triflamide resulting in the formation the main products of bis-triflamidation, or different regioselectivity of halogenation of styrene and trimethyl(vinyl)silane caused by a linear intermediate iodonium cation in the former case and a cyclic one in the latter.

1. Introduction

Reactions of sulfamidation and aziridination of alkenes are atom economy and green routes to C–N coupling reactions [1] widely used in organic synthesis [2,3], resulting in many of biologically active compounds [4,5,6]. Amides are able to react with non-activated alkenes only under harsh conditions [7]; therefore, many attempts were made to discover amidation/aziridination techniques under mild conditions as well as to increase the yield of the target products. At present, there are several approaches for the activation of sulfonamide, in which the sulfonamido-containing precursors are represented by λ3-iodinanes (PhI=NTs) [8,9,10,11,12,13,14] and -brominanes [15], azides [16,17,18], chloramine-T [19,20,21], bromamine-T [22,23], N,N-dichlorosulfonamides [24,25], N-bromosuccinimide (NBS) and N-iodosuccinimide (NIS) [26], nitrido complexes [27,28,29,30] and combinations of sulfonamides with external oxidants [31,32,33,34].
The use of chloramines RSO2NClNa allows to perform amination of unsaturated substrates without metal complex catalysts, whereas aziridination of alkenes and dienes is catalyzed, e.g., by N-bromoamides [35]. TsNIK is a more effective catalyst for aziridination of 4-methylstyrene with chloramine-T (yield 75%) than TsNBrNa (60%) or TsNClNa (45%) due to a weaker N–I bond in TsNIK favoring the formation of the complex of nitrenoid TsNI with Cu salts [36]. The reaction of olefins, chloramine-T and CO2 in the presence of ammonium bromide salts directly affords 5-substituted 2-oxazolidinones [37]. In the presence of transition metal salts, chloramines are excellent hydroxyaminating reagents. Thus, osmium salt-catalyzed reactions of alkenes with chloramine-T result in β-aminoalcohols—important substrates in the synthesis of natural and pharmaceutically valuable compounds [38]. CO2-induced haloamination of olefins with chlor- [39] and bromamines [40] was reported. The same reagents allow to prepare heterocycles of various size with high diastereoselectivity by intramolecular cyclization of γ-iodoolefins [41].
Dihalogenoamines RSO2NX2 are effective haloaminating reagents for activated alkenes, which do not need additional oxidants, but, unlike similar reactions of chloramines, require, as a rule, the presence of metal salt catalysts [42,43,44]. Dichloroamine-T reacts with α,β-unsaturated ketones in MeCN to give the Ritter-type products, which, under the action of KOH or K2CO3, give imidazolidines [45,46,47] or, with Et3N as the base and CuOTf, aziridines [48].
Iodinanes ArI=NSO2Ar’ are used mainly for aziridination of alkenes [9,49,50,51]. Note that aziridination with imino-λ3-bromane CF3SO2N=BrC6H4CF3 is one of a few methods of synthesis of N-triflylaziridines [15,52,53]. Aziridines can be intermediate products, leading finally to various heterocycles. Thus, 2-alkoxy-3,4-dihydro-2H-pyranes react with PhI=NTs with rearrangement and formation of 5-alkoxypyrrolidines [54]. The diversity of possible pathways of the reaction of sulfonamides and their derivatives with alkenes is schematically shown in Scheme 1.
Most of the methods described above require the use of transition metal catalysts. As an alternative, Minakata et al. proposed a new, metal-free procedure for aziridination of alkenes by commercially available sulfonamides as a nitrogen source in the oxidative system (ButOCl + NaI) in up to 95% yield [55]. The active intermediate is ButOI, formed in situ along with other species as was described in an earlier work [56]. In particular, it was found that amidation of trans-2-octene proceeded with complete stereoselectivity, leading to the trans-substituted aziridine as the sole product; the conclusion was made on the completely stereoselective and stereospecific aziridination, consistent with the reaction mechanism involving a cyclic iodonium intermediate.
The scale of applicability of ButOI-mediated amidation is not limited by the preparation of aziridines, since the formation of other valuable heterocycles was also reported [3,57,58,59,60,61,62,63,64,65]. Thus, non-activated alkenes when treated with benzamides in presence of ButOI afford oxazolines [55]. The authors proposed a tentative mechanism with participation of the in situ generated molecules of ButOI, which react with the amide, rather than with alkene, to give N,N-diiodoamides (Scheme 2) [57]. The formation of the latter intermediates was assumed based on the 1H-NMR analysis, which showed that the signal of the amide protons gradually disappeared and the signal of the tert-butyl group of ButOH appeared within one hour at room temperature. Then, RC(O)NI2 react with alkenes to give cyclic iodonium intermediates which capture the amide counter anion. Note the temperature dependence of halogenation of amides: N,N-dibromotosylamide, generated by the reaction of tosylamide with NBS at 25 °C, was completely transformed to N-bromotosylamide upon heating to 40 °C, as proven by 1H-NMR; as a result, the opposite amidation regioselectivity was observed [66]. This is an illustration of the necessity of both theoretical and experimental investigation of the ButOI-mediated amidation mechanism.
In continuation of our studies on the sulfamidation of alkenes and dienes under oxidative conditions using Minakata’s reagent, ButOCl + NaI, we recently reported on oxidative sulfamidation of vinyl silanes with triflamide and arenesulfonamides, in which the products of iodochlorination, aziridination and heterocyclization to 1,4-azasilinanes were obtained in high total yield [61]. Diverse reactivity in the reactions of oxidative sulfamidation leading to numerous and unexpected products prompted us to perform a quantum chemical study of the mechanism of oxidative amidation of alkenes in the presence of ButOCl + NaI. As the model substrates, styrene (1), trimethyl(vinyl)silane (2), dimethyl(divinyl)silane (3) and diphenyl(divinyl)silane (4) were chosen in the reactions with mesylamide (CH3SO2NH2, a), triflamide (CF3SO2NH2 or TfNH2, b), tosylamide (TolpSO2NH2 or TsNH2, c) and p-nosylamide (NO2C6H4pSO2NH2 or NsNH2, d) were used. The performed study does not purport to answer all questions concerning the mechanism of oxidative amidation, but, rather, to shed light on some unresolved experimental observations.

2. Computational Details

The geometries of all molecules were optimized at the B3LYP/DGDZVP level of theory using the Gaussian 09 software [67]. For all the atoms, the all-electron, double-ζ valence polarized basis set (DGDZVP) was used. It has been demonstrated that the DGDZVP basis set applied for iodine-containing organic molecules provides a much better description of the activation barriers than other double-ζ basis sets, and it even outperformed the triple-ζ basis sets [68]. Due to its small size, it is well-suited to studying halogen bonding in large systems. Since the ButOI-mediated reactions of oxidative sulfamidation were carried out, as a rule, in acetonitrile, all calculations were also performed for MeCN as a solvent, using the integral equation formalism polarizable continuum model (IEF-PCM) [69]. For each stationary point, frequency calculations were performed to confirm whether these structures were local minima or transition states (TS). All transition state calculations were accompanied by intrinsic reaction coordinate (IRC) calculations to verify that each TS connected the corresponding reactants and products. The Møller–Plesset second order perturbation theory (MP2) was used for energy refinement. Throughout the Results and Discussion section, if not stated otherwise, “energy” (taking into account the zero-point vibrational energy (ZPVE) correction) refers to the MP2/DGDZVP//B3LYP/DGDZVP calculations. For comparison, the long range corrected density functional theory (DFT) functional, which includes empirical dispersion correction wB97XD with the same basis set, was also tested.

3. Results and Discussion

3.1. Formation of Reactive Iodine-Containing Species

Since for a successful reaction between alkenes and sulfonamides the reagents, which must be introduced first into the reaction system, need to be ButOCl (5) and NaI, we started with the analysis of the reactions occurring between the two, including the formation of ButOI (6) (Scheme 3).
As follows from Scheme 3, two possible pathways for coordination of NaI to the oxygen atom of ButOCl (5) can occur: (i) via the iodine atom, or (ii) via the sodium atom, leading in the formation of ButOI (56) or ICl (56’), respectively. The former reaction proceeds via the transition state TS5-6 (173i cm–1) lying 28.8 kcal/mol above the reagents (ButOCl + NaI). The energy gain for the formation of ButOI 6 is 10.9 kcal/mol. The 56 equilibrium is shifted towards 6 with ΔG = −10.5 kcal/mol. We could not locate a transition state for the reaction 56’. The forced relaxed stepwise rapprochement of ButOCl and NaI that could lead to the formation of ICl (6’) and ButONa was accompanied by a steady energy increase up to 24.5 kcal/mol. Therefore, the only possible reaction in the system ButOCl + NaI is the formation of ButOI (6), while the formation of iodine chloride 6’ apparently occurs in a different way. The results are summarized in Table 1.

3.2. The Effect of the Sulfonamide Structure

In view of numerous examples illustrating specific behavior of triflamide, distinct from that of non-fluorinated sulfonamides [58,60,70,71], it was of interest to compare the relative iodinating activity of different N-iodinated sulfonamides. We performed such an analysis at the MP2//B3LYP level of theory using as a criterion the ease of the N–I bond rupture in both active intermediates capable of acting as iodinating agents, namely, RSO2NHI (7) and RSO2NI2 (8). For the following reactions:
RSO 2 NHI RSO 2 NH + I +
RSO 2 NI 2 RSO 2 NI + I +
where the relative ZPVE-corrected values of ΔE decrease (and, hence, the iodinating activity increases) in the following order: RSO2 = Ts (16.7) ≥ Ms (16.1) > Ns (12.5) >> Tf (0) kcal/mol for (7), and RSO2 = Ts (14.8) ≥ Ms (14.3) > Ns (11.0) >> Tf (0) kcal/mol for (8). Expectedly, the iodinating activity of RSO2NI2 (8) is slightly (by 1.5–1.9 kcal/mol) higher than that of RSO2NHI (7).
The performed theoretical analysis also allows shedding light on another issue: the dependence of the product composition on the structure of the sulfonamide. Thus, while the reactions of non-fluorinated sulfonamides with various alkenes, including styrene 1, in the system ButOCl + NaI gives rise only to aziridines 9 [55], from the reaction of styrene with triflamide the main product of bis-amidation 10 and a substantial amount (21%) of iodoalcohol PhCH(OH)CH2I 11 were isolated [72] (Scheme 4).
To rationalize this unexpected result, we performed calculations of the two following reactions, focusing not on the absolute values of ΔΔE between the two reactions, which itself does not have much sense, but rather on its variation with substituent R:
RSO 2 NH 2 + Bu t OI RSO 2 NHI + Bu t OH
1 + Bu t OI + H 2 O 11 + Bu t OH
The energetic prevalence of the latter process for the reaction with triflamide is much higher than for non-fluorinated sulfonamides [CF3 (4.3) >> CH3 (1.8) > p-NO2C6H4 (0.7) ≥ p-CH3C6H4 (0) kcal/mol] suggesting that the formation of iodoalcohol is relatively much more favorable with triflamide than with other sulfonamides. This is consistent with its formation in our work [72], and with the literature data on its absence in the reaction with non-fluorinated sulfonamides [55].
No less interesting is a question about different regioselectivity of halogenation of styrene 1 and trimethyl(vinyl)silane 2. While the former gives only the β-iodinated product 11 (Scheme 4), compound 2 affords, apart from the products of bis-amidation 12 and dihalogenation 13, both the α- and β-halogenated products 14, 15 in comparable (although small) yields [61] (Scheme 5).
The calculations give an explicit answer to the question about this apparent discrepancy. The reason is the different structure of the corresponding iodonium cations: almost linear cation in the case of styrene and practically cyclic one in the case of silane 2 as follows from the comparison of the bond lengths and bond angles in Figure 1.
The linear (left) structure in Figure 1 is stabilized as the benzylic type cation, which can add a nucleophile only to the α-carbon affording the β-iodinated product, e.g., 11. The cyclic (right) structure is due to the enhanced electron density on the vinyl group of silane 2 owing to the electronodonor effect of the Me3Si group, and can add the amide anion to either of the two carbon atoms resulting in both α- and β-halogenated products, as was experimentally proven (Scheme 5).
The most challenging difference between the reactivities of non-fluorinated sulfonamides and triflamide is the formation of aziridines from styrene 1 or silane 2 as the only products for the former and the main products of bis-triflamidation of 1, 2, and some other alkenes for triflamide [61,72,73], as shown in Scheme 6 on the example of silane 2.
Similar behavior of substantially different substrates in Scheme 4; Scheme 6 proves the generality of the conclusion about different reactivity of triflamide and its nonfluorinated analogues and requires reasonable explanation. For this, we have calculated the reactions of bis-sulfamidation and aziridination for mesylamide and triflamide on the example of silane 2 and compared the thermodynamics of routes (a) and (b) in Scheme 7.
Due to the same composition of the products in Scheme 7, such an approach allows, without loss of generality, to avoid the analysis of numerous possible transition states and intermediates by comparing the free energies for the two pathways. Because of the high ring strain of the aziridine ring, route (a) is energetically 26.4 kcal/mol preferable over route (b) for R = CH3, and even more preferable (by 29.5 kcal/mol) for R = CF3. The entropy factor acts in the opposite direction and decreases the value of ΔG with respect to ΔE from −26.4 to −9.8 kcal/mol for R = CH3 and from −29.5 to −14.0 kcal/mol for R = CF3. Therefore, the entropy factor also favors route (a), the value of ΔΔG being 4.2 kcal/mol. Such a notable free energy difference explains the experimentally observed formation of linear adducts in the reactions with triflamide [61,72] and of aziridines with non-fluorinated sulfonamides [55].

3.3. Interactions between ButOCl(I), Alkene and Sulfonamide

The next step was to investigate possible interactions of ButOCl(I) and the model alkenes 3 and 4 with various sulfonamides. The approaching of ButOCl (5) to dimethyl(divinyl)silane (3) and diphenyl(divinyl)silane (4) is slightly exothermic by 3.0 and 3.8 kcal/mol, respectively, resulting in the formation of π-complexes with one of the vinyl groups; the formation of similar complexes with ButOI (6) is twice as exothermic as that (5.9 and 7.4 kcal/mol, respectively). However, the Gibbs free energy slightly increases due to the loss of entropy, see Table 2.
The reaction of halogenating agents ButOCl(I) with the model sulfonamides a–d proceeds as shown in Scheme 8, via the transition state TS5-16 or TS6-7 (Figure S1). The transition states TS5-16 corresponding to the formation of N-chlorosulfonamides (16) lie very high in energy, from 59.5 (d) to 64.0 kcal/mol (a), suggesting a highly improbable reaction. Still, the formation of products 16a–d is slightly exothermic, with the energy gain varying from 1.8 (b) to 6.6 kcal/mol (c or d, see Table S1).
In contrast, iodination of sulfonamides a–d with ButOI (6) occurring via the transition state TS6-7 is connected with overcoming a much lower activation barrier from 28.6 (b) to 32.0 kcal/mol (a) with respect to the nonreacting system (5 + sulfonamide a–d), which is about twice as low as that for chlorination through the transition state TS5-16 (Scheme 8, Table S2). Hence, the iodination of sulfonamides with ButOI proceeds much easier than their chlorination with ButOCl, in spite of lower N–I energy bond with respect to N–Cl bond (by ~5 kcal/mol). This means that the ease of halogenation is determined by the ease of the O–Hlg bond rupture, rather than the N–Hlg bond formation. N-Iodosubstituted non-fluorinated sulfonamides 7 a,c,d lie 0.2–2.4 kcal/mol lower, but, for N-iodotriflamide 7b, it lies 2.2 kcal/mol higher with respect to the unreacted system (sulfonamide + ButOI). Further iodination affording N,N-diiodosulfonamides 8 through TS7-8 proceeds by ~7 kcal/mol easier than the first one. The equilibrium in the system (sulfonamide + 2ButOI ⇆ 8 +2ButOH) is slightly shifted to the products, the Gibbs free energy ΔG varying from 0.8 (b) to 3.1 kcal/mol (c, d), and the ZPVE-corrected total energies of the products ΔE being from 0.7 (b) to 4.4 kcal/mol (c) below the reagents (Table S2). Taking into account the fact that ButOI is formed in the reaction of ButOCl with NaI, which is exothermic by 10.9 kcal/mol (Table 1), the relative energy of N,N-diiodosulfonamides 8 must be 2 × 10.9 = 21.8 kcal/mol below the energy of the unreacted system (2ButOCl + 2NaI + sulfonamide).

3.4. Analysis of Reactivity of Sulfonamides, Mono- and Diiodosulfonamides

Some geometric parameters of sulfonamides a–d, N-iodosulfonamides 7, and N,N-diiodosulfonamides 8 as well as some electronic effects in their molecules are presented in Table 3.
As follows from Table 3, with the increase of the iodine atoms number (n) from 0 to 2, the S–N bond in sulfonamides is elongated (weakened). Simultaneously, the pyramidality of the nitrogen atom decreases indicating the increase of the p-character of the nitrogen lone pair from sp2.9 to sp2.5 (R = Me, Ar) or from sp2.5 to sp2.2 (R = CF3). The HOMO–LUMO energy gap narrows from the maximum value of 18.4 eV (R = CF3, n = 0) to the minimum of 9.4 eV (R = p-Tol, n =2), while the LUMO goes down and in N,N-diiodosulfonamides (n = 2), it lies below zero (~ –0.4 eV), being localized on the antibonding iodine atom lone pair. Note that the MO localized on the π-bond of the vinyl group in 3 corresponds to HOMO and lies at –9.97 eV, whereas in 4 it corresponds to HOMO-4 and lies at –10.32 eV. With the increase of the iodine atoms, the electron density on the nitrogen atom decreases, whereas its variations on the iodine atom are negligible.
N-Iodosulfonamides 7 are capable of forming π-complexes with the vinyl group of substrates 3 and 4 with lowering the energy of the system by ~7 kcal/mol as compared to unreacted system [7 + 3 (or 4)] (Table S3). These π-complexes are pre-reaction complexes on the potential energy surface of the reaction of iodosulfamidation (Scheme 9). The transition states of the reaction of iodosulfamidation of 3 (4), TS7-17, lie 23 kcal/mol (20 kcal/mol) above the unreacted system [7 + 3 (or 4)]. The products of iodosulfamidation of 3 or 4 (17 or 17) lie ~45 kcal/mol below the unreacted system. N,N-Diiodosulfonamides 8 demonstrate similar reactivity; however, the reaction with 3 or 4 proceeds much easier because the barrier heights (TS8-18) are lower than those of TS7-17, with the activation energies from 4.7 (c) to 9.5 kcal/mol (b) in the case of TS8-18 and from 1.6 (d) to 9.0 kcal/mol (b) in the case of TS8-18’ (Table S4). The adducts of N,N-diiodosulfonamides with 3 or 4N-(2-(dimethyl(vinyl)silyl)-2-iodoethyl)-N-iodosulfonamides 18 or N-(2-(diphenyl(vinyl)silyl)-2-iodoethyl)-N-iodosulfonamides 18’ lie ~46 kcal/mol below the nonreacting system (8 + 3 or 4).
Adducts 18 or 18’ are key intermediates for further development of the reaction, since they can (i) act as strong iodinating agents for sulfonamides, monoiodosulfonamides and other species in the reaction mixture; (ii) add intramolecularly to the second C=C bond due to the presence of the N-I moiety forming 1,4-azasilinanes; (iii) eliminate molecular iodine to give N-sulfonylaziridines (Scheme 10).
The iodinating properties of 18 are due to a weak N–I bond (<40 kcal/mol) and are manifested, e.g., in the exchange reactions with sulfonamides or N-iodosulfonamides, or with ButOH to recover ButOI. The 1817 reaction proceeds through TS18-17 with a very low activation barrier of only circa 4 kcal/mol in the forward direction in the case of 18b. A similar reaction with 18a has an even smaller barrier, of circa 3 kcal/mol.
On the other hand, the N–I iodine atom in 18 can form intramolecular π-complex with the intact vinyl group with subsequent cyclization to 1,4-azasilinanes 19. The corresponding transition state TS18-19 was also found, Figure 2. The 1819 reaction has activation barrier of ~30 kcal/mol in the forward direction; its driving force is the formation of a stable product of cyclization—(3R,5S)-3,5-diiodo-1-organylsulfonyl-1,4-azasilinanes 19 (twist conformer), which lies 33–38 kcal/mol below adducts 18 depending on substituents R and R1 (Table S5). We could not locate TS18-19 for R = CF3 and R1 = CH3; instead, cyclization occurred with synchronous elimination of CF3SO2. For R1 = Ph, the corresponding TS18-19 was found to lie 50 kcal/mol above 18′b making this transformation impossible. Note that product 19 is formed stereospecifically and, initially, in the twist conformation of the ring, which is unfavorable as compared to “chair” by 5–7 kcal/mol.
Finally, adducts 18 can eliminate molecular iodine to afford aziridines 20 (Scheme 9, Table S5). The reaction is preceded by elimination of the iodine atom from nitrogen in 18 via TS18-20. The 1820 activation barriers are only 6–10 kcal/mol in the forward direction. Again, TS18-20 for R =CF3 could not be located for both R1 = CH3 and Ph. Aziridines 20 lie ~3.8 kcal/mol (in average) higher than 18 at the MP2//B3LYP level of theory, whereas wB97XD, vice versa, predicts it to be 2.8 kcal/mol more favorable (for R = R1 = CH3). The energy diagram for transformation of adducts 18 for R = R1 = CH3 is shown in Figure 3.

4. Conclusions

The analysis of the sulfamidation of different substrates with various sulfonamides in the oxidative system (ButOCl + NaI) was performed at the MP2/DGDZVP//B3LYP/DGDZVP level of theory with the solvent effect taken into account by the IEF-PCM method for MeCN as a solvent. The only energetically favorable process in the aforesaid system is the formation of ButOI acting as an active intermediate generating the mono- and diiodinated sulfonamides RSO2NHI and RSO2NI2 from the reagents. Judged from the ease of the N–I bond rupture with generation of electrophilic I+ in the two species, their iodinating activity increases in the order RSO2: Ts ≥ Ms > Ns >> Tf being 1.5–1.9 kcal/mol higher for RSO2NI2 than for RSO2NHI. The intriguing difference between the reactivity of triflamide forming mainly the products of bis-triflamidation and the non-fluorinated sulfonamides leading to aziridines is found to be due to different thermodynamics of the two routes. The energetic preference of addition over cyclization caused by a high strain of the aziridine ring is by 3.1 kcal/mol larger for triflamide than for mesylamide. Smaller entropy loss for triflamide increases this difference to ΔΔG of 4.2 kcal/mol, thus explaining the experimental observations. Different regioselectivity of oxidative addition to styrene and trimethyl(vinyl)silane is explained by a different structure of the intermediate iodonium cations: a linear one formed from styrene and leading to the β-iodinated product, and a cyclic one formed from vinylsilane and opened by nucleophilic attack of RSO2NH on the sterically less hindered β-carbon atom. The mechanism of the reaction of non-fluorinated sulfonamides RSO2NH2 with divinylsilanes 3 and 4 in the oxidative system (ButOCl + NaI) has been studied in detail. The reaction results in the formation of N-sulfonylaziridines and 1,4-azasilinanes as the kinetic and thermodynamic products, respectively. The adducts of N,N-diiodosulfonamides to one of the vinyl groups of the silane are key intermediates of the reaction.

Supplementary Materials

The following are available online, Figure S1 (structure of TS5-16, TS6-7 for R = CH3.), Tables S1–S5 (relative total and free energies of intermediates of transformations 516, 68, 717(17’), 818(18’), 1819(20) and 1819(20).

Author Contributions

Conceptualization, project administration, funding acquisition, and writing—review and editing, B.A.S. Literature analysis, calculations, validation, and writing—original draft preparation, A.V.K. and M.Y.M. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Beller, M.; Tillack, A.; Seayad, J. Catalytic Amination Reactions of Olefins and Alkynes. Chemin 2005, 36, 10–1002. [Google Scholar] [CrossRef]
  2. Li, G.; Kotti, S.R.S.S.; Timmons, C. Recent Development of Regio- and Stereoselective Aminohalogenation Reaction of Alkenes. Chemin 2007, 38, 2745–2758. [Google Scholar] [CrossRef]
  3. Minakata, S. Utilization of N−X Bonds in The Synthesis of N-Heterocycles. Acc. Chem. Res. 2009, 42, 1172–1182. [Google Scholar] [CrossRef] [PubMed]
  4. Ghorab, M.M.; Ismail, Z.H.; Radwan, A.A.; Abdalla, M. Synthesis and pharmacophore modeling of novel quinazolines bearing a biologically active sulfonamide moiety. Acta Pharm. 2013, 63, 1–18. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  5. Tačić, A.; Nikolić, V.; Nikolić, L.; Savić, I. Antimicrobial sulfonamide drugs. Adv. Technol. 2017, 6, 58–71. [Google Scholar] [CrossRef] [Green Version]
  6. Markov, V.I. Synthesis of substituted 1,2,3-triazoles via N-sulfonylaziridines. ARKIVOC 2006, 2005, 89. [Google Scholar] [CrossRef] [Green Version]
  7. Brunet, J.-J.; Neibecker, D.; Agbossou-Niedercorn, F. Functionalisation of alkenes: Catalytic amination of monoolefins. J. Mol. Catal. 1989, 49, 235–259. [Google Scholar] [CrossRef]
  8. Mansuy, D.; Mahy, J.-P.; Duréault, A.; Bedi, G.; Battioni, P. Iron- and manganese-porphyrin catalysed aziridination of alkenes by tosyl- and acyl-iminoiodobenzene. J. Chem. Soc. Chem. Commun. 1984, 1161. [Google Scholar] [CrossRef]
  9. Evans, D.A.; Bilodeau, M.T.; Faul, M.M. Development of the Copper-Catalyzed Olefin Aziridination Reaction. J. Am. Chem. Soc. 1994, 116, 2742–2753. [Google Scholar] [CrossRef]
  10. Cui, Y.; He, C. Efficient Aziridination of Olefins Catalyzed by a Unique Disilver(I) Compound. J. Am. Chem. Soc. 2003, 125, 16202–16203. [Google Scholar] [CrossRef]
  11. Evans, D.A.; Woerpel, K.A.; Hinman, M.M.; Faul, M.M. Bis(oxazolines) as chiral ligands in metal-catalyzed asymmetric reactions. Catalytic, asymmetric cyclopropanation of olefins. J. Am. Chem. Soc. 1991, 113, 726–728. [Google Scholar] [CrossRef]
  12. Lowenthal, R.E.; Masamune, S. Asymmetric copper-catalyzed cyclopropanation of trisubstituted and unsymmetrical cis-1,2-disubstituted olefins: Modified bis-oxazoline ligands. Tetrahedron Lett. 1991, 32, 7373–7376. [Google Scholar] [CrossRef]
  13. Li, Z.; Conser, K.R.; Jacobsen, E.N. Asymmetric alkene aziridination with readily available chiral diimine-based catalysts. J. Am. Chem. Soc. 1993, 115, 5326–5327. [Google Scholar] [CrossRef]
  14. Li, G.; Wei, H.-X.; Kim, S.H. Unexpected copper-catalyzed aminohalogenation reaction of olefins using N-halo-N-metallo-sulfonamide as the nitrogen and halogen sources. Tetrahedron 2001, 57, 8407–8411. [Google Scholar] [CrossRef]
  15. Ochiai, M.; Kaneaki, T.; Tada, N.; Miyamoto, K.; Chuman, H.; Shiro, M.; Hayashi, S.; Nakanishi, W. A New Type of Imido Group Donor: Synthesis and Characterization of Sulfonylimino-λ3-bromane that Acts as a Nitrenoid in the Aziridination of Olefins at Room Temperature under Metal-Free Conditions. J. Am. Chem. Soc. 2007, 129, 12938–12939. [Google Scholar] [CrossRef] [PubMed]
  16. Li, Z.; Quan, R.W.; Jacobsen, E.N. Mechanism of the (Diimine)copper-Catalyzed Asymmetric Aziridination of Alkenes. Nitrene Transfer via Ligand-Accelerated Catalysis. J. Am. Chem. Soc. 1995, 117, 5889–5890. [Google Scholar] [CrossRef]
  17. Omura, K.; Murakami, M.; Uchida, T.; Irie, R.; Katsuki, T. Enantioselective Aziridination and Amination Using p-Toluenesulfonyl Azide in the Presence of Ru(salen)(CO) Complex. Chem. Lett. 2003, 32, 354–355. [Google Scholar] [CrossRef]
  18. Hemming, K.; Chambers, C.S.; Jamshaid, F.; O’Gorman, P.A. Intramolecular Azide to Alkene Cycloadditions for the Construction of Pyrrolobenzodiazepines and Azetidino-Benzodiazepines. Molecules 2014, 19, 16737–16756. [Google Scholar] [CrossRef] [Green Version]
  19. Ando, T.; Minakata, S.; Ryu, I.; Komatsu, M. Nitrogen atom transfer to alkenes utilizing Chloramine-T as a nitrogen source. Tetrahedron Lett. 1998, 39, 309–312. [Google Scholar] [CrossRef]
  20. Albone, D.P.; Aujla, P.S.; Challenger, S.; Derrick, A.M. A Simple Copper Catalyst for Both Aziridination of Alkenes and Amination of Activated Hydrocarbons with Chloramine-T Trihydrate. J. Org. Chem. 1998, 63, 9569–9571. [Google Scholar] [CrossRef]
  21. Maquestiau, A.; Flammang, R.; Flammang-Barbieux, M.; Mispreuve, H.; Howe, I.; Beynon, J. The gas-phase pinacol rearrangement and related reactions in organic cations generated by chemical ionisation. Tetrahedron 1980, 36, 1993–1998. [Google Scholar] [CrossRef]
  22. Vyas, R.; Gao, G.-Y.; Harden, A.J.D.; Zhang, X.P.; Zhang, X.P. Iron(III) Porphyrin Catalyzed Aziridination of Alkenes with Bromamine-T as Nitrene Source. Org. Lett. 2004, 6, 1907–1910. [Google Scholar] [CrossRef] [PubMed]
  23. Chanda, B.M.; Vyas, R.; Bedekar, A.V. Investigations in the transition metal catalyzed aziridination of olefins, amination, and other insertion reactions with Bromamine-T as the source of nitrene. J. Org. Chem. 2001, 66, 30–34. [Google Scholar] [CrossRef] [PubMed]
  24. Ushakova, I.V.; Shainyan, B.A. Chlorotriflamidation of vinylsilanes with N,N-dichlorotriflamide. Mendeleev Commun. 2020, 30, 117–118. [Google Scholar] [CrossRef]
  25. Ushakova, I.V.; Shainyan, B.A. The reaction of chloroalkyl(vinyl)silanes with N,N-dichlorosulfonamides. Mendeleev Commun. 2020, 30. [Google Scholar] [CrossRef]
  26. Yu, W.Z.; Chen, F.; Cheng, Y.A.; Yeung, Y.-Y. Catalyst-Free and Metal-Free Electrophilic Bromoamidation of Unactivated Olefins Using the N-Bromosuccinimide/Sulfonamide Protocol. J. Org. Chem. 2015, 80, 2815–2821. [Google Scholar] [CrossRef]
  27. Groves, J.T.; Takahashi, T. Activation and transfer of nitrogen from a nitridomanganese(V) porphyrin complex. Aza analog of epoxidation. J. Am. Chem. Soc. 1983, 105, 2073–2074. [Google Scholar] [CrossRef]
  28. Minakata, S.; Ando, T.; Nishimura, M.; Ryu, I.; Komatsu, M. ChemInform Abstract: Novel Asymmetric and Stereospecific Aziridination of Alkenes with a Chiral Nitridomanganese Complex. Chemin 2010, 30, 3392–3394. [Google Scholar] [CrossRef]
  29. Nishimura, M.; Minakata, S.; Thongchant, S.; Ryu, I.; Komatsu, M. Selective [2+1] aziridination of conjugated dienes with a nitridomanganese complex: A new route to alkenylaziridines. Tetrahedron Lett. 2000, 41, 7089–7092. [Google Scholar] [CrossRef]
  30. Nishimura, M.; Minakata, S.; Takahashi, T.; Oderaotoshi, Y.; Komatsu, M. Asymmetric N1 Unit Transfer to Olefins with a Chiral Nitridomanganese Complex: Novel Stereoselective Pathways to Aziridines or Oxazolines. J. Org. Chem. 2002, 67, 2101–2110. [Google Scholar] [CrossRef]
  31. Dauban, P.; Sanière, L.; Tarrade, A.A.; Dodd, R.H. Copper-Catalyzed Nitrogen Transfer Mediated by Iodosylbenzene PhIO. J. Am. Chem. Soc. 2001, 123, 7707–7708. [Google Scholar] [CrossRef]
  32. Guthikonda, K.; Du Bois, J. A Unique and Highly Efficient Method for Catalytic Olefin Aziridination. J. Am. Chem. Soc. 2002, 124, 13672–13673. [Google Scholar] [CrossRef] [PubMed]
  33. Han, H.; Bae, I.; Yoo, E.J.; Lee, J.; Do, Y.; Chang, S. Notable Coordination Effects of 2-Pyridinesulfonamides Leading to Efficient Aziridination and Selective Aziridine Ring Opening. Org. Lett. 2004, 6, 4109–4112. [Google Scholar] [CrossRef] [PubMed]
  34. Catino, A.J.; Nichols, J.M.; Forslund, R.E.; Doyle, M.P. Efficient Aziridination of Olefins Catalyzed by Mixed-Valent Dirhodium(II,III) Caprolactamate. Org. Lett. 2005, 7, 2787–2790. [Google Scholar] [CrossRef] [PubMed]
  35. Thakur, V.V.; Sudalai, A. N-Bromoamides as versatile catalysts for aziridination of olefins using chloramine-T. Tetrahedron Lett. 2003, 44, 989–992. [Google Scholar] [CrossRef]
  36. Jain, S.L.; Sain, B. Aziridination of alkenes using N-iodo-N-potassio-p-toluenesulphonamide as a nitrene precursor. Tetrahedron Lett. 2003, 44, 575–577. [Google Scholar] [CrossRef]
  37. Kong, D.-L.; He, L.-N.; Wang, J.-Q. Facile synthesis of oxazolidinones catalyzed by n-Bu4NBr3/n-Bu4NBr directly from olefins, chloramine-T and carbon dioxide. Catal. Commun. 2010, 11, 992–995. [Google Scholar] [CrossRef]
  38. Fokin, V.V.; Sharpless, K.B. A practical and highly efficient aminohydroxylation of unsaturated carboxylic acids. Angew. Chem. Int. Ed. 2001, 40, 3455–3457. [Google Scholar] [CrossRef]
  39. Minakata, S.; Yoneda, Y.; Oderaotoshi, Y.; Komatsu, M. Unprecedented CO2-Promoted Aminochlorination of Olefins with Chloramine-T. Org. Lett. 2006, 8, 967–969. [Google Scholar] [CrossRef]
  40. Hayakawa, J.; Kuzuhara, M.; Minakata, S. CO2-induced amidobromination of olefins with bromamine-T. Org. Biomol. Chem. 2010, 8, 1424. [Google Scholar] [CrossRef]
  41. Minakata, S.; Kano, D.; Oderaotoshi, A.Y.; Komatsu, M. Unique Ionic Iodine Atom Transfer Cyclization: A New Route to Iodomethylated Pyrrolidine Derivatives from γ-Iodoolefin and Chloramine-T. Org. Lett. 2002, 4, 2097–2099. [Google Scholar] [CrossRef] [PubMed]
  42. Chen, D.; Timmons, C.; Chao, S.; Li, G. Regio- and Stereoselective Copper-Catalyzed Synthesis of Vicinal Haloamino Ketones from α,β-Unsaturated Ketones. Chemin 2004, 35, 3097–3101. [Google Scholar] [CrossRef]
  43. Cai, Y.; Liu, X.; Jiang, J.; Chen, W.; Lin, L.; Feng, X. Catalytic Asymmetric Chloroamination Reaction of α,β-Unsaturated γ-Keto Esters and Chalcones. J. Am. Chem. Soc. 2011, 133, 5636–5639. [Google Scholar] [CrossRef] [PubMed]
  44. Cai, Y.; Liu, X.; Zhou, P.; Kuang, Y.; Lin, L.; Feng, X. Iron-catalyzed asymmetric haloamination reactions. Chem. Commun. 2013, 49, 8054–8056. [Google Scholar] [CrossRef] [PubMed]
  45. Wei, H.-X.; Kim, S.H.; Li, G. Electrophilic Diamination of Alkenes by Using FeCl3−PPh3Complex as the Catalyst. J. Org. Chem. 2002, 67, 4777–4781. [Google Scholar] [CrossRef] [PubMed]
  46. Timmons, C.; McPherson, L.; Chen, D.A.; Wei, H.-X.; Li, G. Manganese (IV) oxide-catalyzed electrophilic diamination of electron-deficient alkenes provides an easy synthesis of alpha,beta-diamino acid and ketone derivatives for peptidomimetic studies. J. Pept. Res. 2005, 66, 249–254. [Google Scholar] [CrossRef]
  47. Wu, H.; Ji, X.; Sun, H.; An, G.; Han, J.; Li, G.; Pan, Y. Organocatalyzed regio- and stereoselective diamination of functionalized alkenes. Tetrahedron 2010, 66, 4555–4559. [Google Scholar] [CrossRef]
  48. Li, G.; Chen, D.; Timmons, C.; Guo, L.; Xu, X. One-Pot Stereoselective Synthesis ofanti3-Alkyl and 3-Aryl-N-p-tosyl-aziridine-2-ketones and 3-Aryl-N-p-tosyl-aziridine-2-carboxylates. Synthesis 2004, 2004, 2479–2484. [Google Scholar] [CrossRef]
  49. Cho, D.-J.; Jeon, S.-J.; Kim, H.-S.; Kim, T.-J. Catalytic Cyclopropanation and Aziridination of Alkenes by a Cu(I) Complex of Ferrocenyldiimine. Synlett 1998, 1998, 617–618. [Google Scholar] [CrossRef]
  50. Amisial, L.D.; Dai, X.; Kinney, R.A.; Krishnaswamy, A.; Warren, T.H. Cu(I) β-Diketiminates for Alkene Aziridination: Reversible CuArene Binding and Catalytic Nitrene Transfer from PhINTs. Inorg. Chem. 2004, 43, 6537–6539. [Google Scholar] [CrossRef]
  51. Jeon, H.-J.; Nguyen, S.T. ChemInform Abstract: Nitrene-Transfer to Olefins Catalyzed by Methyltrioxorhenium: A Universal Catalyst for the [1 + 2] Cycloaddition of C-, N-, and O-Atom Fragments to Olefins. Chemin 2001, 32, 235–236. [Google Scholar] [CrossRef]
  52. Hoque, M.; Miyamoto, K.; Tada, N.; Shiro, M.; Ochiai, M. Synthesis and Structure of Hypervalent Diacetoxybromobenzene and Aziridination of Olefins with Imino-λ3-bromane Generated in Situ under Metal-Free Conditions. Org. Lett. 2011, 13, 5428–5431. [Google Scholar] [CrossRef] [PubMed]
  53. Ochiai, M.; Naito, M.; Miyamoto, K.; Hayashi, S.; Nakanishi, W. Imination of Sulfides and Sulfoxides with Sulfonylimino-λ3-Bromane under Mild, Metal-Free Conditions. Chem. A Eur. J. 2010, 16, 8713–8718. [Google Scholar] [CrossRef]
  54. Armstrong, A.; Cumming, G.R.; Pike, K. Aminative Rearrangement of 2-Alkoxy-3,4-dihydro-2H-pyrans: A Novel Stereocontrolled Route to Substituted Pyrrolidines. Chemin 2004, 35, 812–813. [Google Scholar] [CrossRef]
  55. Minakata, S.; Morino, Y.; Oderaotoshi, Y.; Komatsu, M. Novel aziridination of olefins: Direct synthesis from sulfonamides using t-BuOI. Chem. Commun. 2006, 3337–3339. [Google Scholar] [CrossRef] [PubMed]
  56. Tanner, D.D.; Gidley, G.C.; Das, N.; Rowe, J.E.; Potter, A. On the structure of tert-butyl hypoiodite. J. Am. Chem. Soc. 1984, 106, 5261–5267. [Google Scholar] [CrossRef]
  57. Minakata, S.; Morino, Y.; Ide, T.; Oderaotoshi, Y.; Komatsu, M. Direct synthesis of oxazolines from olefins and amides using t-BuOI. Chem. Commun. 2007, 3279–3281. [Google Scholar] [CrossRef]
  58. Shainyan, B.A.; Moskalik, M.Y.; Astakhova, V.V.; Schilde, U. Novel design of 3,8-diazabicyclo[3.2.1]octane framework in oxidative sulfonamidation of 1,5-hexadiene. Tetrahedron 2014, 70, 4547–4551. [Google Scholar] [CrossRef]
  59. Shainyan, B.A.; Moskalik, M.Y.; Astakhova, V.V.; Sterkhova, I.V.; Ushakov, I.A. Oxidative addition of trifluoroacetamide to alkenes, 2,5-dimethylhexa-2,4-diene and conjugated cyclic dienes. Tetrahedron 2015, 71, 8669–8675. [Google Scholar] [CrossRef]
  60. Moskalik, M.Y.; Astakhova, V.V.; Schilde, U.; Sterkhova, I.V.; Shainyan, B.A. Assembling of 3,6-diazabicyclo[3.1.0]hexane framework in oxidative triflamidation of substituted buta-1,3-dienes. Tetrahedron 2014, 70, 8636–8641. [Google Scholar] [CrossRef]
  61. Astakhova, V.V.; Shainyan, B.A.; Moskalik, M.Y.; Sterkhova, I.V. Oxidative sulfamidation of vinyl silanes: A route to diverse silylated N-Heterocycles. Tetrahedron 2019, 75, 4531–4541. [Google Scholar] [CrossRef]
  62. Shainyan, B.A. Unsaturated Derivatives of Trifluoromethanesulfonamide. Eur. J. Org. Chem. 2018, 2018, 3594–3608. [Google Scholar] [CrossRef]
  63. 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]
  64. 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]
  65. 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]
  66. Yu, W.Z.; Cheng, Y.A.; Wong, M.W.; Yeung, Y.-Y. Atmosphere- and Temperature-Controlled Regioselective Aminobromination of Olefins. Adv. Synth. Catal. 2017, 359, 234–239. [Google Scholar] [CrossRef]
  67. Frisch, M.J.; Trucks, G.W.; Schlegel, H.B.; Scuseria, G.E.; Robb, M.A.; Cheeseman, J.R. ; Scalmani, G. ; Barone, V. ; Mennucci, B. ; Petersson, G.A.; Nakatsuji, H.; Caricato, M.; Li, X.; Hratchian, H.P.; Izmaylov, A.F.; Bloino, J.; Zheng, G.; Sonnenberg, J.L.; Hada, M.; Ehara, M.; Toyota, K.; Fukuda, R.; Hasegawa, J.; Ishida, M.; Nakajima, T.; Honda, Y.; Kitao, O.; Nakai, H.; Vreven, T.; Montgomery, J.A.; Peralta, J.E.; Ogliaro, F.; Bearpark, M.; Heyd, J.J.; Brothers, E.; Kudin, K.N.; Staroverov, V.N.; Kobayashi, R.; Normand, J.; Raghavachari, K.; Rendell, A.; Burant, J.C.; Iyengar, S.S.; Tomasi, J.; Cossi, M.; Rega, N.; Millam, J.M.; Klene, M.; Knox, J.E.; Cross, J.B.; Bakken, V.; Adamo, C.; Jaramillo, J.; Gomperts, R.; Stratmann, R.E.; Yazyev, O.; Austin, A.J.; Cammi, R.; Pomelli, C.; Ochterski, J.W.; Martin, R.L.; Morokuma, K.; Zakrzewski, V.G.; Voth, G.A.; Salvador, P.; Dannenberg, J.J.; Dapprich, S.; Daniels, A.D.; Farkas, O.; Foresman, J.B.; Ortiz, J.V.; Cioslowski, J.; Fox, D.J. Gaussian 09, Revision, A.02;, Wallingford, CT, 2009.
  68. Poleshchuk, O.K.; Yureva, A.; Filimonov, V.D.; Frenking, G. Study of a surface of the potential energy for processes of alkanes free-radical iodination by B3LYP/DGDZVP method. J. Mol. Struct. THEOCHEM 2009, 912, 67–72. [Google Scholar] [CrossRef]
  69. Tomasi, J.; Mennucci, B.; Cancès, E. The IEF version of the PCM solvation method: An overview of a new method addressed to study molecular solutes at the QM ab initio level. J. Mol. Struct. THEOCHEM 1999, 464, 211–226. [Google Scholar] [CrossRef]
  70. Shainyan, B.A.; Tolstikova, L.L. Trifluoromethanesulfonamides and Related Compounds. Chem. Rev. 2012, 113, 699–733. [Google Scholar] [CrossRef]
  71. 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]
  72. Shainyan, B.A.; Moskalik, M.Y.; Starke, I.; Schilde, U. Formation of unexpected products in the attempted aziridination of styrene with trifluoromethanesulfonyl nitrene. Tetrahedron 2010, 66, 8383–8386. [Google Scholar] [CrossRef]
  73. Moskalik, M.Y.; Shainyan, B.A.; Schilde, U. Reaction of trifluoromethanesulfonamide with alkenes and cycloocta-1,5-diene under oxidative conditions. Direct assembly of 9-heterobicyclo[4.2.1]nonanes. Russ. J. Org. Chem. 2011, 47, 1271–1277. [Google Scholar] [CrossRef]
Scheme 1. Versatile reaction pathways of alkenes with sulfonamide derivatives.
Scheme 1. Versatile reaction pathways of alkenes with sulfonamide derivatives.
Molecules 25 04877 sch001
Scheme 2. Tentative mechanism of formation of iodonium cations by the reaction of alkenes with N,N-diiodocarboxamides or -sulfonamides as the key intermediates.
Scheme 2. Tentative mechanism of formation of iodonium cations by the reaction of alkenes with N,N-diiodocarboxamides or -sulfonamides as the key intermediates.
Molecules 25 04877 sch002
Scheme 3. Possible transformations in the system ButOCl + NaI.
Scheme 3. Possible transformations in the system ButOCl + NaI.
Molecules 25 04877 sch003
Scheme 4. Different courses of the reaction of styrene with triflamide and non-fluorinated sulfonamides. Conditions: (ButOCl + NaI), MeCN, −10 °C.
Scheme 4. Different courses of the reaction of styrene with triflamide and non-fluorinated sulfonamides. Conditions: (ButOCl + NaI), MeCN, −10 °C.
Molecules 25 04877 sch004
Scheme 5. Different regioselectivity of halotriflamidation of trimethyl(vinyl)silane 2.
Scheme 5. Different regioselectivity of halotriflamidation of trimethyl(vinyl)silane 2.
Molecules 25 04877 sch005
Figure 1. Linear (a) and cyclic (b) iodonium cations from electrophilic iodination of styrene 1 and trimethyl(vinyl)silane 2.
Figure 1. Linear (a) and cyclic (b) iodonium cations from electrophilic iodination of styrene 1 and trimethyl(vinyl)silane 2.
Molecules 25 04877 g001
Scheme 6. Different courses of the reaction of trimethylvinylsilane with triflamide and non-fluorinated sulfonamides.
Scheme 6. Different courses of the reaction of trimethylvinylsilane with triflamide and non-fluorinated sulfonamides.
Molecules 25 04877 sch006
Scheme 7. Bis-sulfamidation versus aziridination of trimethyl(vinyl)silane.
Scheme 7. Bis-sulfamidation versus aziridination of trimethyl(vinyl)silane.
Molecules 25 04877 sch007
Scheme 8. Halogenation of sulfonamides with ButOI(Cl).
Scheme 8. Halogenation of sulfonamides with ButOI(Cl).
Molecules 25 04877 sch008
Scheme 9. Formation of the products of iodosulfamidation by the reaction of N-iodo- (7) and N,N-diiodosulfonamides (8) with dimethyldivinyl- (3) and diphenyldivinylsilane (4).
Scheme 9. Formation of the products of iodosulfamidation by the reaction of N-iodo- (7) and N,N-diiodosulfonamides (8) with dimethyldivinyl- (3) and diphenyldivinylsilane (4).
Molecules 25 04877 sch009
Scheme 10. Possible transformations of N-(2-((vinyl)silyl)-2-iodoethyl)-N-iodosulfonamides 18.
Scheme 10. Possible transformations of N-(2-((vinyl)silyl)-2-iodoethyl)-N-iodosulfonamides 18.
Molecules 25 04877 sch010
Figure 2. The structure of the transition state TS18-19, R = R1 = CH3.
Figure 2. The structure of the transition state TS18-19, R = R1 = CH3.
Molecules 25 04877 g002
Figure 3. Energy diagram for conversion of adduct 18 to 1,4-azasilinane 19 and aziridine 20. R = R1 = CH3 (refer to Scheme 10 for details).
Figure 3. Energy diagram for conversion of adduct 18 to 1,4-azasilinane 19 and aziridine 20. R = R1 = CH3 (refer to Scheme 10 for details).
Molecules 25 04877 g003
Table 1. Relative energies (kcal/mol) of molecules 6 and 6’ formed in the system ButOCl + NaI.
Table 1. Relative energies (kcal/mol) of molecules 6 and 6’ formed in the system ButOCl + NaI.
Parameter5TS5-666’
MP2/DGDZVP//B3LYP/DGDZVP
ΔE+ZPVE0.028.8−10.924.5
ΔG0.038.8−10.523.6
wB97XD/DGDZVP
ΔE+ZPVE0.025.4−13.020.0
ΔG0.034.3−12.819.4
Table 2. Relative energies (kcal/mol) of π-complexes of divinylsilanes 3, 4 with t-BuOCl 5 and t-BuOI 6.
Table 2. Relative energies (kcal/mol) of π-complexes of divinylsilanes 3, 4 with t-BuOCl 5 and t-BuOI 6.
Parameter5 + 3 or 45···35···46 + 3 or 46···36···4
MP2/DGDZVP//B3LYP/DGDZVP
ΔE+ZPVE0.0−3.0−3.8−10.9−16.8−18.3
ΔG0.06.95.0−10.5−6.0−7.7
wB97XD/DGDZVP
ΔE+ZPVE0.0−13.0−19.2−21.2
ΔG0.0−12.8−8.6−8.3
Table 3. Geometric and electronic parameters of sulfonamides, mono- and diiodosulfonamides RSO2X, X = NH2–nIn (MP2/DGDZVP//B3LYP/D GDZVP). HOMO (highest occupied molecular orbital), LUMO (lowest unoccupied molecular orbital), CM5 (charge model 5), NBO (natural bond order), LP (lone pair).
Table 3. Geometric and electronic parameters of sulfonamides, mono- and diiodosulfonamides RSO2X, X = NH2–nIn (MP2/DGDZVP//B3LYP/D GDZVP). HOMO (highest occupied molecular orbital), LUMO (lowest unoccupied molecular orbital), CM5 (charge model 5), NBO (natural bond order), LP (lone pair).
ParameterR = CH3R = CF3R = p-TolR = p-NO2C6H4
X= NH2NHINI2NH2NHINI2NH2NHINI2NH2NHINI2
S–N, Å1.6781.7071.7451.6461.6591.6791.6831.7171.7611.6731.7021.740
N–I, Å2.1152.1362.1032.1212.1162.1392.1112.134
∠SNI(H)I(H)126.5126.6137.0135.1141.0155.3124.9123.7135.5127.7127.4138.6
ΣN335.5337.3344.9343.3348.6355.2333.8334.6343.7336.6338.1346.0
HOMO, eV−12.720−10.480−10.242−13.453−10.685−10.382−9.734−9.835−9.881−10.375−10.349 −10.289
LUMO, eV5.1410.644−0.4754.9460.317−0.7352.1610.694−0.4350.1970.116−0.545
LUMO–HOMO gap, eV17.86111.1249.76718.39811.0019.64711.89510.5299.44910.57110.4669.744
CM5 charge on N−0.600−0.503−0.412−0.574−0.501−0.428−0.601−0.502−0.407−0.593−0.501−0.412
CM5 charge on I0.2230.2230.2790.2720.2080.2120.2300.232
Electrophilicity (I)1.8661.6002.1681.8201.8191.5381.9861.568
NBO (LPN occupancy)1.9001.9091.9181.8661.8671.8671.8971.9081.9181.8881.8981.907
Sample Availability: Samples of the compounds are not available from the authors.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Kuzmin, A.V.; Moskalik, M.Y.; Shainyan, B.A. Trifluoromethanesulfonamide vs. Non-Fluorinated Sulfonamides in Oxidative Sulfamidation of the C=C Bond: An In Silico Study. Molecules 2020, 25, 4877. https://doi.org/10.3390/molecules25214877

AMA Style

Kuzmin AV, Moskalik MY, Shainyan BA. Trifluoromethanesulfonamide vs. Non-Fluorinated Sulfonamides in Oxidative Sulfamidation of the C=C Bond: An In Silico Study. Molecules. 2020; 25(21):4877. https://doi.org/10.3390/molecules25214877

Chicago/Turabian Style

Kuzmin, Anton V., Mikhail Yu. Moskalik, and Bagrat A. Shainyan. 2020. "Trifluoromethanesulfonamide vs. Non-Fluorinated Sulfonamides in Oxidative Sulfamidation of the C=C Bond: An In Silico Study" Molecules 25, no. 21: 4877. https://doi.org/10.3390/molecules25214877

Article Metrics

Back to TopTop