Next Article in Journal
Pressurized-Liquid Extraction as an Efficient Method for Valorization of Thymus serpyllum Herbal Dust towards Sustainable Production of Antioxidants
Next Article in Special Issue
Synthesis of the [11]Cyclacene Framework by Repetitive Diels–Alder Cycloadditions
Previous Article in Journal
Effects of Essential Oils of Elettaria cardamomum Grown in India and Guatemala on Gram-Negative Bacteria and Gastrointestinal Disorders
Previous Article in Special Issue
Maleimide Self-Reaction in Furan/Maleimide-Based Reversibly Crosslinked Polyketones: Processing Limitation or Potential Advantage?
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Hetero-Diels-Alder Reactions of In Situ-Generated Azoalkenes with Thioketones; Experimental and Theoretical Studies †

by
Grzegorz Mlostoń
1,*,
Katarzyna Urbaniak
1,
Malwina Sobiecka
1,‡,
Heinz Heimgartner
2,
Ernst-Ulrich Würthwein
3,*,
Reinhold Zimmer
4,
Dieter Lentz
4 and
Hans-Ulrich Reissig
4,*
1
Department of Organic and Applied Chemistry, Faculty of Chemistry, University of Lodz, 12 Tamka Street, 91-403 Lodz, Poland
2
Department of Chemistry, University of Zurich, Winterthurerstrasse 190, CH-8057 Zurich, Switzerland
3
Organisch-Chemisches Institut and Center for Multiscale Theory and Computation (CMTC), Westfälische Wilhelms-Universität Münster, Corrensstrasse 40, 48149 Münster, Germany
4
Institut für Chemie und Biochemie, Freie Universität Berlin, Takustrasse 3, 14195 Berlin, Germany
*
Authors to whom correspondence should be addressed.
Dedicated to Professor Mieczysław Mąkosza (Warsaw) in recognition of his outstanding achievements in organic chemistry.
Part of Master Thesis, Sobiecka, M. Hetero-Diels-Alder reactions with exploration of α-azoethylenes as heterodienes and thioketones as heterodienophiles. University of Łódź, 2020, Łódź, Poland.
Molecules 2021, 26(9), 2544; https://doi.org/10.3390/molecules26092544
Submission received: 1 April 2021 / Revised: 20 April 2021 / Accepted: 22 April 2021 / Published: 27 April 2021
(This article belongs to the Special Issue Diels-Alder Reaction in Organic Synthesis)

Abstract

:
The hetero-Diels-Alder reactions of in situ-generated azoalkenes with thioketones were shown to offer a straightforward method for an efficient and regioselective synthesis of scarcely known N-substituted 1,3,4-thiadiazine derivatives. The scope of the method was fairly broad, allowing the use of a series of aryl-, ferrocenyl-, and alkyl-substituted thioketones. However, in the case of N-tosyl-substituted cycloadducts derived from 1-thioxo-2,2,4,4-tetramethylcyclobutan-3-one and the structurally analogous 1,3-dithione, a more complicated pathway was observed. By elimination of toluene sulfinic acid, the initially formed cycloadducts afforded 2H-1,3,4-thiadiazines as final products. Advanced DFT calculations revealed that the observed high regioselectivity was due to kinetic reaction control and that the (4 + 2)-cycloadditions of sterically less unhindered thiones occurred via highly unsymmetric transition states with shorter C..S-distances (2.27–2.58 Å) and longer N..C-distances (3.02–3.57 Å). In the extreme case of the sterically very hindered 2,2,4,4-tetramethylcyclobutan-1,3-dione-derived thioketones, a zwitterionic intermediate with a fully formed C‒S bond was detected, which underwent ring closure to the 1,3,4-thiadiazine derivative in a second step. For the hypothetical formation of the regioisomeric 1,2,3-thiadiazine derivatives, the DFT calculations proposed more symmetric transition states with considerably higher energies.

Graphical Abstract

1. Introduction

In the most recent three decades, the poor reputation of thioketones (Figure 1) as smelly and unstable sulfur analogues of ketones changed dramatically and currently, many of them belonging to aromatic 1ad, ferrocenyl 1eh, or cycloaliphatic 1ik representatives are considered to be superior reagents for cycloaddition chemistry and related applications (Figure 1) [1]. Recently, thiochalcones, considered α,β-unsaturated analogues of thioketones, e.g., Figure 1 (1l), were demonstrated to act as active C=S dieno- and dipolarophiles [2,3,4].
Based on the results of kinetic studies, R. Huisgen named them first ‘superdipolarophiles’ [5] and almost at the same time, his former PhD student J. Sauer coined the term ‘superdienophiles’ for thioketones [6]. In our very recent publications, diverse thioketones were demonstrated to act as reactive building blocks yielding five-membered S-heterocycles (or products of their secondary conversions) not only with typical 1,3-dipoles such as thiocarbonyl S-methanides [7,8], diazo alkanes [9,10,11], nitrile imines [12,13], nitrile oxides [14], etc., but also with donor-acceptor cyclopropanes in the presence of an activating Lewis acid [15]. Moreover, trienamine-mediated asymmetric hetero-Diels-Alder reactions of thioketones leading to optically active 4H-thiopyran derivatives were described [16]. In numerous cases, mechanistic studies demonstrated that both (3 + 2)- as well as (4 + 2)-cycloadditions with aromatic and ferrocenyl thioketones do not follow the standard concerted pathways [17,18] but they occur via step-wise mechanisms involving diradicaloid or zwitterionic intermediates [7,9,19,20,21].
In a very recent study from our laboratory, some thioketones, such as 1a,1e,1k (Scheme 1) as well as thiochalcone 1l (Scheme 1), were shown to behave as reactive hetero-dienophiles towards the in situ-generated α-nitrosostyrene 2 (Scheme 1), and the desired 1,5,2-oxathiazines Scheme 1 (3ad) were obtained under mild conditions in a regioselective manner, in good to high yields (Scheme 1) [3,22].
Azoalkenes (1,2-diaza-1,3-dienes) 4 (Scheme 2) are structurally and electronically related to 1-oxa-2-aza-1,3-dienes 2 (Scheme 2) and currently attract great attention as versatile substrates for hetero-Diels-Alder reactions performed with various dienophiles. The importance of this research is reflected by a large number of excellent review articles, which appeared in the last two decades [23,24,25,26,27]. Numerous original publications appear regularly. In addition, very recent reports focused on asymmetric cycloaddition reactions with this type of hetero-dienes and a few of them can be cited as representative examples [28,29,30,31].
In spite of this fact, the (4 + 2)-cycloadditions of azoalkenes 4 (Scheme 2) with ‘superdienophilic’ thioketones (Scheme 1 (1)) are practically unknown. However, a single, brief report by B. Bonini et al. on the hetero-Diels-Alder reaction of thiofluorenone (Scheme 2 (1b)) with some stable, isolable azoalkenes (Scheme 2 (4a,b)) appeared in 1981. These reactions were performed at room temperature in benzene solution and regioselectively led to 3,6-dihydro-2H-1,3,4-thiadiazines (Scheme 2 (5a,b)) in high to fair yields (Scheme 2) [32]. Notably, thiofluorenone S-oxide (sulfine), derived from the studied aromatic thioketone, reacted with Scheme 2 (4a) and Scheme 2 (4b) to give the corresponding S-oxides Scheme 2 (6c) and Scheme 2 (6d), respectively, derived from isomeric 1,2,3-thiadiazines as major products, however (Scheme 2, below). Thus, the (4 + 2)-cycloaddition onto the C=S(O) bond of these dienophiles occurred with inversed regioselectivity. Two other aromatic thioketones derived from thiobenzophenone (Scheme 1 (1a)), bearing two 4-MeC6H4 and 4-O2NC6H4 groups, respectively, failed to react with the tested azoalkenes (Scheme 2 (4a,b)). Therefore, the synthetic utility of this reported method seems to be very limited.
The goal of the present study was the examination of hetero-Diels-Alder reactions of in situ-generated, highly reactive azoalkenes (Scheme 3 (7a,b)) with a larger series of aromatic, ferrocenyl, and cycloaliphatic thioketones (Scheme 1 (1)). The regioselectivity of these cycloadditions was of particular interest since either the scarcely known 1,3,4-thiadiazines (Scheme 3 (9)) or the similarly rare 1,2,3-thiadiazines (Scheme 3 (10)) can be formed (Scheme 3). A careful study of typical examples of these (4 + 2)-cycloadditions by appropriate computational methods should help to understand the involved reaction mechanism(s) and the observed regioselectivity.

2. Results and Discussion

2.1. Experimental Work

The reactive azoalkenes (Scheme 3 (7a,7b)) were generated in situ from hydrazones (Scheme 3 (8a,8b)) derived from α-chloroacetophenone and p-toluenesulfonyl hydrazide or benzoylhydrazine, respectively, following a well-known and widely applied protocol [33]. In both cases, elimination of hydrogen chloride, leading to the required azoalkene, occurs smoothly at room temperature in the presence of potassium carbonate in CH2Cl2 solution (Scheme 3).
As a typical reaction with thiobenzophenone (1a) and hydrazone (8a), the solution was stirred in the presence of potassium carbonate overnight, and after that time, the characteristic blue color of 1a disappeared. After filtration and evaporation of the solvent, the 1H-NMR spectrum of the crude mixture revealed the presence of a characteristic singlet located at 3.21 ppm, which was attributed to the CH2 group present in cycloadduct (9a), formed as the sole product. After chromatographic purification, colorless crystals were isolated and identified as compound 9a (Scheme 4, 54% yield). The same reaction carried out in THF using Cs2CO3 as a base for deprotonation of (8a) led to the same product in 45% yield, only. However, comparison of both procedures in two parallel performed experiments showed that the reaction carried out in CH2Cl2 was more efficient (shorter reaction time, better yield) and for that reason, further experiments were performed in CH2Cl2 using K2CO3 as an inexpensive and easily available base. The structure of 9a was confirmed by further spectroscopic data and the most important indication was offered by signals in the 13C-NMR spectrum at 143.7 ppm, which was assigned to the C atom of the C=N group, and at 24.4 ppm attributed to the 6-CH2 unit. In addition, a less intense but characteristic signal of the C-2 atom was found at 77.7 ppm. The elemental analysis confirmed the molecular formula of 9a as C28H24N2O2S2. Interestingly, while heating in a capillary, the colorless 9a turned blue, indicating a retro-cycloaddition reaction leading to the extrusion of thiobenzophenone (1a).
An analogous course of the reaction was observed when hydrazone 8b served as the precursor of azoalkene 7b, which was trapped in situ by (1a), and cycloadduct (9b) was obtained as the sole product. After chromatographic purification, the yield of the latter compound was determined to be 59%. There was no evidence that regioisomeric 1,2,3-thiadiazine derivatives 10a or 10b were formed in these experiments.
In analogy to 1a, the (4 + 2)-cycloadducts 9ch were obtained as crystalline products from aromatic thioketones 1bd using 8a or 8b as precursors of the in situ-generated azoalkenes 7a and 7b, respectively. Finally, the molecular structures of 9c and 9g were unambiguously confirmed by X-ray analyses (Figure 2).
Quite unexpectedly, two other aromatic thioketones, namely 2,3,6,7-dibenzocycloheptene-1-thione (1m) and its 10,11-dihydro analogue (1n), which have been applied in our earlier studies focused on hetero-Diels-Alder reactions [35], failed to react with 7a even after 3 days at rt and no discoloration of the blue solutions was observed after that time.
Ferrocenyl-substituted thioketones 1eh reacted similarly to aromatic representatives, and the reactions carried out at rt in CH2Cl2 with 8a, applied as a precursor of the reactive azoalkene 7a, led to the ferrocenyl substituted 1,3,4-thiadiazine derivatives (9i,k,m,o) in 39–56% yield (Scheme 5). A similar course of the studied reactions, regioselectively leading to cycloadducts 9j,l,n, and 9p, was observed with the N-benzoyl functionalized azoalkene 7b. In this series, the lower yields of isolated products 9 can be explained by the fact, that along with the desired (4 + 2)-cycloadducts, substantial amounts (40–45%) of the corresponding ferrocenyl ketones were formed as side products and isolated during chromatographic separation of the crude mixtures as more polar fractions. However, the detailed mechanism of their formation under the applied reaction conditions (simple hydrolysis or oxidation of the C=S group) is currently unknown. In accordance with the postulated structure of heterocycles 9ip, the 6-CH2 group (the H atoms present in this group are diastereotopic in products 9ip appeared in the 1H-NMR spectra in all cases as AB-system in the region of 3.08–3.67 ppm with a coupling constant of approximately 2JH,H = 17 Hz.
In a recent publication from our group, a possible non-concerted pathway of hetero-Diels-Alder reactions of aryl/ferrocenyl thioketones with a trienamine was discussed [16]. In the light of this discussion, the reaction of azoalkene 7b with the cyclopropyl ferrocenyl thioketone 1h, which is used for the first time in the herein studied hetero-Diels-Alder reaction, deserves a brief comment. The cyclopropyl substituent is known to act as so called ‘radical clock’ [36] and, therefore, cyclopropyl-substituted dienophiles found application in mechanistic studies aimed at the detection of a postulated diradicaloid intermediate [37,38] in the course of a non-concerted cycloaddition reaction. Isolation of 9o as the sole product of the studied reaction of 1h with 8b suggests that its formation results from the concerted hetero-Diels-Alder reaction. Notably, diferrocenyl thioketone [39], which gave the desired (4 + 2)-cycloadduct in the earlier reported reaction with α-nitrosostyrene in a good yield of 75% [22], failed to react with both azoalkenes 7a and 7b. It is very likely that steric hindrance was the reason for the negative results in both cases.
The reactions of the cycloaliphatic thioketones 1ik with azoalkenes 7a and 7b were performed analogously following the standard, overnight procedure. After a typical work-up, the 1H-NMR analysis of the crude reaction mixture obtained from 1i and 7a revealed the presence of two products with characteristic singlets localized at 3.50 and 6.54 ppm, respectively. Whereas one of them (the high-field shifted signal) might be attributed to the 6-CH2 group of (4 + 2)-cycloadduct 9p (Scheme 5), the second, low-field shifted singlet suggested formation of a new product with a hitherto unknown structure. After column chromatography on silica gel, the product showing the high-field signal completely disappeared and the isolated material consisted exclusively of yellowish crystals with mp 93–95 °C. Notably, after its isolation, a substantial reduction of the mass was observed. Further spectroscopic data were required to solve the structure of this compound. In the 1H-NMR spectrum, neither signals characteristic for the tosyl moiety could be found nor those for the 6-CH2 group of 3,6-dihydro-2H-1,3,4-thiadiazine derivatives of type 9. Instead, a new singlet at 6.54 ppm suggested the presence of an olefinic fragment. On the other hand, the 13C-NMR spectrum showed, along with other absorptions characteristic for the Ph ring at 125.0 (2CH), 128.8 (2CH), 129.1 (1CH), and 134.7 (Car) ppm, two signals localized at 107.4 and 151.8 ppm. In accordance with the 1H-NMR data, they were attributed to two C atoms of the alkene fragment –S–CH=C(Ph)–N=. In addition, the signal of the carbonyl group was found in the 13C-NMR spectrum at 217.6 ppm. This information allowed to postulate the structure of the isolated product as 2H-1,3,4-thiadiazine derivative (11a) (Scheme 6), which was supported by combustion analysis confirming the molecular formula C16H18N2OS. We assume that the elimination was favorable in these two examples since it relieved steric compression exerted by the close proximity of the N-tosyl group and the four methyl groups. Compounds with a 2H-1,3,4-thiadiazine substructure are very rare and only derivatives anellated to arene rings are reported in literature. Compounds such as the orange-colored 11a can be regarded as cis-configured azoalkenes, which is in conjugation with a thioether moiety. The UV/Vis spectrum of 11a in dichloromethane shows a strong absorption at 239 nm (log ε = 4.1 with a shoulder at 283 nm) and a weaker absorption at 358 nm (log ε = 2.8 with at 430 nm).
The analogous structure 11b was attributed to the product isolated after an attempted hetero-Diels-Alder reaction of 1j with 7a (molar ratio of 1j:7a = 1:1) (Scheme 6). The presence of the C=S group was proven by the characteristic orange-red color of the product and an absorption in the 13C-NMR spectrum at 277.4 ppm. Unexpectedly, the same 1:1 product 11b was obtained in another experiment aimed at the synthesis of the bis-cycloadduct, starting with a molar ratio of 1j:7a = 1:2. Apparently, steric hindrance in the initially formed cycloadduct 9r prevents an efficient interaction of azoalkene 7a with the second, unconverted C=S group. In contrast to 7a, the N-benzoyl substituted azoalkene 7b reacted with both thioketones 1i and 1j, yielding complex mixtures of products, which undergo decomposition during attempted chromatographic separation. It is worth mentioning that the 1H-NMR spectra of the crude product mixtures revealed the presence of a product with a characteristic AB-system found at 3.77 and 4.14 ppm, respectively, with J = 12.4 Hz. These signals were accompanied by a less intense singlet located at 3.83 ppm. It is very likely that similar products were formed in both reactions, but all attempts aimed at their isolation, either by column chromatography or by treatment with organic solvents, were unsuccessful and in both experiments, only further decomposition was observed. Similarly, adamantanethione (1k) did not yield stable products neither with 7a nor 7b.
The experiments performed with thiochalcone 1l and both azoalkenes 7 deserve a brief comment. Whereas the N-tosyl-substituted 7a formed an unstable cycloadduct which could not be isolated in a pure form, the N-benzoyl analogue 7b delivered the 1,3,4-thiadiazine derivative 9s as a stable, crystalline product, isolated after chromatography in 38% yield (Scheme 7).
The (4 + 2)-cycloaddition also occurred regioselectively in this case leading to 9s as the sole product. It was isolated as a stable crystalline material with no tendency to decomposition at ambient conditions. Its structure was easily elucidated based on spectroscopic data. For example, diagnostic signals of the 6-CH2 fragment appeared in the 1H-NMR spectrum as the expected AB-system localized at 3.50 and 3.62 ppm (2JH,H = 17.0 Hz), and the styryl group appeared as two doublets found at 6.67 and 6.97 ppm (3JH,H = 16.0 Hz). Furthermore, the C=N and C=O absorptions appeared in the 13C-NMR spectrum at 140.6 ppm and 170.2 ppm, respectively. As discussed in our publication about the hetero-Diels-Alder reactions of thiochalcones with α-nitrosostyrene [3], the combination of these two 4π systems can theoretically afford eight isomers, but it provided selectively one isomer. Apparently, azoalkene 7b shows a similarly high regio- and periselectivity.
In summary of this experimental section, one can conclude that in most of the studied cases, the ‘superdienophilic’ thioketones 1 react with azoalkenes 7a,b yielding a series of scarcely known 6H-1,3,4-thiadiazine derivatives of type 9 and 11. In the latter case, the initially formed (4 + 2)-cycloadducts 9q,r with the N-tosyl group located at N-3 underwent elimination of sulfinic acid and converted into the isolated 2H-1,3,4-thiadiazine derivatives 11. Remarkably, the stability of the cycloadducts obtained with thiochalcone 1l depends on the substitution pattern, and only in the case of N-benzoyl azoethylene 7b could a stable product 9s be isolated. All successful (4 + 2)-cycloadditions proceeded with high regioselectivity and the conceivable formation of 1,2,3-thiadiazine derivatives 10 was experimentally not observed.

2.2. Mechanistic Investigations by DFT Calculations

The central question within this project concerns the experimentally observed explicit regioselectivity in favor of 1,3,4-thiadiazines derivatives 9 over the possible formation of 1,2,3-thiadiazines 10 (see Scheme 8) with respect to possible mechanisms involved, e.g., a concerted Diels-Alder reaction, a stepwise cycloaddition via zwitterions, or alternatives. In the following, the observed experimental outcome will be discussed on the basis of the Gibbs free energy surface as investigated by a comprehensive quantum chemical DFT study. Thus, mechanistic details, e.g., intermediates and reaction barriers via transition states, and the underlying thermodynamic and kinetic data will be in the center of investigation.
Starting with B3LYP/6-31G(d) [40,41] + GD3BJ [42,43] geometry optimizations for the gas phase, PBE1PBE/def2tzvp [44,45,46,47,48] + GD3BJ optimizations for the PCM solvent sphere of dichloromethane [49] were performed. In accordance with the reaction conditions (room temperature, no irradiation), only closed shell calculations were done, although diradical intermediates cannot principally be excluded. In the following, we discuss differences in Gibbs free energies (ΔG298) (kcal/mol) (see also Supplementary Materials for details); all relative energies refer to the sum of the starting materials. The structure and conformation of the observed 1,3,4-thiadiazines 9 for the calculations was derived from the X-ray structure of 9c (Figure 2, left side). Furthermore, minima for van der Waals complexes formed from the two starting compounds were localized on the Gibbs free energy surface (see Supplementary Materials). They indicate substantial noncovalent interactions (van der Waals forces) between both reaction partners due to π-interactions of the aromatic moieties and London dispersion interactions of the chemical groups involved (for an example, see Figure 3). Their relative energy (3.0–4.3 kcal/mol) is low and only slightly dependent on the substitution pattern (Table 1, entries 1–6).
Two principal classes of thioketones were in the center of interest for the reactions with azoalkene 7a, namely 1a, 1b, 1m, and 1n (Scheme 8, Table 1, entries 1–4) as examples of aryl-substituted thioketones and 2,2,4,4-tetramethyl-3-thioxocyclobutan-1-one (1i) (Scheme 9, Table 1, entries 5–6) as a prototype of a sterically hindered and strained aliphatic thioketone. Table 1 summarizes the computational results obtained. It is interesting to see that the calculated Gibbs free reaction energies of all 1,3,4-thiadiazines (9) and 1,2,3-thiadiazines (10) obtained are very similar; thus, thermodynamic control of these reactions can be excluded. From the thermodynamic point of view, the formation of both isomeric products is likely, with small differences due to the substitution pattern of the respective thioketone.
However, the Gibbs free activation energies differed significantly in favor of the 1,3,4-thiadiazines 9, indicating strong kinetic control with substantially lower activation barriers compared to the corresponding 1,2,3-thiadiazine regioisomers 10. The barriers of (4 + 2)-cycloadditions of thioketones 1m and 1n, respectively leading to 1,3,4-thiadiazine 9m and 9n, were only slightly higher than those of 1a and 1b, but still not excessively high (compare entries 3 and 4 with entries 1 and 2). Hence, these calculations cannot easily explain the failed cycloadditions of 7a with 1m and 1n.
Concerning the reaction mechanisms of the cyclizations to 9 or 10, substantial differences were found with regard to the atomic distances of the forming bonds in the respective transition states (Table 2). Thus, the transition states A for the formation of the 1,3,4-thiadiazines 9 were characterized by very unsymmetric structures with relatively small C..S-distances (2.27–2.58 Å) and quite large N..C-distances (3.02–3.57 Å) (differences from 0.62 to 1.30 Å). As an example, the transition state of the cycloaddition of azoalkene 7a with thiobenzophenone (1a) is depicted in Figure 3 (right side), which illustrates the fairly unsymmetric formation of the new bonds. Computationally, these transition states are not easy to trace, since they are localized on a very flat energy surface (see below for an extreme example). In contrast, the hypothetical formation of the 1,2,3-thiadiazine isomers 10 is well in accord with a classical concerted hetero-Diels-Alder reaction with atomic distances of 2.15–2.55 Å and a maximal difference of 0.36 Å in transition states B (Figure 4).
A special example of the unsymmetric formation of the 1,3,4-thiadiazine 9 is the reaction of 7a with the aliphatic, sterically hindered, ring-strained 2,2,4,4-tetramethyl-3-thioxocyclobutan-1-one (1i) involving the C=S bond, which was studied in more detail (see Scheme 9, Table 1 and Table 2, entry 5). The maximal difference of the distances between the reacting atoms is quite large (1.3 Å), indicating a two-step mechanism involving an open-chain zwitterionic intermediate instead of a concerted (4 + 2)-cycloaddition. Indeed, an intermediate Z and two transition states TS A1 and TS A2 could be localized on the energy surface (Figure 5). The energy of stabilized zwitterion Z lies only 9.5 kcal above that of the starting materials and its barrier for cyclization to 9q is low. Apparently, it is the steric hindrance exhibited by the four methyl groups of 1i that hampers the approach of the thiocarbonyl C-atom to the azoalkene N-atom and thus prevents the concerted (4 + 2)-cycloaddition in this case. These computational results imply further experimental studies in order to react zwitterionic intermediates such as Z with suitable trapping agents. The final elimination of the sulfinic acid 12 from 9q affording 11a was calculated to be slightly endergonic but supported by the experimental conditions (presence of base and of silica gel were not considered in the calculations).
As an alternative, the cycloadditions of 7a to the C=O bond of 1i were also considered. According to the computational results, such reaction modes are very unlikely due to quite high activation barriers and to low free reaction energies. Thus, not unexpectedly, the C=O bonds do not compete with the much more reactive C=S bonds for the formation of six-membered heterocyclic products (see Table 1 and Table 2, entries 6).
In summary, the computational investigations presented here interpret the experimental findings as the result of strong dominance of kinetic over thermodynamic control. The transition states leading to the observed 1,3,4-thiadiazine derivatives 9 obtained are quite asymmetric with relatively short C..S-distances but long N..C-distances. In extreme cases, when high steric hindrance comes into play as in 9q, a two-step mechanism involving a zwitterionic intermediate Z dominates over the concerted (4 + 2)-cycloaddition reaction. In the other cases, competition of both pathways seems possible. Surprisingly, although the (4 + 2)-cycloaddition fits from the bond distances well to more “conventional” reaction pathways, this route to the 1,2,3-thiadiazines 10 is kinetically strongly disfavored. The proposed intermediacy of the zwitterionic intermediates discussed here should experimentally be verified by employing suitable trapping reagents.

3. Materials and Methods

3.1. Materials

The starting thioketones 1 were prepared by thionation of the corresponding ketones with Lawesson’s reagent (in the cases of 1a, 1ch, and 1l) [50], a mixed stream of H2S/HCl (for 1b) [51], or by treatment with P2S5 in pyridine solution (for 1ik) [52]. The precursors of azoalkenes 7a and 7b, i.e., hydrazones 8a [53] and 8b [54], respectively, were synthesized from commercially available 2-chloroacetophenone (Chemat, Gdańsk, Poland) and the respective hydrazides (TolSO2NHNH2 or PhCONHNH2, respectively) in CH2Cl2 solution at rt (overnight stirring). After evaporation of the solvent, the residual oily materials were treated with a small portion of MeOH. The next day, the colorless crystals formed were filtered off and air-dried over a few hours.

3.2. Analytical Methods and Equipment

General information. All commercially available solvents and reagents were used as received. If not stated otherwise, reactions were performed in flame-dried flasks under an argon atmosphere and addition of the reactants by using a syringe; subsequent manipulations were conducted in the air. NMR spectra were taken with Bruker AVIII (1H-NMR (600 MHz); 13C-NMR (151 MHz); Bruker, Billerica, MA, USA). Chemical shifts are given relative to solvent residual peaks and integrals in accordance with assignments and coupling constants J in Hz. For detailed peak assignments, 2D spectra were measured (COSY, HMQC). The UV-Vis spectra were recorded with a JASCO V-630 spectrometer (JASCO, Easton, MD, USA) in CH2Cl2 solutions. The HRMS spectra were registered with a Varian 500-MS LC Ion Trap (Palo Alto, CA, USA) or with a Waters Synapt G2-Si mass spectrometer. The MS/ESI spectra were measured using a Varian 500 MS LS Ion Trap sp apparatus. IR measurements were performed with an Agilent Cary 630 FTIR spectrometer, in neat (Agilent, Santa Clara, CA, USA). Elemental analyses were obtained with a Vario EL III instrument. Melting points were determined in capillaries with an Aldrich Melt-Temp II apparatus and are uncorrected.

3.3. Quantum Chemical Calculations

Quantum chemical calculations (PBE1PBE/def2-TZVP+PCM (dichloromethane)+GD3BJ dispersion correction) [43,44,45,46,47,48,49,50] were performed on the basis of preceding B3LYP/6-31G(d) [41,42] +GD3BJ-geometry optimizations using the Gaussian 16, Revision B.01 [55], package of programs. To obtain the most reliable structural information, several conformers of each isomer were calculated, often after MM2-conformational analysis. The transition state localizations are based on reaction path calculations by stepwise, independent elongation of both relevant bonds beginning with the cycloadducts (“retro-Diels-Alder reaction”) and full optimization of all other parameters. Transition state searches or QST2 calculations on the basis of the calculated 3D-surfaces followed. In several cases, IRC-calculations were performed in order to characterize the transition states obtained. Minima for van der Waals complexes of the reacting starting material were also localized on the energy surface.

3.4. Synthesis

Reactions of Azoalkenes 7a,b with Thioketones 1ak and Thiochalcone 1l—General Procedures

The appropriate thioketone 1 (1 mmol) and 1.1 mmol of the corresponding precursor 8a or 8b were dissolved in 2 mL of freshly distilled dichloromethane and an excess of pre-dried potassium carbonate was added in small portions. The resulting suspension was stirred magnetically at room temperature under argon atmosphere until the color of the thioketone disappeared (overnight stirring is recommended). After completion of the reaction, the potassium carbonate was filtered off on a filter paper and the solvent was evaporated in a vacuum evaporator. The crude mixtures of products were purified by chromatography (SiO2) using a 1:4 mixture of petroleum ether and dichloromethane (PE/DCM) as an eluent. Analytically pure samples were obtained by crystallization of the isolated material from petroleum ether or hexane with a small admixture of dichloromethane.
2,2,5-Triphenyl-3-(p-tolylsulfonyl)-6H-1,3,4-thiadiazine (9a): Yield: 260 mg (54%). Colorless crystals (petroleum ether/CH2Cl2). M.p. 178–182 °C. 1H-NMR: δ 2.46 (s, CH3), 3.21 (s, CH2), 7.27, 7.73 (AB system, 2JH,H = 8.1 Hz, 4CHarom), 7.35–7.44 (m, 9CHarom), 7.57–7.60 (m, 6CHarom) ppm. 13C-NMR: δ 21.7 (CH3), 24.4 (CH2), 77.7 (Cq), 125.5, 128.0, 128.1, 128.5, 128.6, 129.0, 129.2, 129.4 (19CHarom), 137.3, 137.5, 139.4, 141.7, 143.7 (5Carom and C=N) ppm. IR: ν 1595 (m), 1489 (m), 1355 (s), 1172 (s), 1090 (m), 1008 (s), 924 (m), 902 (m), 760 (s), 660 (s), 577 (m), 540 (vs) cm−1. C28H24N2O2S2 (484.63): calcd. C 69.39, H 4.99, N 5.78, S 13.23; found C 69.37, H 4.98, N 5.80, S 13.17.
3-Benzoyl-2,2,5-triphenyl-6H-1,3,4-thiadiazine (9b): Yield: 255 mg (59%). Colorless crystals (petroleum ether/CH2Cl2). M.p. 177–179 °C. 1H-NMR: δ 3.34 (s, CH2), 7.26–7.61 (m, 18CHarom), 7.86–7.88 (m, 2CHarom) ppm. 13C-NMR: δ 24.9 (CH2); 73.6 (Cq), 125.2, 127.6, 128.2, 128.3, 128.4, 128.5, 129.0, 130.3, 131.1 (20CHarom), 135.6, 137.3, 139.2, 140.5 (4Carom and C=N), 170.8 (C=O) ppm. IR: ν 1699 (C=O, s), 1605 (m), 1481 (m), 1311 (m), 1289 (s), 1155 (m), 1077 (m), 957 (m), 812 (m), 711 (s), 693 (vs) cm−1. MS/ESI: m/e 435 (100%, [M + 1]+). C28H22N2OS (434.56): calcd. C 77.39, H 5.10, N 6.45, S 7.38; found C 77.31, H 5.04, N 6.61, S 7.18.
5-Phenyl-3-(p-tolylsulfonyl)spiro [6H-1,3,4-thiadiazine-2,9′-fluorene] (9c): Yield: 345 mg (72%). Pale yellow crystals (petroleum ether/CH2Cl2). M.p. 179–182 °C. 1H-NMR: δ 2.33 (s, CH3), 3.95 (s, CH2), 7.00, 7.82 (AB system, 2JH,H = 8.1 Hz, 4CHarom), 7.16–7.20 (m, 4CHarom), 7.28–7.30 (m, 2CHarom), 7.45–7.51 (m, 5CHarom), 7.88–7.89 (m, 2CHarom) ppm. 13C-NMR: δ 21.5 (CH3), 23.8 (CH2), 66.8 (Cq), 120.8, 123.3, 125.5, 127.8, 128.7, 128.8, 129.3, 129.7 (17CHarom), 134.7, 137.0, 138.2, 141.9, 143.7, 145.2 (7Carom and C=N) ppm. IR: ν 1587 (m), 1446 (m), 1353 (s), 1168 (s), 108 (m), 1002 (m), 775 (s), 743 (s), 667 (s), 577 (s), 548 (vs) cm−1. C28H22N2O2S2 (482.62): calcd. C 69.68, H 4.59, N 5.80, S 13.29; found C 69.42, H 4.56, N 5.93, S 13.23.
3-Benzoyl-5-phenyl-spiro[6H-1,3,4-thiadiazine-2,9′-fluorene] (9d): Yield: 245 mg (57%). Yellowish crystals (petroleum ether/CH2Cl2). M.p. 168 °C (dec.). 1H-NMR: δ 4.06 (s, CH2), 7.29–7.32 (m, 1CHarom), 7.37–7.52 (m, 11CHarom), 7.64–7.67 (m, 2CHarom), 7.76–7.79 (m, 2CHarom), 7.85–7.87 (m, 2CHarom) ppm. 13C-NMR: δ 23.3 (CH2); 65.6 (Cq), 121.0, 121.6, 125.4, 127.5, 127.9, 128.7, 128.9, 129.4, 130.4, 131.0 (18CHarom), 134.8, 137.2, 138.2, 139.8, 146.3 (6Carom and C=N), 168.8 (C=O) ppm. IR: ν 1714 (C=O, s), 1673 (m), 1599 (m), 1449 (m), 1285 (m), 1189 (m), 1095 (m), 917 (m), 807 (m), 730 (vs), 693 (s) cm−1. MS/ESI: m/e 433 (100%, [M + 1]+). C28H20N2OS (432.54): calcd. C 77.75, H 4.66, N 6.48, S 7.41; found C 77.48, H 4.68, N 6.74, S 7.16.
5-Phenyl-3-(p-tolylsulfonyl)spiro[6H-1,3,4-thiadiazine-2,9′-xanthene] (9e): Yield 340 mg (68%). Yellow–green crystals (petroleum ether/CH2Cl2). M.p. 145–147 °C. 1H-NMR: δ 2.42 (s, CH3), 3.66 (s, CH2), 7.03–7.05 (m, 2CHarom), 7.20–7.21 (m, 2CHarom), 7.24, 7.60 (AB system, 2JH,H = 8.1 Hz, 4CHarom), 7.28–7.29 (m, 2CHarom), 7.31–7.39 (m, 2CHarom), 7.46–7.49 (m, 3CHarom), 7.79–7.81 (m, 2CHarom) ppm. 13C-NMR: δ 21.6 (CH3), 23.7 (CH2), 60.3 (Cq), 117.0, 122.9, 125.6, 126.6, 128.7, 128.8, 129.3, 129.5, 129.8 (17CHarom), 134.9, 135.9, 136.9, 141.8, 144.2, 149.0 (7Carom and C=N) ppm. IR: ν 1599 (m), 1476 (m), 1444 (s), 1302 (s), 1168 (s), 1086 (s), 1010 (m), 749 (vs), 579 (s), 546 (vs) cm−1. C28H22N2O3S2 (498.62): calcd. C 67.45, H 4.45, N 5.62, S 12.86; found C 67.28, H 4.46, N 5.66, S 12.84.
3-Benzoyl-5-phenylspiro[6H-1,3,4-thiadiazine-2,9′-xanthene] (9f): Yield: 280 mg (62%). Yellowish crystals (petroleum ether/CH2Cl2). M.p. 181 °C (dec). 1H-NMR: δ 3.67 (s, CH2), 7.06–7.08 (m, 2CHarom), 7.25–7.27 (m, 4CHarom), 7.32–7.36 (m, 2CHarom), 7.38–7.40 (m, 3CHarom), 7.53–7.54 (m, 1CHarom), 7.61–7.63 (m, 2CHarom), 7.85–7.87 (m, 2CHarom) ppm. 13C-NMR: δ = 23.1 (CH2), 58.7 (Cq), 117.3, 123.3, 124.2, 125.5, 127.6, 128.6, 123.3, 129.5, 130.2, 130.9 (18CHarom), 123.9, 135.1, 137.1, 141.0, 149.7 (6Carom and C=N), 169.1 (C=O) ppm. IR: ν 1673 (s), 1599 (m), 1303 (m), 1474 (s), 1446 (s), 1311 (m), 1289 (s), 1159 (m), 1099 (m), 1017 (m), 887 (m), 797 (s), 749 (s), 712 (s), 689 (s) cm−1. MS/ESI: m/e 475 (100, [M]+). C28H20N2O2S∙1.5 H2O (475.56): calcd. C 70.72, H 4.87, N 5.89, S 6.74; found C 70.87, H 4.36, N 6.20, S 6.98.
5-Phenyl-3-(p-tolylsulfonyl)spiro[6H-1,3,4-thiadiazine-2,9′-thioxanthene] (9g): Yield: 375 mg (73%). Greenish crystals (petroleum ether/CH2Cl2). M.p. 161–163 °C. 1H-NMR: δ 2.46 (s, CH3), 3.36 (s, CH2), 7.17–7.19 (m, 2CHarom), 7.30–7.33 (m, 4CHarom), 7.40–7.42 (m, 4CHarom), 7.47–7.51 (m, 3CHarom), 7.80–7.82 (m, 4CHarom) ppm. 13C-NMR: δ 21.7 (CH3), 22.5 (CH2), 67.7 (Cq), 125.6, 125.7, 126.0, 127.5, 128.5, 128.8, 128.9, 129.5, 129.8 (17CHarom), 128.4, 133.6, 136.1, 137.0, 142.4, 144.4 (7Carom and C=N) ppm. IR: ν 1597 (m), 1444 (s), 1303 (m), 1168 (s), 1087 (s), 1008 (m), 771 (s), 742 (s), 652 (s), 557 (s), 546 (vs) cm−1. C28H22N2O2S3 (514.68): calcd. C 65.34, H 4.31, N 5.44, S 18.69; found C 65.25, H 4.36, N 5.57, S 18.69.
3-Benzoyl-5-phenylspiro[6H-1,3,4-thiadiazine-2,9′-thioxanthene] (9h): Yield: 320 mg (69%). Yellowish crystals (petroleum ether/CH2Cl2). M.p. 186 °C (dec). 1H-NMR: δ = 3.36 (s, CH2), 7.21–7.22 (m, 2CHarom), 7.27–7.32 (m, 5CHarom), 7.36–7.38 (m, 3CHarom), 7.46–7.48 (m, 2CHarom), 7.50–7.52 (m, 2CHarom), 7.57–7.59 (m, 2CHarom), 7.96–7.98 (m, 2CHarom) ppm. 13C-NMR: δ = 22.3 (CH2), 65.5 (Cq), 125.1, 125.6, 126.3, 126.8, 127.7, 127.9, 128.6, 129.5, 130.3, 131.0 (18CHarom), 129.4, 133.5, 135.1, 137.0, 142.7 (6Carom and C=N), 169.4 (C=O) ppm. IR: ν 1670 (C=O, s), 1602 (m), 1461 (m), 1386 (m), 1334 (m), 1285 (s), 1144 (m), 1073 (m), 913 (m), 793 (m), 745 (vs), 699 (s) cm−1. C28H20N2OS2∙1.5 H2O (491.60): calcd. C 68.41., H 4.72, N 5.70, S 13.04; found C 68.62, H 4.85, N 6.01, S 13.20.
2-Ferrocenyl-2-methyl-5-phenyl-3-(p-tolylsulfonyl)-6H-1,3,4-thiadiazine (9i): Yield: 295 mg (56%). Beige crystals (petroleum ether/CH2Cl2). M.p. 124 °C (dec). 1H-NMR: δ 2.41, 2.50 (2s, 2CH3), 3.34, 3.67 (AB system, 2JH,H = 16.6 Hz, CH2), 4.17–4.19 (m, 2CHFc), 4.24–4.25 (m, 1CHFc), 4.33 (s, 5CHFc), 4.49–4.50 (m, 1CHFc), 7.23, 7.70 (AB system, 3JH,H = 8.2 Hz, 4CHarom), 7.37–7.39 (m, 3CHarom), 7.59–7.61 (m, 2CHarom) ppm. 13C-NMR: δ 21.6, 28.5 (2CH3), 24.3 (CH2), 66.7, 67.3, 68.6, 69.7, 69.8 (9CHFc), 67.4 (Cq), 90.6 (CFc), 125.5, 128.4, 128.9, 129.2, 129.6 (9CHarom), 136.8, 137.2, 143.3, 144.2 (3Carom and C=N) ppm. IR: ν 1597 (m), 1444 (m), 1357 (m), 1174 (s), 1120 (m), 1094 (m), 986 (m), 809 (s), 747 (m), 669 (s), 562 (s), 542 (vs) cm−1. MS-ESI: m/e 530 (100, [M]+), 531 (85, [M + 1]+), 553 (46, [M + Na]+). C27H26FeN2O2S2 (530.48): calcd. C 61.13, H 4.94, N 5.28, S 12.09; found C 60.94, H 4.97, N 5.43, S 11.99.
3-Benzoyl-2-ferrocenyl-2-methyl-5-phenyl-6H-1,3,4-thiadiazine (9j): Yield: 185 mg (39%). Beige crystals (petroleum ether/CH2Cl2). M.p. 154 °C (dec). 1H-NMR: δ 2.45 (s, CH3), 3.45, 3.80 (AB system, 2JH,H = 17.3 Hz, CH2), 4.19–4.20 (m, 1CHFc), 4.23–4.25 (m, 1CHFc), 4.35 (s, 5CHFc), 4.37–4.39 (m, 2CHFc), 7.27–7.29 (m, 4CHarom), 7.39–7.41 (m, 3CHarom), 7.44–7.47 (m, 1CHarom), 7.63–7.64 (m, 2CHarom) ppm. 13C-NMR: δ 24.0 (CH2), 26.1 (CH3), 62.6 (Cq), 66.1 (1CHFc), 66.5 (1CHFc), 68.4 (1CHFc), 68.7 (1CHFc), 69.6 (5CHFc), 91.7 (CFc), 125.1, 127.4, 128.4, 128.8, 129.5, 130.2 (10CHarom), 136.9, 137.3, 139.1 (2Carom and C=N), 170.9 (C=O) ppm. IR: ν 1689 (C=O, s), 1610 (m), 1481 (m), 1341 (m), 1337 (m), 1289 (s), 1174 (m), 1103 (m), 1002 (m), 902 (m), 752 (m), 748 (vs), 711 (s) cm−1. C27H24FeN2OS (480.41): calcd. C 67.50, H 5.04, N 5.83, S 6.67; found C 67.32, H 5.17, N 5.85, S 6.51.
2-Ferrocenyl-2,5-diphenyl-3-(p-tolylsulfonyl)-6H-1,3,4-thiadiazine (9k): Yield: 230 mg (39%). Beige crystals (petroleum ether/CH2Cl2). M.p. 158–160 °C. 1H-NMR: δ 2.44 (s, CH3), 3.08, 3.48 (AB system, 2JHH = 16.9 Hz, CH2), 4.11 (s, 1CHFc), 4.30 (d, J = 5.6 Hz, 2CHFc), 4.39 (s, 6CHFc), 7.21–8.20 (m, 14CHarom) ppm. 13C-NMR: δ 21.6 (CH3), 25.2 (CH2), 67.4, 68.0, 69.0, 70.2, 72.7 (9CHFc), 74.5 (Cq), 89.4 (CFc), 125.5, 127.8, 128.1, 128.2, 128.4, 128.5, 128.8, 129.0 (14CHarom); 137.1, 137.6, 141.0, 141.5, 143.4 (4Carom and C=N) ppm. IR: ν 1597 (m), 1446 (m), 1351 (m), 1304 (m), 1169 (s), 1088 (s), 1008 (m), 810 (m), 751 (m), 689 (s), 568 (s), 539 (vs) cm−1. MS-ESI: m/z 593 (100, [M + 1]+), 615 (27, [M + Na]+. C32H28FeN2O2S2 (592.55): calcd. C 64.86, H 4.76, N 4.73, S 10.82; found C 64.99, H 4.59, N 4.95, S 10.91.
3-Benzoyl-2-ferrocenyl-2,5-diphenyl-6H-1,3,4-thiadiazine (9l): Yield: 175 mg (32%). Beige crystals (petroleum ether/CH2Cl2). M.p. 157 °C (dec). 1H-NMR: δ 3.17, 3.55 (AB system, 2JHH = 17.1 Hz, CH2), 3.99–4.01 (m, 1CHFc), 4.20–4.22 (m, 1CHFc), 4.29–4.30 (m, 1CHFc), 4.45 (s, 5CHFc), 4.59–4.61 (m, 1CHFc), 7.24–7.29 (m, 4CHarom), 7.35–7.44 (m, 7CHarom), 7.47–7.51 (m, 2CHarom), 7.70–7.74 (m, 2CHarom) ppm. 13C-NMR: δ 25.8 (CH2), 66.8 (Cq), 67.0 (1CHFc), 68.4 (1CHFc), 69.6 (5CHFc), 70.5 (1CHFc), 71.3 (1CHFc), 92.0 (CFc), 125.2, 127.5, 127.9, 128.0, 128.2, 128.4, 128.9, 130.0, 130.7 (15CHarom), 136.2, 137.3, 139.1, 139.6 (3Carom, C=N), 170.5 (C=O) ppm. IR: ν 1677 (C=O, s), 1611 (m), 1490 (m), 1446 (m), 1394 (m), 1341 (m), 1282 (s), 1155 (m), 989 (m), 820 (m), 741 (s), 693 (vs) cm−1. MS/ESI: m/e 542 (100%, [M + 1]+). C32H26FeN2OS (542.48): calcd. C 70.85, H 4.83, N 5.16, S 5.91; found C 70.79, H 4.95, N 5.09, S 6.17.
2-Ferrocenyl-5-phenyl-2-(2-thienyl)-3-(p-tolylsulfonyl)-6H-1,3,4-thiadiazine (9m): Yield: 290 mg (48%). Brown crystals (petroleum ether/CH2Cl2). M.p. 126–128 °C. 1H-NMR: δ 2.43 (s, CH3), 3.33, 3.58 (AB system, 2JH,H = 16.9 Hz, CH2), 4.31–4.33 (m, 4CHFc), 4.35–4.40 (m, 5CHFc), 6.94 (pseudo t, J = 4.3 Hz, 1CHarom), 7.23–7.25 (m, 2CHarom), 7.38–7.42 (m, 5CHarom), 7.65–7.67 (m, 4CHarom) ppm. 13C-NMR: δ 21.5 (CH3), 25.3 (CH2), 67.4, 68.0, 68.7, 70.4, 73.1 (9CHFc), 71.3 (Cq), 89.2 (CFc), 125.5, 125.7, 126.8, 128.4, 128.6, 128.8, 129.0, 129.1 (12CHarom), 137.0, 137.4, 141.5, 143.5, 145.3 (4Carom and C=N) ppm. IR: ν 1599 (m), 1439 (m), 1353 (s), 1169 (s), 1090 (m), 1003 (m), 939 (m), 805 (s), 751 (s), 665 (s), 565 (s), 538 (vs) cm−1. C30H26FeN2O2S3 (598.58): calcd. C 60.30, H 4.38, N 4.68, S 16.07; found C 60.36, H 4.30, N 4.71, S 16.02.
3-Benzoyl-2-ferrocenyl-5-phenyl-2-thienyl-6H-1,3,4-thiadiazine (9n): Yield: 205 mg (37%). Beige crystals (petroleum ether/CH2Cl2). M.p. 161 °C (dec). 1H-NMR: 3.48, 3.68 (AB system, 2JH,H = 17.2 Hz, CH2), 4.19–4.20 (m, 1CHFc), 4.19–4.20 (m, 1CHFc), 4.22–4.23 (m, 1CHFc), 4.29–4.30 (m, 1CHFc), 4.46 (s, 5CHFc), 4.57–4.60 (m, 1CHFc), 6.90–6.92 (m, 1CHarom), 7.02–7.44 (m, 1CHarom), 7.25–7.31 (m, 4CHarom), 7.39–7.41 (m, 5CHarom), 7.46–7.47 (m, 1CHarom), 7.67–7.69 (m, 2CHarom) ppm. 13C-NMR: δ 26.1 (CH2), 67.1 (Cq), 66.7 (1CHFc), 67.3 (1CHFc), 68.4 (CFc), 70.1 (5CHFc), 71.7 (1CHFc), 71.3 (1CHFc), 91.9 (CFc), 125.2, 125.9, 126.4, 127.5, 128.4, 128.8, 128.9, 130.3, 130.9 (15CHarom), 136.0, 137.2, 139.3, 144.0 (3Carom and C=N), 170.3 (C=O) ppm. IR: ν 1677 (C=O, s), 1609 (m), 1492 (m), 1444 (m), 1381 (m), 1282 (vs), 1244 (m), 1165 (m), 1108 (m), 988 (m), 820 (m), 726 (vs), 689 (vs) cm−1. C30H24FeN2OS2 (548.50): calcd. C 65.69, H 4.41, N 5.11, S 11.69; found C 65.70, H 4.51, N 5.21, S 11.42.
2-Cyclopropyl-2-ferrocenyl-3-(p-tolylsulfonyl)-6H-1,3,4-thiadiazine (9o): Yield: 290 mg (48%). Yellowish crystals (petroleum ether/CH2Cl2). M.p. 105‒106 °C (decomp.). 1H-NMR: δ 0.70–0.82 (m, 4CHCp), 0.83–0.91 (m, 1CHCp), 1.06–1.13 (m, 1CHCp), 2.33–2.41 (m, 1CHCP), 2.42 (s, CH3), 3.41, 3.90 (AB system, J = 16.7 Hz, CH2), 4.23–4.25 (m, 2CHFc), 4.28–4.29 (m, 1CHFc), 4.31 (s, 5CHFc), 4.77–4.78 (m, 1CHFc), 7.22, 7.68 (AB system, JH,H = 8.2 Hz, 4CHarom), 7.36–7.42 (m, 3CHarom), 7.59–7.61 (m, 2CHarom). 13C-NMR: δ 1.8, 4.3 (2CH2(Cp)), 21.0 (CHCp), 21.8 (CH3), 24.2 (CH2), 66.8 (Cq), 66.9, 68.1, 68.2, 69.8 (4CHFc), 70.7 (5CHFc), 90.3 (CFc), 125.5, 128.4, 128.5, 128.8, 129 (9CHarom), 137.2, 137.4, 139.1, 141.2 (3Carom and C=N). IR: ν 1608 (m), 1440 (m), 1346 (s), 1170 (s), 1092 (s), 1006 (m), 943 (m), 816 (m), 753 (m), 690 (s), 667 (vs) cm−1. C29H28FeN2O2S2 (556.52): calcd. C 62.59, H 5.07, N 5.03, S 11.52; found C 62.22, H 5.23, N 5.20, S 11.37.
3-Benzoyl-2-cyclopropyl-2-ferrocenyl-5-phenyl-6H-1,3,4-thiadiazine (9p): Yield: 125 mg (23%). Orange thick oil. 1H-NMR: δ 0.6–0.9 (m, 4CHCp), 2.32–2.34 (m, 1CHCp), 3.59, 4.09 (AB system, 2JH,H = 17.1 Hz, CH2), 4.24–4.26 (m, 2CHFc), 4.35 (s, 5CHFc), 4.40–4.41 (m, 1CHFc), 4.75–4.77 (m, 1CHFc), 7.27–7.30 (m, 3CHarom), 7.36–7.39 (m, 4CHarom), 7.42–7.43 (m, 1CHarom), 7.55–7.57 (m, 2CHarom). 13C-NMR: δ 1.7, 3.9 (2CH2(Cp)), 21.2 (CHCp), 24.6 (CH2), 66.6, 66.7, 68.4, 69.6, 69.7 (9CHFc), 67.3 (Cq), 91.7 (CFc), 125.1, 127.3, 128.4, 128.8, 129.4, 130.1 (10CHarom), 137.2, 137.4, 139.1 (2Carom and C=N), 170.9 (C=O). IR: ν 1677 (C=O, s), 1613 (m), 1446 (m), 1405 (m), 1341 (m), 1285 (vs), 1241 (m), 1166 (m), 1107 (m), 1073 (m), 1028 (m), 950 (m), 820 (m), 758 (s), 693 (vs) cm−1. C29H26FeN2OS (506.45): calcd. C 68.78, H 5.17, N 5.53, S 6.33; found C 68.58, H 5.13, N 5.80, S 6.34.
1,1,3,3-Tetramethyl-7-phenyl-5-thia-8,9-diazaspiro[3.5]nona-6,8-dien-2-one (11a): Yield: 160 mg (56%). Orange crystals (petroleum ether/CH2Cl2). M.p. 93–95 °C. 1H-NMR: δ 1.34, 1.39 (2s, 4CH3), 6.54 (s, CH-S), 7.41–7.42 (m, 1CHarom), 7.46–7.49 (m, 2CHarom), 7.77–7.79 (m, 2CHarom) ppm. 13C-NMR: δ 21.5, 21.6 (4CH3), 64.7, 70.0 (3Cq), 107.4 (CH-S), 125.0, 128.8, 129.1 (5CHarom), 134.7, 151.8 (Carom and C=N), 217.6 (C=O) ppm. IR: ν 2965 (m), 1785 (vs), 1446 (m), 1422 (m), 1362 (m), 1203 (m), 1041 (m), 1030 (m), 1002 (m), 747 (vs), 691 (vs), 656 (s) cm−1. UV-Vis: λ1: 239 (ε1 = 1.3 × 104); λ2: 283 (ε2 = 3.7 × 103), λ3: 358 (ε3 = 7.4 × 102)); λ4: 430 (ε4 = 1.6 × 102) nm. MS-ESI: m/z 287 (100, [M + 1]+), 309 (18, [M + Na]+). C16H18N2OS (286.39): calcd. C 67.10, H 6.34, N 9.78, S 11.19; found C 67.06, H 6.33, N 9.75, S 11.15.
1,1,3,3-Tetramethyl-7-phenyl-5-thia-8,9-diazaspiro[3.5]nona-6,8-diene-2-thione (11b): Yield: 210 mg (69%). Orange crystals (petroleum ether/CH2Cl2). M.p. 71–74 C. 1H-NMR: δ 1.40, 1.45 (2s, 4CH3), 6.54 (s, CH-S), 7.40–7.42 (m, 1CHarom), 7.47–7.50 (m, 2CHarom), 7.8–7.81 (m, 2CHarom) ppm. 13C-NMR: δ 25.6, 25.7 (4CH3), 67.3, 73.3 (3Cq), 107.3 (CH-S), 124.9, 128.7, 129.0 (5CHarom), 134.7, 151.7 (Carom and C=N), 277.4 (C=O) ppm. IR: ν 2961 (m), 2931 (m), 1459 (m), 1429 (m), 1422 (m), 1358 (m), 1300 (s), 1157 (m), 1025 (m), 749 (vs), 691 (vs) cm−1. UV-Vis: λ1 = 268 (ε1 = 1.36 × 104); λ2 = 364 (ε2 = 1.91 × 103); λ3 = 426 (ε3 = 5.8 × 102) nm. C16H18N2S2 (302.45): calcd. C 63.54, H 6.00, N 9.26, S 21.20; found C 63.53, H 5.92, N 9.20, S 21.17.
3-Benzoyl-2,5-diphenyl-2-styryl-6H-1,3,4-thiadiazine (9s): Yield: 175 mg (38%). Thick yellowish oil. 1H-NMR: δ 3.50, 3.62 (AB system, 2JH,H = 17.0 Hz, CH2), 6.67, 6.97 (2d, 3JH,H = 16.0 Hz, 2CH=), 7.28–7.43 (m, 16CHarom), 7.69–7.70 (m, 2CHarom), 7.89–7.91 (m, 2CHarom) ppm. 13C-NMR: δ 24.0 (CH2), 69.1 (Cq), 125.3, 126.4, 127.9, 127.5, 127.6, 128.2, 128.3, 128.5, 128.6, 128.7, 129.2, 130.2, 131.0, 131.7 (20CHarom and 2CH=), 135.5, 135.9, 137.3, 140.6 (3Carom and C=N), 170.2 (C=O) ppm. IR: ν 1669 (C=O, s), 1590 (m), 1492 (m), 1285 (s), 1226 (m), 1148 (m), 1073 (m), 961 (m), 726 (s), 689 (vs) cm−1. ESI-MS for C30H25N2OS [M + 1]+: cald. mass: 461.1688, found mass: 461.1687. ESI-MS for C30H24N2OSNa [M + Na]+: cald. mass: 483.1507, found mass: 483.1509.

4. Conclusions

A large number of new reviews [26,27,56,57] and original publications [31,58,59,60,61,62], which appeared in the last four years, demonstrates the growing interest in the chemistry of azoalkenes and their applications in hetero-Diels-Alder reactions. The presented study showed important mechanistic aspects and demonstrated that scarcely known 3,6-dihydro-2H-1,3,4-thiadiazines can efficiently be prepared by regioselective hetero-Diels-Alder reactions of thioketones with in situ-generated azoalkenes, bearing an electron-withdrawing substituent at the terminal N-atom. In comparison to aromatic and ferrocenyl-substituted thioketones, sterically hindered cycloaliphatic representatives differ in their reactivity and no stable cycloadducts were obtained with adamantanethione (1k). Interestingly, the primary products derived from cyclobutanethiones 1i,j suffered an elimination of toluene sulfinic acid to provide the rare 2H-1,3,4-thiadiazines. Differing stability of the target (4 + 2)-cycloadducts was also observed in experiments with thiochalcone 1l and azoalkenes 7a and 7b, respectively.
A comprehensive computational study demonstrated that the preferred formation of 1,3,4-thiadiazines 9 from azoalkene 7a and thioketones 1 is best explained by strong kinetic control of the respective cycloaddition over the formation of 1,2,3-thiadiazines 10. The transition states involved for 1,3,4-thiadizine formation are quite unsymmetric with short C..S and long N..C-distances. In the case of high sterical hindrance at the thioketone carbon, even a two-step pathway via a zwitterionic intermediate could be traced. In contrast, the formation of the unobserved 1,2,3-thiadiazines 10 would involve a concerted hetero-Diels-Alder reaction. However, although thermodynamics are not unfavorable for these regioisomers, substantial kinetic barriers prevented formation of 10 under the experimental conditions. The proposed intermediacy of zwitterions suggests further experiments in order to trap these transient species by suitable reagents.
It has to be underlined that 1,3,4-thiadiazines and fused systems containing this motif attract attention as biologically active compounds [63]. However, in spite of the growing interest in the application of thia-Diels-Alder reactions for the synthesis of six-membered S-heterocycles [3,16,21,22,64,65], efficient (4 + 2)-cycloadditions with easily accessible thioketones as well as thiochalcones with in situ-generated, reactive azoalkenes are to date practically unexplored for the preparation of the 1,3,4-thiadiazine skeleton. Therefore, the described experimental and theoretical results open new horizons for the development of syntheses of this type of N,S-heterocycles, including an asymmetric approach in the case of prochiral ferrocenyl/aryl and ferrocenyl/alkyl thioketones.
In general, one can conclude that—in spite of the fact that the first study by Diels and Alder was published almost 100 years ago [66,67]—the exploration of (4 + 2)-cycloaddition reactions [68] as a universal tool for construction of six-membered carbo- and heterocyclic systems still inspires new generations of chemists for the development of new methods in organic synthesis and offers a plethora of theoretical problems, which are of crucial importance for better understanding the mechanisms of organic reactions.

Supplementary Materials

The Supplementary Materials are available online and they contain the scanned 1H and 13C NMR spectra for the described and isolated compounds. The X-ray crystallography data for compounds 9c and 9g are deposited as CSD Communication under deposition number 2072033 and 2072034.

Author Contributions

G.M.—project foundation, laboratory work coordination and consultancy, founds availability, reference collection, and checking manuscript preparation; K.U.—laboratory work, collection of spectroscopic data, the manuscript, and SM part preparation; M.S.—part of the laboratory work and collection of spectroscopic data; H.H.—consultancy, manuscript preparation; E.-U.W.—computational work, main manuscript, and SM part preparation; R.Z.—substrate preparation; D.L.—X-ray analysis; H.-U.R.—project foundation, reference collection and checking their contents, and manuscript preparation. All authors have read and agreed to the published version of the manuscript.

Funding

The work was partially supported by the National Science Center (Cracow) within the project Beethoven-2 (grant #2016/23/G/ST5/04115/l) (for G.M.) and Alexander von Humboldt Foundation (Bonn) within the Institutspartnerschaft project.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Reported data are available at Authors via e-mail contact.

Acknowledgments

Authors thank Małgorzta Celeda (Łódź) for skillful help in the preparation of starting thioketones. G.M. thanks Wolfgang Weigand (Jena) for fruitful discussions on the chemistry of ferrocenyl thioketones (the AvH Project ‘Institutspartnerschaft’).

Conflicts of Interest

Authors do not report any conflict of interest.

References

  1. Mlostoń, G.; Grzelak, P.; Hamera-Faldyga, R.; Jasiński, M.; Pipiak, P.; Urbaniak, K.; Albrecht, Ł.; Hejmanowska, J.; Heimgartner, H. Aryl, hetaryl, and ferrocenyl thioketones as versatile building blocks for exploration in the organic chemistry of sulfur. Phosphorus Sulfur Silicon Relat. Elem. 2017, 192, 204–211. [Google Scholar] [CrossRef]
  2. Huisgen, R.; Fisera, L.; Giera, H.; Sustmann, R. Thiones as superdipolarophiles. Rates and equilibria of nitrone cycloadditions to thioketones. J. Am. Chem. Soc. 1995, 117, 9671–9685. [Google Scholar] [CrossRef]
  3. Mlostoń, G.; Urbaniak, K.; Jasiński, M.; Würthwein, E.-U.; Heimgartner, H.; Zimmer, R.; Reissig, H.-U. The [4+2]-cycloaddition of α-nitrosoalkenes with thiochalcones as a prototype of periselective hetero-Diels-Alder reactions—Experimental and computational studies. Chem. Eur. J. 2020, 26, 237–248. [Google Scholar] [CrossRef] [PubMed]
  4. Grzelak, P.; Utecht, G.; Jasiński, M.; Mlostoń, G. First (3+2)-cycloadditions of thiochalcones as C=S dipolarophiles: Efficient synthesis of 1,3,4-thiadiazoles via reactions with fluorinated nitrile imines. Synthesis 2017, 49, 2129–2137. [Google Scholar]
  5. Huisgen, R.; Li, X.; Giera, H.; Langhals, E. “Thiobenzophenone S-methylide” (=(diphenylmethylidenesulfonio)methanide), and C,C multiple bonds: Cycloadditions and dipolarophilic reactivities. Helv. Chim. Acta 2001, 84, 981–999. [Google Scholar] [CrossRef]
  6. Rohr, U.; Schatz, J.; Sauer, J. Thio- and selenocarbonyl compounds as “superdienophiles” in [4+2] cycloadditions. Eur. J. Org. Chem. 1998, 1998, 2875–2883. [Google Scholar] [CrossRef]
  7. Mlostoń, G.; Urbaniak, K.; Linden, A.; Heimgartner, H. Selenophen-2-yl substituted thiocarbonyl ylides—At the borderline of dipolar and diradical reactivity. Helv. Chim. Acta 2015, 98, 453–461. [Google Scholar] [CrossRef] [Green Version]
  8. Mlostoń, G.; Pipiak, P.; Linden, A.; Heimgartner, H. Studies on the reactions of thiocarbonyl S-methanides with hetaryl thioketones. Helv. Chim. Acta 2015, 98, 462–473. [Google Scholar] [CrossRef] [Green Version]
  9. Mlostoń, G.; Pipiak, P.; Heimgartner, H. Diradical reaction mechanisms in [3+2]-cycloadditions of hetaryl thioketones with alkyl- or trimethylsilyl-substituted diazomethanes. Beilstein J. Org. Chem. 2016, 12, 716–724. [Google Scholar] [CrossRef] [Green Version]
  10. Mlostoń, G.; Hamera-Fałdyga, R.; Celeda, M.; Heimgartner, H. Efficient synthesis of ferrocifenes and other ferrocenyl substituted ethylenes via a ‘sulfur approach’. Org. Biomol. Chem. 2018, 16, 4350–4356. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  11. Mlostoń, G.; Jasiński, R.; Kula, K.; Heimgartner, H. A DFT study on the Barton-Kellogg reaction—The molecular mechanism of the formation of thiiranes in the reaction between diphenyldiazomethane and diaryl thioketones. Eur. J. Org. Chem. 2020, 2020, 176–182. [Google Scholar] [CrossRef]
  12. Mlostoń, G.; Urbaniak, K.; Utecht, G.; Lentz, D.; Jasiński, M. Trifluoromethylated 2,3-dihydro-1,3,4-thiadiazoles via the regioselective [3+2]-cycloadditions of fluorinated nitrile imines with aryl, hetaryl, and ferrocenyl thioketones. J. Fluor. Chem. 2016, 192, 147–154. [Google Scholar] [CrossRef]
  13. Ali, K.A.; Mlostoń, G.; Urbaniak, K.; Linden, A.; Heimgartner, H. [3+2]-Cycloadditions of nitrile imines with hetaryl thioketones. J. Sulfur Chem. 2017, 38, 604–613. [Google Scholar] [CrossRef] [Green Version]
  14. Mlostoń, G.; Kowalski, M.K.; Obijalska, E.; Heimgartner, H. Efficient synthesis of fluoroalkylated 1,4,2-oxathiazoles via regioselective [3+2]-cycloaddition of fluorinated nitrile oxides with thioketones. J. Fluor. Chem. 2017, 199, 92–96. [Google Scholar] [CrossRef] [Green Version]
  15. Mlostoń, G.; Kowalczyk, M.; Augustin, A.U.; Jones, P.G.; Werz, D.B. Ferrocenyl substituted tetrahydrothiophenes via formal [3+2]-cycloaddition reactions of ferrocenyl thioketones with donor-acceptor cyclopropanes. Beilstein J. Org. Chem. 2020, 16, 1288–1295. [Google Scholar] [CrossRef]
  16. Hejmanowska, J.; Jasiński, M.; Mlostoń, G.; Albrecht, Ł. Taming of thioketones—The first asymmetric thio-Diels-Alder reaction with thioketones as heterodienophiles. Eur. J. Org. Chem. 2017, 2017, 950–954. [Google Scholar] [CrossRef]
  17. Huisgen, R. 1,3-Dipolar cycloadditions. Past and future. Angew. Chem. Int. Ed. Engl. 1963, 2, 565–598. [Google Scholar] [CrossRef]
  18. Woodward, R.B.; Hoffmann, R. The conservation of orbital symmetry. Angew. Chem. Int. Ed. Engl. 1969, 8, 781–853. [Google Scholar] [CrossRef]
  19. McKee, M.L.; Mlostoń, G.; Urbaniak, K.; Heimgartner, H. Dimerization reactions of aryl selenophen-2-yl-substituted thiocarbonyl S-methanides as diradical processes: A computational study. Beilstein J. Org. Chem. 2017, 13, 410–416. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  20. Jasiński, R.; Dresler, E. On the question of zwitterionic intermediates in the [3+2] cycloaddition reactions: A critical review. Organics 2020, 1, 49–69. [Google Scholar] [CrossRef]
  21. Mlostoń, G.; Grzelak, P.; Linden, A.; Heimgartner, H. Thia-Diels–Alder reactions of hetaryl thioketones with nonactivated 1,3-dienes leading to 3,6-dihydro-2H-pyrans: Evidence for a diradical mechanism. Chem. Heterocycl. Compd. 2017, 53, 518–525. [Google Scholar] [CrossRef] [Green Version]
  22. Mlostoń, G.; Urbaniak, K.; Zimmer, R.; Reissig, H.-U.; Heimgartner, H. Hetero-Diels-Alder reactions of conjugated nitrosoalkenes with ferrocenyl, hetaryl and cycloaliphatic thioketones. Chem. Sel. 2018, 3, 11724–11728. [Google Scholar]
  23. Attanasi, O.A.; De Crescentini, L.; Filippone, P.; Mantellini, F.; Santeusanio, S. 1,2-Diaza-1,3-butadienes; just a nice class of compounds, or powerful tools in organic chemistry? Reviewing an experience. ARKIVOC 2002, xi, 274–292. [Google Scholar] [CrossRef] [Green Version]
  24. Lemos, A. Addition and cycloaddition reactions of phosphinyl- and phosphonyl-2H-azirines, nitrosoalkenes and azoalkenes. Molecules 2009, 14, 4098–4119. [Google Scholar] [CrossRef] [Green Version]
  25. Attanasi, O.A.; De Crescentini, L.; Favi, G.; Filippone, P.; Mantellini, F.; Perrulli, F.R.; Santeusanio, S. Cultivating the passion to build heterocycles from 1,2-diaza-1,3-dienes; the force of imagination. Eur. J. Org. Chem. 2009, 2009, 3109–3127. [Google Scholar] [CrossRef]
  26. Lopes, S.M.M.; Cardoso, A.L.; Lemos, A.; Pinho e Melo, T.M.V.D. Recent advances in the chemistry of conjugated nitrosoalkenes and azoalkenes. Chem. Rev. 2018, 118, 11324–11352. [Google Scholar] [CrossRef]
  27. Wei, L.; Shen, C.; Hu, Y.-Z.; Tao, H.-Y.; Wang, C.-J. Enantioselective synthesis of multi-nitrogen containing heterocycles using azoalkenes as key intermediates. Chem. Commun. 2019, 55, 6672–6684. [Google Scholar] [CrossRef]
  28. Gao, S.; Chen, J.-R.; Hu, X.-Q.; Cheng, H.-G.; Lu, L.-Q.; Xiao, W.-J. Copper-catalyzed enantioselective inverse electron-demand hetero-Diels–Alder reactions of diazadienes with enol ethers: Efficient synthesis of chiral pyridazines. Adv. Synth. Catal. 2013, 355, 3539–3544. [Google Scholar] [CrossRef]
  29. Zhong, X.; Lv, J.; Luo, S. [4+2] Cycloaddition of in situ generated 1,2-diaza-1,3-dienes with simple olefins: Facile approaches to tetrahydropyridazines. Org. Lett. 2015, 17, 1561–1564. [Google Scholar] [CrossRef] [PubMed]
  30. Huang, R.; Tao, H.-Y.; Wang, C.-J. PPh3-Mediated [4+2]- and [4+1]-annulations of maleimides with azoalkenes: Access to fused tetrahydropyridazine/pyrrolidinedione and spiro-dihydropyrazole/pyrrolidinedione derivatives. Org. Lett. 2017, 19, 1176–1179. [Google Scholar] [CrossRef] [PubMed]
  31. Mei, G.-J.; Zheng, W.; Gonçalves, T.P.; Tang, X.; Huang, K.-W.; Lu, Y. Catalytic Asymmetric formal [3+2] cycloaddition of azoalkenes with 3-vinylindoles: Synthesis of 2,3-dihydropyrroles. iScience 2020, 23, e100873. [Google Scholar] [CrossRef] [Green Version]
  32. Bonini, B.F.; Maccagnani, G.; Mazzanti, G.; Rosini, G.; Foresti, E. Cycloaddition reactions of sulphines and thiones with azoalkenes. J. Chem. Soc. Perkin Transl. I 1981, 1, 2322–2327. [Google Scholar] [CrossRef]
  33. Quan, B.-X.; Zhuo, J.-R.; Zhao, J.-Q.; Zhang, M.-L.; Zhou, M.-Q.; Zhang, X.-M.; Yuan, W.-C. [4+1] annulation reaction of cyclic pyridinium ylides with in situ generated azoalkenes for the construction of spirocyclic skeletons. Org. Biomol. Chem. 2020, 18, 1886–1891. [Google Scholar] [CrossRef]
  34. Johnson, C.K. ORTEP II, Report ORNL-5138; Oak Ridge National Laboratory: Oak Ridge, TN, USA, 1976. [Google Scholar]
  35. Mlostoń, G.; Grzelak, P.; Maciej, M.; Linden, A.; Heimgartner, H. Hetero-Diels-Alder reactions of hetaryl and aryl thioketones with acetylenic dienophiles. Beilstein J. Org. Chem. 2015, 11, 576–582. [Google Scholar] [CrossRef] [Green Version]
  36. Nonhebel, D.C. The chemistry of cyclopropylmethyl and related radicals. Chem. Soc. Rev. 1993, 22, 347–359. [Google Scholar] [CrossRef]
  37. Firestone, R.A. The low energy of concert in many symmetry-allowed cycloadditions supports a stepwise-diradical mechanism. Int. J. Chem. Kinet. 2013, 45, 415–428. [Google Scholar] [CrossRef]
  38. Firestone, R.A. The diradical mechanism for 1,3-dipolar cycloadditions and related thermal pericyclic reactions. Tetrahedron 1977, 33, 3009–3039. [Google Scholar] [CrossRef]
  39. Matczak, P.; Mlostoń, G.; Hamera-Fałdyga, R.; Görls, H.; Weigand, W. Structure of diferrocenyl thioketone: From molecule to crystal. Molecules 2019, 24, e3950. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  40. Becke, A.D. Density-functional thermochemistry. III. The role of exact exchange. J. Chem. Phys. 1993, 98, 5648–5652. [Google Scholar] [CrossRef] [Green Version]
  41. Lee, C.; Yang, W.; Parr, R.G. Development of the Colle-Salvetti correlation-energy formula into a functional of the electron density. Phys. Rev. B 1988, 37, 785–789. [Google Scholar] [CrossRef] [Green Version]
  42. Grimme, S.; Ehrlich, S.; Goerigk, L. Effect of the damping function in dispersion corrected density functional theory. J. Comput. Chem. 2011, 32, 1456–1465. [Google Scholar] [CrossRef]
  43. Grimme, S.; Hansen, A.; Brandenburg, J.G.; Bannwarth, C. Dispersion-corrected mean-field electronic structure methods. Chem. Rev. 2016, 116, 5105–5154. [Google Scholar] [CrossRef] [Green Version]
  44. Perdew, J.P.; Burke, K.; Ernzerhof, M. Generalized gradient approximation made simple. Phys. Rev. Lett. 1996, 77, 3865–3868. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  45. Perdew, J.P.; Burke, K.; Ernzerhof, M. Errata: Generalized gradient approximation made simple. Phys. Rev. Lett. 1997, 78, 1396. [Google Scholar] [CrossRef] [Green Version]
  46. Adamo, C.; Barone, V. Toward reliable density functional methods without adjustable parameters: The PBE0 model. J. Chem. Phys. 1999, 110, 6158–6169. [Google Scholar] [CrossRef]
  47. Ernzerhof, M.; Scuseria, G.E. Assessment of the Perdew-Burke-Ernzerhof exchange-correlation functional. J. Chem. Phys. 1999, 110, 5029–5036. [Google Scholar] [CrossRef] [Green Version]
  48. Weigend, R.; Ahlrichs, R. Balanced basis sets of split valence, triple zeta valence and quadruple zeta valence quality for H to Rn: Design and assessment of accuracy. Phys. Chem. Chem. Phys. 2005, 7, 3297–3305. [Google Scholar] [CrossRef]
  49. Tomasi, J.; Mennucci, B.; Cammi, R. Quantum mechanical continuum solvation models. Chem. Rev. 2005, 105, 2999–3093. [Google Scholar] [CrossRef] [PubMed]
  50. Pedersen, B.S.; Scheibye, S.; Nilsson, N.H.; Lawesson, S.O. Studies on organophosphorus compounds. 10. Syntheses of thioketones. Bull. Soc. Chim. Belg. 1978, 87, 223–228. [Google Scholar] [CrossRef]
  51. Campaigne, E.; Reid, W.B. Thiocarbonyls. 2. Thiofluorenone. J. Amer Chem. Soc. 1946, 68, 769–770. [Google Scholar] [CrossRef]
  52. Elam, E.U.; Davis, H.E. Chemistry of dimethylketene dimer. 7. Dimers of dimethylthioketene. J. Org. Chem. 1967, 32, 1562–1565. [Google Scholar] [CrossRef]
  53. Hatcher, J.M.; Coltart, D.M. Copper(I)-catalyzed addition of Grignard reagents to in situ derived N-sulfonyl azoalkenes: An umpolung alkylation procedure applicable to the formation of up to three contiguous quaternary centers. J. Am. Chem. Soc. 2010, 132, 4546–4547. [Google Scholar] [CrossRef]
  54. Lai, E.C.K.; Mackay, D.; Taylor, N.J.; Kenneth, N.; Watson, K.N. Competitive [4+2] and [3+2] cycloadditions of nitrosoalkenes to the lmino bond of bicyclic 1,2-oxazines. J. Chem. Soc. Perkin Trans. 1 1990, 1990, 1497–1506. [Google Scholar] [CrossRef]
  55. 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 16; Revision, B.01; Gaussian, Inc.: Wallingford, CT, USA, 2016. [Google Scholar]
  56. Chen, Z.; Meng, L.; Ding, Z.; Hu, J. Construction of Versatile N-heterocycles from in situ generated 1,2-diaza-1,3-dienes. Curr. Org. Chem. 2019, 23, 164–187. [Google Scholar] [CrossRef]
  57. Grosso, C.; Lieber, M.; Brigas, A.F.; Pinho e Melo, T.M.V.D.; Lemos, A. Regioselectivity in hetero Diels−Alder reactions. J. Chem. Educ. 2019, 96, 148–152. [Google Scholar] [CrossRef]
  58. Zhao, H.-W.; Pang, H.-L.; Zhao, Y.-D.; Liu, Y.-Y.; Zhao, L.-J.; Chen, X.-Q.; Song, X.-Q.; Feng, N.-N.; Du, J. Construction of 2,3,4,5-tetrahydro-1,2,4-triazines via [4+2] cycloaddition of α-halogeno hydrazones to imines. RSC Adv. 2017, 7, 9264–92760. [Google Scholar] [CrossRef] [Green Version]
  59. Wie, L.; Zhu, Q.; Song, Z.-M.; Liu, K.; Wang, C.-J. Catalytic asymmetric inverse electron demand Diels–Alder reaction of fulvenes with azoalkenes. Chem. Commun. 2018, 54, 2506–2509. [Google Scholar]
  60. Emamian, S.; Soleymani, M.; Moosavi, S.S. Copper(I)-catalyzed asymmetric aza Diels–Alder reactions of azoalkenes toward fulvenes: A molecular electron density theory study. New J. Chem. 2019, 43, 4765–4776. [Google Scholar] [CrossRef]
  61. Ciccolini, C.; Giacomo Mari, G.; Francesco, G.; Gatti, F.G.; Giuseppe Gatti, G.; Gianluca Giorg, G.; Fabio Mantellini, F.; Gianfranco Favi, G. Synthesis of polycyclic fused indoline scaffolds through a substrate-guided reactivity switch. J. Org. Chem. 2020, 85, 11409–11425. [Google Scholar] [CrossRef] [PubMed]
  62. Mei, G.-J.; Tang, X.; Tasdan, Y.; Lu, Y. Enantioselective dearomatization of indoles by an azoalkene-enabled (3+2) reaction: Access to pyrroloindolines. Angew. Chem. Int. Ed. 2020, 59, 648–652. [Google Scholar] [CrossRef]
  63. Novikova, A.P.; Perova, N.M.; Chupakhin, O.N. Synthesis and properties of functional derivatives of 1,3,4-thiadiazines and condensed systems based on these systems (Review). Chem. Heterocycl. Comp. 1991, 27, 1159–1172. [Google Scholar] [CrossRef]
  64. Blond, G.; Gulea, M.; Mamane, V. Recent contributions to hetero Diels-Alder reactions. Curr. Org. Chem. 2016, 20, 2161–2210. [Google Scholar] [CrossRef]
  65. Sachse, F.; Gebauer, K.; Schneider, C. Continuous flow synthesis of 2H-thiopyrans via thia-Diels–Alder reactions of photochemically generated thioaldehydes. Eur. J. Org. Chem. 2021, 2021, 64–71. [Google Scholar] [CrossRef]
  66. Diels, O.; Alder, K. Synthesen in der hydroaromatische Reihe. I. Mitteilung: Anlagerungen von ‘Di-en‘-kohlenwasserstoffen. Justus Liebigs Ann. Chem. 1928, 460, 98–122. [Google Scholar] [CrossRef]
  67. Norton, J.A. The Diels-Alder chemistry was reviewed in 1942 for the first time, see in the Diels-Alder diene synthesis. Chem. Rev. 1942, 31, 319–523. [Google Scholar] [CrossRef]
  68. Von Euler, H.; Josephson, K.O. Historically, first report on a (4+2)-cycloaddition reaction originates from a Swedish laboratory: Über Kondensationen an Doppelbindungen. I.: Über die Kondensation von Isopren mit Benzochinon. Ber. Deutsch. Chem. Gesell. 1920, 53, 822–826. [Google Scholar] [CrossRef] [Green Version]
Figure 1. Synthetically useful representatives of aromatic 1ad; ferrocenyl 1eh; cycloaliphatic 1ik; and α,β-unsaturated 1l thioketones.
Figure 1. Synthetically useful representatives of aromatic 1ad; ferrocenyl 1eh; cycloaliphatic 1ik; and α,β-unsaturated 1l thioketones.
Molecules 26 02544 g001
Scheme 1. Regioselective (4 + 2)-cycloadditions of thioketones (1a, 1e, 1k) and thiochalcone (1l) with in situ-generated α-nitrosoethylene (2) leading to 1,5,2-oxathiazines (3ad) [3,22].
Scheme 1. Regioselective (4 + 2)-cycloadditions of thioketones (1a, 1e, 1k) and thiochalcone (1l) with in situ-generated α-nitrosoethylene (2) leading to 1,5,2-oxathiazines (3ad) [3,22].
Molecules 26 02544 sch001
Scheme 2. Comparison of the course of hetero-Diels-Alder reactions of thiofluorenone (1b) and the corresponding S-oxide (sulfine) with stable, isolable azoalkenes (4a,b) [32].
Scheme 2. Comparison of the course of hetero-Diels-Alder reactions of thiofluorenone (1b) and the corresponding S-oxide (sulfine) with stable, isolable azoalkenes (4a,b) [32].
Molecules 26 02544 sch002
Scheme 3. Generation of the reactive azoalkenes (7a,7b) from hydrazones (8a,8b), respectively, and their (4 + 2)-cycloadditions with thioketones 1 leading competitively either to 1,3,4-thiadiazines (9) or the regioisomeric 1,2,3-thiadiazines (10).
Scheme 3. Generation of the reactive azoalkenes (7a,7b) from hydrazones (8a,8b), respectively, and their (4 + 2)-cycloadditions with thioketones 1 leading competitively either to 1,3,4-thiadiazines (9) or the regioisomeric 1,2,3-thiadiazines (10).
Molecules 26 02544 sch003
Scheme 4. Hetero-Diels-Alder reactions of aromatic thioketones 1ad with azoalkenes 7a and 7b regioselectively leading to 3,6-dihydro-2H-1,3,4-thiadiazine derivatives 9ah.
Scheme 4. Hetero-Diels-Alder reactions of aromatic thioketones 1ad with azoalkenes 7a and 7b regioselectively leading to 3,6-dihydro-2H-1,3,4-thiadiazine derivatives 9ah.
Molecules 26 02544 sch004
Figure 2. ORTEP Plots [34] of the molecular structures of compounds 9c (left) and 9g (right). Displacement ellipsoids are drawn at the 50% probability level.
Figure 2. ORTEP Plots [34] of the molecular structures of compounds 9c (left) and 9g (right). Displacement ellipsoids are drawn at the 50% probability level.
Molecules 26 02544 g002
Scheme 5. Hetero-Diels-Alder reactions of ferrocenyl-substituted thioketones 1eh with azoalkenes (7a) and (7b) leading to 2-ferrocenyl-substituted 3,6-dihydro-2H-1,3,4-thiadiazines (9ip).
Scheme 5. Hetero-Diels-Alder reactions of ferrocenyl-substituted thioketones 1eh with azoalkenes (7a) and (7b) leading to 2-ferrocenyl-substituted 3,6-dihydro-2H-1,3,4-thiadiazines (9ip).
Molecules 26 02544 sch005
Scheme 6. Elimination of sulfinic acid MeC6H4SO2H from initially formed cycloadducts of 9q,r, leading to 2H-1,3,4-thiadiazine derivatives 11a,b, respectively.
Scheme 6. Elimination of sulfinic acid MeC6H4SO2H from initially formed cycloadducts of 9q,r, leading to 2H-1,3,4-thiadiazine derivatives 11a,b, respectively.
Molecules 26 02544 sch006
Scheme 7. Regio- and periselective formation of the highly substituted cycloadduct 9s from azoalkene 7b and thiochalcone 1l.
Scheme 7. Regio- and periselective formation of the highly substituted cycloadduct 9s from azoalkene 7b and thiochalcone 1l.
Molecules 26 02544 sch007
Scheme 8. Cycloadditions of azoalkene (7a) with aryl-substituted thioketones (1) and pathways for both possible regioisomers (9) and (10).
Scheme 8. Cycloadditions of azoalkene (7a) with aryl-substituted thioketones (1) and pathways for both possible regioisomers (9) and (10).
Molecules 26 02544 sch008
Figure 3. Van der Waals complex of azoalkene 7a and thiobenzophenone (1a) (left side) and its transition state A (right side) leading to 1,3,4-thiadiazine derivative 9a.
Figure 3. Van der Waals complex of azoalkene 7a and thiobenzophenone (1a) (left side) and its transition state A (right side) leading to 1,3,4-thiadiazine derivative 9a.
Molecules 26 02544 g003
Figure 4. Transition states B of cycloadditions of azoalkene 7a with thiobenzophenone (1a) (left side) or thiocyclobutanone derivative 1i (right side) hypothetically leading to 1,2,3-thiadiazine derivative 10a or 10iB, respectively.
Figure 4. Transition states B of cycloadditions of azoalkene 7a with thiobenzophenone (1a) (left side) or thiocyclobutanone derivative 1i (right side) hypothetically leading to 1,2,3-thiadiazine derivative 10a or 10iB, respectively.
Molecules 26 02544 g004
Figure 5. Transition states of the two-step cycloaddition of azoalkene 7a with 3-thioxocyclobutanone derivative 1i; left side: TS A1 leading to zwitterion Z and TS A2 for the cyclization of Z leading to 1,3,4-thiadiazine derivative 9q.
Figure 5. Transition states of the two-step cycloaddition of azoalkene 7a with 3-thioxocyclobutanone derivative 1i; left side: TS A1 leading to zwitterion Z and TS A2 for the cyclization of Z leading to 1,3,4-thiadiazine derivative 9q.
Molecules 26 02544 g005
Scheme 9. Hypothetical cycloaddition of azoalkene 7a with sterically hindered thioketone 1i to 1,2,3-thiadiazine derivative 10q and step-wise reaction involving zwitterion Z leading to 1,3,4-thiadiazine derivative 9q and elimination to 2H-1,3,4-thiadiazine 11a. ((ΔG298) in kcal/mol).
Scheme 9. Hypothetical cycloaddition of azoalkene 7a with sterically hindered thioketone 1i to 1,2,3-thiadiazine derivative 10q and step-wise reaction involving zwitterion Z leading to 1,3,4-thiadiazine derivative 9q and elimination to 2H-1,3,4-thiadiazine 11a. ((ΔG298) in kcal/mol).
Molecules 26 02544 sch009
Table 1. DFT calculations of the cycloadditions of azoalkene 7a with different thioketones 1 (Gibbs free energies (ΔG298) in kcal/mol).
Table 1. DFT calculations of the cycloadditions of azoalkene 7a with different thioketones 1 (Gibbs free energies (ΔG298) in kcal/mol).
EntryStarting MaterialsVan der Waals ComplexesTransition State A1,3,4-Thiadiazine 9Transition State B1,2,3-Thiadiazine 10
17a + 1a
[0.0]
4.38.2−20.819.3−19.3
27a + 1b
[0.0]
3.010.0−24.017.6−22.7
37a + 1m
[0.0]
3.313.1−15.633.6−18.7
47a + 1n
[0.0]
3.911.9−19.433.5−15.3
57a + 1i
[0.0]
4.313.2 [a]
12.0
−19.125.9−22.8
6 [b]7a + 1i
[0.0]
4.332.1−4.1 [c]38.813.8 [d]
[a] See discussion and Scheme 9; [b] Addition to the CO double bond of 1i; [c] 1,3,4-Oxadiazine; [d] 1,2,3-Oxadiazine.
Table 2. DFT-calculated distances of the reacting atoms in the transition states A or B for the cycloadditions of azoalkene 7a with different thioketones 1 leading to 9 or 10.
Table 2. DFT-calculated distances of the reacting atoms in the transition states A or B for the cycloadditions of azoalkene 7a with different thioketones 1 leading to 9 or 10.
EntryStarting MaterialsTransition State A to 1,3,4-Thiadiazine 9 TS-Distances (Å)Transition State B to 1,2,3-Thiadiazine 10 TS-Distances (Å)
C..SN..CC..CS..N
11a + 7a [0.0]2.5783.1992.4932.151
21b + 7a [0.0]2.3933.0972.3702.226
31m + 7a [0.0]2.3193.0292.4712.342
41n + 7a [0.0]2.3733.0442.5462.190
51i + 7a2.2763.5732.4272.366
1.8332.529
61i + 7a [a]1.695 [b]3.214 [b]2.770 [c]1.939 [c]
[a] Addition to the CO double bond of 1i; [b] TS for 1,3,4-oxadiazine; [c] TS for 1,2,3-oxadiazine.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Mlostoń, G.; Urbaniak, K.; Sobiecka, M.; Heimgartner, H.; Würthwein, E.-U.; Zimmer, R.; Lentz, D.; Reissig, H.-U. Hetero-Diels-Alder Reactions of In Situ-Generated Azoalkenes with Thioketones; Experimental and Theoretical Studies. Molecules 2021, 26, 2544. https://doi.org/10.3390/molecules26092544

AMA Style

Mlostoń G, Urbaniak K, Sobiecka M, Heimgartner H, Würthwein E-U, Zimmer R, Lentz D, Reissig H-U. Hetero-Diels-Alder Reactions of In Situ-Generated Azoalkenes with Thioketones; Experimental and Theoretical Studies. Molecules. 2021; 26(9):2544. https://doi.org/10.3390/molecules26092544

Chicago/Turabian Style

Mlostoń, Grzegorz, Katarzyna Urbaniak, Malwina Sobiecka, Heinz Heimgartner, Ernst-Ulrich Würthwein, Reinhold Zimmer, Dieter Lentz, and Hans-Ulrich Reissig. 2021. "Hetero-Diels-Alder Reactions of In Situ-Generated Azoalkenes with Thioketones; Experimental and Theoretical Studies" Molecules 26, no. 9: 2544. https://doi.org/10.3390/molecules26092544

Article Metrics

Back to TopTop