Next Article in Journal
Microwave-Assisted Atom Transfer Radical Cyclization in the Synthesis of 3,3-Dichloro-γ- and δ-Lactams from N-Alkenyl-Tethered Trichloroacetamides Catalyzed by RuCl2(PPh3)3 and Their Cytotoxic Evaluation
Previous Article in Journal
Effects of Central Metal Ion on Binuclear Metal Phthalocyanine-Based Redox Mediator for Lithium Carbonate Decomposition
Previous Article in Special Issue
Design, Synthesis, and Evaluation of B-(Trifluoromethyl)phenyl Phosphine–Borane Derivatives as Novel Progesterone Receptor Antagonists
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Novel Bicyclic P,S-Heterocycles via Stereoselective hetero-Diels–Alder Reactions of Thiochalcones with 1-Phenyl-4H-phosphinin-4-one 1-Oxide †

by
Grzegorz Mlostoń
1,*,
Katarzyna Urbaniak
1,
Marcin Palusiak
2,
Elżbieta Łastawiecka
3,
Sławomir Frynas
3,
Kazimierz Michał Pietrusiewicz
3,* and
Heinz Heimgartner
4
1
Department of Organic and Applied Chemistry, Faculty of Chemistry, University of Lodz, Tamka 12, 91-403 Lodz, Poland
2
Department of Physical Chemistry, Faculty of Chemistry, University of Lodz, Pomorska 163/165, 90-236 Lodz, Poland
3
Department of Organic and Crystal Chemistry, Institute of Chemical Sciences, Faculty of Chemistry, Maria Curie-Skłodowska University, Gliniana 33, 20-614 Lublin, Poland
4
Department of Chemistry, University of Zurich, Winterthurerstrasse 190, CH-8057 Zurich, Switzerland
*
Authors to whom correspondence should be addressed.
In memory of Professor Jan Epsztajn (Lodz).
Molecules 2024, 29(9), 2036; https://doi.org/10.3390/molecules29092036
Submission received: 15 March 2024 / Revised: 23 April 2024 / Accepted: 24 April 2024 / Published: 28 April 2024
(This article belongs to the Special Issue Recent Development of Organophosphorus Chemistry)

Abstract

:
Thiochalcones undergo cycloaddition reactions in THF solution at 60 °C with the synthetically unexplored 1-phenyl-4H-phosphinin-4-one 1-oxide in a highly regio- and stereoselective manner, yielding hitherto unknown bicyclic P,S-heterocycles containing fused thiopyran and phosphinine rings. The stereochemical structures of two of the obtained (4+2)-cycloadducts were unambiguously assigned by means of the X-ray single-crystal analysis. Based on these assignments, a concerted mechanism of the hetero-Diels–Alder reaction via the preferred endo approach of the heterodiene from the less hindered P=O side of the phosphininone molecule is postulated to explain the established rac-(4RS,8SR,9SR,10SR)-configured (4+2)-cycloadducts isolated as major products.

Graphical Abstract

1. Introduction

Aromatic thioketones such as thiofluorenone (1a) and thiobenzophenone (1b) are recognized as superior reagents for trapping 1,3-dipoles and dienes, hence termed as ‘superdipolarophiles’ (Rolf Huisgen 1995) [1,2,3] and ‘superdienophiles’ (Jürgen Sauer 1998) [4,5], respectively. The α,β-unsaturated analogs of aromatic thioketone 1, known as thio-chalcones, exist in solution as a mixture of the monomeric 2 and two dimeric forms, 3 and 3′ (Figure 1). In general, contrary to chalcones, they have been less explored in the current organic synthesis, particularly in the cycloaddition chemistry (e.g., [6,7,8]).
However, in our recent publications, diverse thiochalcones (as monomeric forms 2) bearing aryl, hetaryl, or ferrocenyl groups have been described for the first time as active dipolarophiles in (3+2)-cycloadditions with electron-deficient, fluorinated nitrile imines [9] and as dienophiles in (4+2)-cycloadditions with electron-deficient α-nitroso alkenes [10]. Additionally, thiochalcones were successfully applied as dienophiles or S-heterodienes in some asymmetric hetero-Diels–Alder reactions [11,12]. Moreover, they smoothly undergo (4+2)-cycloaddition reactions with acetylenic carboxylates [13,14] and some 1,4-quinones, e.g., with naphthoquinone to give fused 4H-thiopyran 4 in high yields after spontaneous oxidation of the initial (4+2)-cycloadducts [15] (Scheme 1).
Many years ago, a rare six-membered phosphorus heterocycle (4H-phosphinin-4-one) was reported as a stable 1-oxide 5a [16]. In recent publications, the synthesis of its dihydro derivative 5b was also described [17,18] (Figure 2). The structures of both compounds resemble the 1,4-quinone framework, strongly suggesting their potential reactivity toward dienes/heterodienes and, thus, a possible application for the synthesis of new, fused P-heterocyclic systems. This expectation is reinforced by the fact that five-membered ring congeners of phosphinine 1-oxides 5a and 5b, namely phospholene 1-oxide 6a and its derivatives 6b and 6c, as well as the dihydrophosphinine 1-oxide 7, have already been reported to act as dieno- and dipolarophiles in the cycloaddition chemistry (Scheme 2).
Cycloaddition reactions involving the exploration of C,P-heterocycle 1-oxides as dieno- or dipolarophiles are rarely described. Representative examples of such reactions, leading to a variety of polycylic P-heterocycles, as illustrated in Scheme 2, Scheme 3 and Scheme 4, have already been reported.
As shown in Scheme 2, the highly activated phospholene 1-oxide 6a easily enters the Diels–Alder reaction with 7-methoxy-4-vinyl-1,2-dihydronaphthalene (diene), stereoselectively yielding the tetracyclic cycloadduct 8 with a 15-P-steroid structure in 25% yield (Scheme 2A) [19]. Similarly, phospholene 1-oxide 6b was reported to undergo the Diels–Alder reaction with in situ-generated α-oxy-o-xylylene (diene), leading to the tricyclic P-heterocycle 9 in 90% yield (Scheme 2B) [20].
The same substrate 6b was successfully employed as a dipolarophile in the (3+2)-cycloaddition with an acyclic nitrone (C-phenyl-N-methylnitrone), yielding the P-containing fused heterocycle 10 in 72% yield with high stereoselectivity (Scheme 3) [21,22].
Furthermore, the phospholene 1-oxide 6c undergoes dimerization according to the rules of (4+2)-cycloadditions, yielding the tricyclic P-heterocycle 11 in suitable yield (66%) with high stereoselectivity [23,24] (Scheme 4A).
The only known example of a (4+2)-cycloaddition involving the six-membered phosphinine 1-oxide 7 is depicted in Scheme 4B. However, in this instance, the observed dimerization-like cycloaddition did not occur between two molecules of 7. Instead, 7 was trapped by its double bond isomer 7′ present in the reaction mixture, leading to the stereo-selective formation of the tricyclic product 12 (Scheme 4B) [25]. This unexpected reaction provides an intriguing example of so-called cross-dimerizations.
The objective of the present study was to investigate the reactivity of 4H-phosphinin-4-one 1-oxides 5a and 5b toward selected thiochalcone 2, which were employed as reactive heterodienes. Specifically, we aimed to elucidate the regiochemistry of the anticipated hetero-Diels–Alder reactions and the stereochemical structure of the resulting (4+2)-cycloadducts containing two heteroatoms (P and S). Additionally, the investigation of the reaction mechanism was of particular interest.

2. Results and Discussion

The test experiment was carried out in THF solution at 60 °C using 1-phenyl-3-(thien-2-yl)propene-1-thione (2a) and 1-phenylphosphinin-4-one 1-oxide (5a) in a 1:1 ratio. TLC analysis revealed complete consumption of 2a after approximately 20 h. After the removal of THF in vacuum, the crude reaction mixture was analyzed by 1H NMR, indicating the presence of two products formed in unequal amounts. The major product exhibited three signals in the non-aromatic region at 3.63–3.66, 4.03–4.05, and 4.54–4.57 ppm, which can be attributed to aliphatic protons. Additionally, three similar but less intense signals were observed for the minor product at 3.69‒3.71, 4.06–4.08, and 4.49–4.50 ppm, respectively (Supplementary Figure S1). The close resemblance of the corresponding signals of the major and minor products strongly suggested that they shared the same framework with identical positions of both heteroatoms (S and P), differing most likely only in the relative stereochemistry at stereogenic centers. A comparison of the intensities of the integration lines of multiplets attributed to the major and minor components suggested a ca. 7:1 ratio of these isomeric cycloadducts.
Chromatographic separation of the crude reaction mixture on a SiO2 column afforded the major product with a melting point of 184 °C (decomposition) in a yield of 57%, using a petroleum ether/dichloromethane mixture as the eluent. The 31P NMR spectrum showed a resonance signal at 20.9 ppm for the P atom. The HRMS spectrum of this product revealed a molecular mass [M + 1] = 435.0642, corresponding to the molecular formula C24H20O2PS2, indicative of the (4+2)-cycloadduct of 2a and 5a. Finally, single-crystal X-ray diffraction analysis confirmed the structure of the fused P,S-heterocycle 13a, with the heteroatoms located at S(1) and P(8) positions, respectively (for conventional atom numbering, see Scheme 5). In contrast to quinones (Scheme 1), no spontaneous oxidation of the initially formed cycloadducts was observed. Based on the X-ray analysis, the stereochemical structure of this diastereoisomer was determined as rac-(4RS,8SR,9SR,10SR) (Scheme 5 and Figure 3).
The aforementioned three sets of signals of the minor product could tentatively be attributed to an isomeric cycloadduct 14a with rac-(4SR,8SR,9SR,10SR) configuration (Scheme 5). However, attempts to isolate this product by column chromatography were unsuccessful.
Reactions of thiochalcones 2b2d with 5a (molar ratio 1:1) were carried out under analogous conditions to those described for 2a. In all cases, as in the experiment with 2a, the 1H NMR spectra obtained for the crude reaction mixtures revealed the presence of two isomeric cycloadducts of types 13 and 14, exhibiting similar sets of signals in the previously described high-field regions of the 1H NMR spectra. However, in all cases, the yields of the postulated cycloadducts 14b14d were consistently very low (<10%), and their attempted isolation was again unsuccessful. The major product 13b, bearing two phenyl substituents, was isolated chromatographically in 51% yield. Subsequent crystallization from petroleum ether/CH2Cl2 yielded single crystals suitable for the X-ray analysis. The analysis confirmed both the regiochemical structure of 13b and the configurations of the four stereogenic centers at C(4), P(8), C(9), and C(10), indicating again the endo mode of addition in full analogy to 13a (Scheme 5 and Figure 4).
It is worth mentioning that in separate NMR experiments, the stereochemical stability of 13b was tested in CDCl3 solution at room temperature and at 60 °C, and in both cases, unchanged 13b was observed thereafter. Moreover, addition of catalytic amounts of either TFA or pyridine to the sample did not result in any detectable isomerization of 13b, neither at room temperature nor after heating at 60 °C. The observed lack of isomerization thus allowed us to exclude the possibility that the obtained minor products were formed via keto-enol equilibration of the corresponding major products under the reaction conditions. Considering the complete face selectivity recorded thus far for all known cycloadditions of cyclic organophosphorus dienophiles, which are exclusively attacked by a diene from their P=O bearing side (cf. Scheme 5 and Scheme 6), it became most plausible that the stereochemistry of the minor product 14 reflects the exo mode of their formation, as tentatively assigned (cf. Scheme 5).
Notably, the attempted hetero-Diels–Alder reaction of 5a with the isomeric thiochalcone 2′a (3-phenyl-1-(thien-2-yl)propene) in THF at 60 °C was unsuccessful, resulting in the decomposition of the starting materials. Apparently, the substitution pattern in the starting thiochalcone also influences the outcome of the studied reactions.
Similarly, the attempted (4+2)-cycloaddition of thiochalcone 2b with 2,3-dihydro-4H-phosphinin-4-one 1-oxide 5b was unsuccessful, and no defined product could be identified in the crude reaction mixtures.
Structure analysis of cycloadducts 13a and 13b: In our research endeavor, we successfully acquired crystal samples of compounds 13a and 13b, which proved to be suitable for conducting X-ray single-crystal experiments. Both compounds exhibit crystal structures within the centrosymmetric P21/c space group. Consequently, the crystal samples applied in the analysis contain racemic mixtures of inversion-equivalent enantiomeric forms, aligning with the anticipated outcome within the context of our synthesis procedure. For a detailed insight into the absolute configuration of all four chiral centers in molecules of 13a and 13b, refer to Scheme 5. Furthermore, our crystallographic analysis revealed the presence of two molecules within the symmetry-independent cell unit in both crystals. There is no pseudo-symmetry between the molecules in the asymmetric unit. Each molecule possesses its own structural characterization in both experimentally treated samples.
Remarkably, these molecules within the symmetry-independent unit display identical absolute configurations.
The molecular architecture of both compounds is characterized by two fused heterocyclic rings, as depicted in Figure 3 and Figure 4. Notably, the C(9) and C(10) centers serve as common elements in both heterocycles. The hydrogen atoms attached to these carbon atoms adopt a cis configuration, leading to a distinctive nonplanar geometry of this primary molecular moiety and, consequently, influencing the overall molecular structure. This nonplanarity is reflected in the formation of dihedral angles of 79.74° and 75.16° for compounds 13a and 13b, respectively. In both crystal samples, the heterocyclic rings adopt a half-boat conformation, with the C(9) and C(10) atoms positioned above the mean plane of the rings. The distance between the carbon atoms at the apex position and the planar fragments of the rings is in the range of 0.654(3)–0.756(3) Å. Importantly, the heteroatoms S and P are always in the plane of their parent rings.
The planarity of other rings conforms to expectations, with their positioning relative to the bearing fragment determined by crystal packing, exhibiting no anomalous behavior.
An intriguing observation arises from the absence of classic H-bond-donating groups in both molecules, thereby precluding the presence of traditional H-bonds (e.g., O–HO or N–HO type) responsible for stabilizing the crystal lattice. Instead, we identified numerous weak short contacts of C–HA (where A = S and O), potentially serving as stabilizing interactions. Notably, our attention is particularly drawn to a short contact of P=OCsp2 type, with OC distances of 2.994(2) Å and 3.006(2) Å for compounds 13a and 13b, respectively. The corresponding P=OC angles are measured to be 130.6(1)° and 137.5(1)°. Importantly, the involved carbon atoms belong to strongly polarized CO carbonyl groups, suggesting that these interactions may rival hydrogen bonding in terms of strength. As already postulated [26], such interactions exhibit a comparable strength to hydrogen bonding and are notably directional due to the partial atomic charge distribution within the interacting fragments. We are expecting that this specific P=OC interaction may be a leading one among all, which stabilizes crystals of 13a and 13b.
Mechanistic consideration: The (4+2)-cycloadditions of electron-rich thiochalcone 2 with the phosphorus-containing dienophile 5a, considered as an electron-deficient dienophile, are reported for the first time. Therefore, the mechanism of these hetero-Diels–Alder reactions depicted in Scheme 6 deserves a brief comment.
In recent decades, mechanisms of cycloaddition reactions have been studied intensively. Alongside the ‘classical’ interpretation based on the assumption of concerted pathways, stepwise processes via zwitterionic or diradical intermediates have also been discussed [27,28,29,30]. It is well known that the large energy gap between HOMO and LUMO energies of both reactants, i.e., 1,3-dipole and dipolarophile in the case of (3+2)-cycloadditions [2], and diene and dienophile in the case of Diels–Alder reactions [31], favors stepwise mechanisms.
In our opinion, the studied reactions follow the classical concerted (but asynchronous) mechanism, and the preferred endo-attack of the heterodiene from the less hindered P=O side (face selectivity) of the organophosphorus compound leads to the isolated major cycloadduct rac-(4RS,8SR,9SR,10SR)-13 (Scheme 6). The reactions occur with complete regioselectivity and lead to the formation of the C–S bond exclusively at the β-C atom. Moreover, it has already been well documented that the α,β-unsaturated P-heterocyclic dienophiles undergo cycloaddition reactions with 1,3-dipoles [21,22] as well as with dienes [19,20] via the approach from the P=O bearing face, and this spatial preference is observed for both the endo and the exo mode of addition [21,22]. The alternative exo attack results in the formation of (4+2)-cycloadduct 14 with an inverted configuration at the C(4) atom (Scheme 6).

3. Materials and Methods

3.1. Experimental: X-ray Structure Determination of 13a and 13b

X-ray diffraction data for 13a and 13b were collected on an XtaLAB Synergy, Dualflex, HyPix diffractometer (Rigaku, Akishima, Japan). Integration of the intensities and corrections for Lorentz effects, polarization effects, and analytical absorption were performed with CrysAlis PRO [32]. Using Olex2 [33], the structures were solved with the SHELXT [34] structure solution program using Intrinsic Phasing and refined with the SHELXL [35] refinement package using Least Squares minimization. The hydrogen atoms were introduced in the calculated positions with an idealized geometry and constrained using a rigid body model with isotropic displacement parameters equal to 1.2 of the equivalent displacement parameters of their parent atoms. The molecular geometries were calculated by the PLATON program [36]. The relevant crystallographic data are provided in Table S1. Atomic coordinates, displacement parameters, and structural factors of the analyzed crystal structures are deposited with the Cambridge Crystallographic Data Centre CCDC (reference numbers: 2298978 and 2298979) [37] (for further details, see Supplementary Materials).

3.2. General Information

Commercial chemicals and solvents were used as received. If not stated otherwise, products were purified by filtration through short silica gel plugs (200–400 mesh) by using freshly distilled solvents as eluents or by recrystallization. Melting points were determined in capillaries with an Aldrich Melt-Temp II (St. Louis, MO, USA), and they are uncorrected. NMR spectra were taken with a Bruker AVIII spectrometer (Billerica, MA, USA, 1H NMR (600 MHz), 13C NMR (151 MHz), and 31P NMR (243 MHz)); chemical shifts (δ) are expressed in parts per million, and they relate to residual undeuterated solvent peaks (CDCl3: 1H NMR δ = 7.26, 13C NMR δ = 77.16) or to an external standard (H3PO4: 31P NMR δ = 0.00). IR spectra are presented in cm‒1, and they were measured with an Agilent Cary 630 FTIR spectrometer (Santa Clara, CA, USA) in neat. Mass spectra (ESI) were registered with a Varian 500-MS LC Ion Trap (Palo Alto, CA, USA). Elemental analyses were obtained with a Vario EL III (Elementar Analysensysteme GmbH, Langenselbold, Germany) instrument.
Starting materials: Thiochalcones 2a2d were prepared by treatment of the corresponding chalcones [38,39] with Lawesson’s reagent in boiling THF for 2.5–3 h. Crude products were purified by column chromatography and used for the studied reactions as unseparated mixtures of monomeric and dimeric forms 2/3/3′ [9,13]. 1-Phenyl-4H-phosphinin-4-one 1-oxide (5a) and 1-phenyl-2,3-dihydro-4H-phosphinin-4-one 1-oxide (5b) were synthesized following the published procedures [17,18].
General procedure for reactions of thiochalcones 2a2d with 4H-phosphinin-4-one 1-oxide 5a: A solution containing 1.1 mmol of the corresponding thiochalcone 2 and 1 mmol of 4H-phosphinin-4-one 1-oxide 5a dissolved in 4 mL of dry THF was heated in an oil bath at 60 °C overnight (about 20 h). After this time, the TLC test showed that starting 5a was completely consumed. The solvent was evaporated, and the remaining mixture, after the 1H NMR examination, was purified by column chromatography (petroleum ether with increasing amounts of CH2Cl2). Analytically pure products were obtained by crystallization from petroleum ether/CH2Cl2.
Structures of the isolated cycloadducts are presented in Figure 5, Figure 6, Figure 7 and Figure 8.
  • rac-RS,8SR,9SR,10SR)-2,8-Diphenyl-4-(thiophen-2-yl)-4,4a,8a-trihydrophosphinino [2,3-b]thiopyran-5-one 8-oxide (13a). Yield = 250 mg (57%), pale yellow crystals, mp = 184 °C (decomp.) (petroleum ether/CH2Cl2).
1H NMR (CDCl3): δ 3.63–3.66 (m, 1CH); 4.03 (brs, 1CH); 4.55 (td, J = 15.6 Hz, J = 2.7 Hz, 1CH); 6.36 (d, J = 2.4 Hz, 1CH); 6.77 (dd, J = 37.1 Hz, J = 12.9 Hz, 1CH); 6.87‒6.92 (m, 1CH); 6.97 (dd, J = 5.1 Hz, J = 3.5 Hz, 1CH); 7.04 (d, J = 3.5 Hz, CH); 7.22 (dd, J = 5.1 Hz, J = 1.0 Hz, 1CH); 7.33‒7.38 (m, 3CHarom); 7.51–7.54 (m, 2CHarom); 7.60–7.68 (m, 2CH); 7.75 (dt, J = 7.4 Hz, J = 1.2 Hz, 1CH); 7.92–7.96 (m, 2CH).
13C{1H} NMR (CDCl3): δ 41.2 (JC,P = 10 Hz), 48.6 (JC,P = 4 Hz) (2d, C(4) and C(10)—assignment uncertain); 44.5 (d, 1JC,P = 66 Hz, C(9)); 120.8, 124.6, 126.7, 128.5, 128.7, 130.9, 131.5, 138.4, 143.0 (9 signals), 126.3, (d, 1JC,P = 47 Hz, C(7)), 129.7 (d, JC,P = 15 Hz), 131.1 (d, JC,P = 2 Hz), 133.8 (d, JC,P = 2 Hz), 128.5, (d, 1JC,P = 105 Hz, Car), 132.0 (d, J = 7 Hz); 144.2 (d, J = 2 Hz), 191.4 (d, 3JC,P = 11 Hz, C=O).
31P NMR (CDCl3): δ 20.9.
IR: 3024m, 3005m, 1694s (C=O), 1437s, 1239m, 1187m, 1168s, 1112m, 830m, 711s, 693vs.
HRMS for [M + 1]+ [C24H20O2PS2]: calc.: 435.0642; found: 435.0646.
EA for C24H19O2PS2 (434.51): calc. C 66.34, H 4.41, S 14.76 found C 66.09, H 4.54, S 14.98.
  • rac-(RS,8SR,9SR,10SR)-2,4,8-Triphenyl-4,4a,8a-trihydrophosphinino[2,3-b]thiopyran-5-one 8-oxide (13b). Yield: 220 mg (51%), beige crystals, mp = 212 °C (decomp.) (petroleum ether/CH2Cl2).
1H NMR (CDCl3): δ 3.63–3.66 (m, 1H); 3.78 (brs, 1H); 4.57–4.61 (m, 1H); 6.46 (d, J = 2.2 Hz, 1H); 6.69 (dd, J = 37.4 Hz, JH,H = 13.0 Hz, 1H); 6.86–6.89 (m, 1H); 7.25–7.27 (m, 1H); 7.34–7.40 (m, 5H); 7.43–7.44 (m, 2H); 7.57–7.59 (m, 2H); 7.65–7.68 (m, 2H); 7.74–7.76 (m, 1H); 7.92–7.96 (m, 2H).
13C{1H} NMR (CDCl3): δ 44.9 (d, 1JC,P = 73 Hz, C(9)), 45.5 (JC,P = 2 Hz), 48.5 (JC,P = 5 Hz) (2d, C(4) and C(10)—assignment uncertain); 120.5, 126.7, 126.9, 128.1, 128.5, 128.6 129.2, 138.8, 143.3 (9 signals); 128.6 (d, 1JC,P = 104 Hz, Car), 129.6 (d, 2JC,P = 15 Hz), 131.05 (d, 2JC,P = 10 Hz, =C(6)), 131.1 (d, 1JC,P = 87 Hz, C(7)), 132.6 (d, JC,P = 7 Hz), 133.7 (d, JC,P = 2 Hz), 140.7 (d, JC,P = 2 Hz); 191.8 (d, 3JC,P = 10 Hz, C=O).
31P NMR (CDCl3): δ 21.1.
IR: 2997m, 2893m, 1653s (C=O), 1567m, 1495m, 1432m, 1123vs, 1098m, 767vs, 698vs, 513s.
HRMS for [M + 1]+ [C26H22O2PS]: calc.: 429.1078; found: 429.1066.
  • rac-(RS,8SR,9SR,10SR)-2,8-Diphenyl-4-(4-methylphenyl)-4,4a,8a-trihydrophosphinino[2,3-b]thiopyran-5-one 8-oxide (13c). Yield: 245 mg (51%), cream-colored crystals, mp = 201 °C (decomp.) (petroleum ether/CH2Cl2).
1H NMR (CDCl3): δ 2.35 (s, CH3); 3.60–362 (m, 1CH); 3.74 (brs, 1CH); 4.57 (d, J = 12.9 Hz, 1CH); 6.44 (brs, 1CH); 6.68 (dd, J = 37.3 Hz, J = 12.9 Hz, 1CH); 6.84–6.88 (m, 1CH); 7.16, 7.56 (AB-signal pattern, J = 7.7 Hz, 4CHarom); 7.32–7.39 (m, 5CH); 7.66–7.68 (m, 2CH); 7.73–7.76–(m, CH); 7.91–7.95 (m, 2CH).
13C{1H} NMR (CDCl3): δ 21.0 (CH3), 44.8 (d, 1JC,P = 66 Hz, C(9)); 44.7 (JC,P = 9 Hz), 48.5 (d, JC,P = 4 Hz) (2d, C(4) and C(10)—assignment uncertain); 120.7, 126.7, 128.5, 128.6, 128.8, 129.1, 136.6, 138.8, 143.4 (9 signals), 128.7 (d, 1JC,P = 104 Hz Car), 129.5 (d, J = 12 Hz), 130.5 (1JC,P = 83 Hz), 131.1 (d, J = 10 Hz), 132.2 (d, J = 7 Hz),133.7 (d, J = 2 Hz), 137.5 (d, J = 2 Hz), 191.9 (d, 3JC,P = 10 Hz, C=O).
31P NMR (CDCl3): δ 21.4.
IR: 3023m, 3011m, 2897m, 1702s (C=O); 1593m, 1501s, 1493m, 1234m, 1166m, 892s, 745s, 693vs.
HRMS for [M + 1]+ [C27H24O2PS]: calc.: 443.1235; found: 443.1235.
EA for C27H23O2PS (442.12): calc. C 73.29, H 5.24, S 7.24; found C 73.21, H 5.17, S 7.10.
  • rac-(RS,8SR,9SR,10SR)-4-(4-Bromophenyl)-2,8-diphenyl-4,4a,8a-trihydrophosphinino[2,3-b]thiopyran-5-one 8-oxide (13d). Yield: 310 mg (61%), yellow crystals, mp = 178 °C (decomp.) (petroleum ether/CH2Cl2).
1H NMR (CDCl3): δ 3.56‒3.59 (m, 1CH); 3.79 (brs, 1CH); 4.55 (td, J = 15.4 Hz, J = 2.7 Hz, 1CH); 6.36 (d, J = 2.4 Hz, 1CH); 6.69 (dd, J = 37.3 Hz, J = 12.9 Hz, 1CH); 6.86–6.90 (m, 1CH); 7.31, 7.46 (AB-signal pattern, J = 8.5 Hz, 4CHarom); 7.36–7.40 (m, 3CHarom); 7.55–7.58 (m, 2CHarom); 7.65–7.69 (m, 2CHarom); 7.75–7.77 (m, 1CHarom); 7.91–7.95 (m, 2CHarom).
13C{1H} NMR (CDCl3): δ 44.5 (d, 1JC,P = 66 Hz, C(9)), 44.4 (JC,P = 9 Hz), 48.4 (JC,P = 4 Hz) (2d, C(4) and C(10)—assignment uncertain); 119.7, 120.8, 126.6, 128.6, 128.8, 131.1, 131.5, 138.7, 143.2 (9 singlet signals); 128.5 (d, 1JC,P = 104 Hz, Car) 129.5 (d, J = 12 Hz), 129.6 (d, 1JC,P = 83 Hz, C(7)), 131.2 (d, J = 8 Hz), 133.8 (d, J = 3 Hz), 133.2 (d, J = 7 Hz), 139.7 (d, J = 2 Hz), 191.9 (d, 3JC,P = 10 Hz, C=O).
31P NMR (CDCl3): δ 20.9.
IR: 3013m, 3009m, 2894m, 1692s (C=O); 1589m, 1487s, 1434m, 1183m, 1110m, 883m, 728s, 691vs.
HRMS for [M + 1]+ [C26H21BrO2PS]: calc.: 507.0183; found: 507.0200.
EA for C26H20BrO2PS (506.01): calc. C 61.55, H 3.97, S 6.32 found C 61.56, H 3.92, S 6.33.
Attempted reaction of thiochalcone 2b with 2,3-dihydro-4H-phosphinin-4-one 1-oxide (5b): A solution containing 246 mg (1.1 mmol) of thiochalcone 2b and 206 mg (1 mmol) of phosphinin-4-one 1-oxide 5b dissolved in 4 mL of dry THF was heated overnight at 60 °C (about 20 h). The formation of tarry products was observed, and no defined product could be isolated from the crude reaction mixture.
Preparation of crystals of 13a and 13b for X-ray measurement: Suitable crystals of 13a and 13b were obtained by slow evaporation of the solvent from a hexane/dichloromethane solution under room conditions. Samples used for crystallization were initially purified by the PLC method directly from the crude reaction mixture. Once the crystals were formed, samples of single crystals suitable for X-ray measurements were selected using a stereomicroscope with polarized-light functionality. During the X-ray measurement, crystals were mounted on loops using dedicated oil.

4. Conclusions

The presented work should be regarded as a continuation of a series of our studies on the preparation and structural investigation of new P-functionalized sulfur heterocycles and P,S-heterocycles. This series is based on the utilization of thiocarbonyl compounds as universal building blocks for diverse cycloaddition reactions, showcasing the significant utility of relatively unexplored thiochalcones [40,41,42].
The current study demonstrated that hitherto unexplored, prochiral phosphinin-4-one 1-oxide 5a undergoes hetero-Diels–Alder reactions with thiochalcones 2 in a regiospecific and highly stereoselective manner, thereby providing access to the previously unknown P,S-heterocycles 13. These compounds are identified as new, fused derivatives of thiopyran and phosphinine. Both diverse phosphorus-containing heterocycles and thio-pyrans are recognized as important pharmacophores, bio-isosteres, and prodrugs [43,44,45], and therefore, the described cycloadducts 13 may be of potential interest for further investigations as biologically active compounds, particularly in medicinal chemistry and crop protection agriculture science. Additionally, the described chiral P,S-heterocycles, available in enantiopure form, may be of interest for phosphorus-based catalysis, an area undergoing dynamic development in recent decades [46].
In summary, this work complements the latest contributions on synthetically useful applications of thiochalcones [47] and underlines their significance in the development of methods for the preparation of novel, sulfur-containing heterocycles.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/molecules29092036/s1: Copies of the 1H, 13C, and 31P NMR spectra of all new compounds 13a13d, as well as the X-ray structure determination for cycloadduct 13a,b.
Figure S1. Fragment of the 1H NMR spectra registered for unseparated crude mixture obtained after (4+2)-cycloaddition of 2a and 5a. Figure S2. The 1H NMR spectrum for cycloadduct 6a. Figure S3. The 13C NMR spectrum for cycloadduct 6a. Figure S4. The 31P NMR spectrum for cycloadduct 6a. Figure S5. The 1H NMR spectrum for cycloadduct 6b. Figure S6. The 13C NMR spectrum of cycloadduct 6b. Figure S7. The 31P NMR spectrum registered for cycloadduct 6b. Figure S8. The 1H NMR spectrum registered for cycloadduct 6c. Figure S9. The 13C NMR spectrum registered for cycloadduct 6c. Figure S10. The 31P NMR spectrum registered for cycloadduct 6c. Figure S11. The 1H NMR spectrum registered for cycloadduct 6d. Figure S12. The 13C NMR spectrum registered for cycloadduct 6d. Figure S14. Molecular structure of the 4-(thien-2-yl) substituted cycloadduct 6a. Atoms are represented by thermal ellipsoids (50%). Figure S15. Molecular structure of the 4-phenyl substituted cycloadduct 6b. Atoms are represented by thermal ellipsoids (50%). Table S1. Crystal data and structure refinement for 6a and 6b.

Author Contributions

Conceptualization, G.M. and K.M.P.; methodology, organic synthesis, G.M. K.U., E.Ł. and S.F.; crystallography, M.P.; investigation, synthesis, K.U. and G.M.; writing—original draft preparation, G.M., K.M.P., M.P. and H.H.; writing—review and editing, G.M. and H.H.; supervision, G.M. and K.M.P. All authors have read and agreed to the published version of the manuscript.

Funding

The presented research was funded by the University of Lodz (Poland) within project IDUB 2023–2024.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data are contained within the article and Supplementary Materials.

Acknowledgments

The authors thank Jakub Wręczycki (Lodz) for his help in the last stage of preparation of this manuscript for submission. Skillful help in laboratory work by Małgorzata Celeda (University of Lodz) is acknowledged.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. 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, 9671–9678. [Google Scholar] [CrossRef]
  2. Fisera, L.; Huisgen, R.; Kalwinsch, I.; Langhals, E.; Li, X.; Mlostoń, G.; Polborn, K.; Rapp, J.; Sicking, W.; Sustmann, R. New thione chemistry. Pure Appl. Chem. 1996, 68, 789–798. [Google Scholar] [CrossRef]
  3. 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]
  4. Breu, J.; Höcht, P.; Rohr, U.; Schatz, J.; Sauer, J. Thiocarbonyl Compounds in [4+2] Cycloadditions: Preparative Aspects. Eur. J. Org. Chem. 1998, 861–2873. [Google Scholar] [CrossRef]
  5. Rohr, U.; Schatz, J.; Sauer, J. Thio- and Selenocarbonyl Compounds as “Superdienophiles” in [4+2] Cycloadditions. Eur. J. Org. Chem. 1998, 2875–2883. [Google Scholar] [CrossRef]
  6. Karakasa, T.; Motoki, S. Chemistry of α,β-Unsaturated Thione Dimers. 1. Preparation of α,β-Unsaturated Thione Dimers and Thermolysis of These Dimers in the Presence of Acrylonitrile or Acrylamide. J. Org. Chem. 1978, 43, 4147–4150. [Google Scholar] [CrossRef]
  7. Karakasa, T.; Motoki, S. Chemistry of α,β-Unsaturated Thione Dimers. 2. Reactions of Thiochalcones and 2-Arylidene-1-thiotetralones with Some Olefins and the Parent Ketones of the Thiones. J. Org. Chem. 1979, 44, 4151–4155. [Google Scholar] [CrossRef]
  8. Motoki, S.; Saito, T.; Karakasa, T.; Kato, H.; Matsushita, T.; Hayashibe, S. Thermal and Lewis Acid-promoted Novel Asymmetric Hetero Diels-Alder Reactions of a 1-Thiabuta-1,3-diene System (Thiochalcones) with (–)-Dimenthyl fumarate. J. Chem. Soc. Perkin Trans. 1 1991, 2281–2283. [Google Scholar] [CrossRef]
  9. Mlostoń, G.; Grzelak, P.; Utecht, G.; Jasiński, M. 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] [CrossRef]
  10. 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, e2544. [Google Scholar] [CrossRef]
  11. Hejmanowska, J.; Jasiński, M.; Mlostoń, G.; Albrecht, Ł. Taming of Thioketones: The First Asymmetric Thia-Diels-Alder reaction with Thioketones as Heterodienophiles. Eur. J. Org. Chem. 2017, 950–954. [Google Scholar] [CrossRef]
  12. Hejmanowska, J.; Jasiński, M.; Wojciechowski, J.; Mlostoń, G.; Albrecht, Ł. The first organocatalytic, ortho-regioselective inverse-electron-demand hetero-Diels-Alder reaction. Chem. Commun. 2017, 53, 11472–11475. [Google Scholar] [CrossRef] [PubMed]
  13. Mlostoń, G.; Grzelak, P.; Heimgartner, H. Hetero-Diels-Alder reactions of hetaryl thiochalcones with acetylenic dienophiles. J. Sulfur Chem. 2017, 38, 1–10. [Google Scholar] [CrossRef]
  14. Mlostoń, G.; Hamera-Fałdyga, R.; Heimgartner, H. First synthesis of ferrocenyl-substituted thiochalcones and their [4 + 2]-cycloadditions with acetylenic dipolarophiles. J. Sulfur Chem. 2018, 39, 322–331. [Google Scholar] [CrossRef]
  15. Mlostoń, G.; Urbaniak, K.; Urbaniak, P.; Marko, A.; Linden, A.; Heimgartner, H. First thia-Diels-Alder reactions of thiochalcones with 1,4-quinones. Beilstein J. Org. Chem. 2018, 14, 1834–1839. [Google Scholar] [CrossRef]
  16. Märkl, G.; Olbricht, H. Derivates of Phospha-2,5-cyclohexadien-4-one and 4,4′-Diphosphabi-(2,5-cyclohexadienylidene). Angew. Chem. Int. Ed. 1966, 5, 588–589. [Google Scholar] [CrossRef]
  17. Łastawiecka, E.; Frynas, S.; Pietrusiewicz, K.M. Desymmetrization Approach to the Synthesis of Optically Active P-Stereogenic Phosphin-2-en-4-ones. J. Org. Chem. 2021, 86, 6195–6206. [Google Scholar] [CrossRef] [PubMed]
  18. Łastawiecka, E.; Włodarczyk, A.; Kozioł, A.E.; Małuszyńska, H.; Pietrusiewicz, K.M. Resolution of P-Stereogenic 1-Phenylphosphin-2-en-4-one 1-Oxide into Two Enantiomers by (R,R)-TADDOL and Conformational Diversity of the Phosphinenone Ring and TADDOL in the Crystal State. Molecules, 2021; 26, 6873. [Google Scholar] [CrossRef]
  19. Kashman, Y.; Awerbouch, O. Phosphasteroids-I. Cycloaddition reactions of phospholenes with various dienes. Tetrahedron Lett. 1973, 14, 3217–3220. [Google Scholar] [CrossRef]
  20. Pietrusiewicz, K.M.; Koprowski, M.; Drzazga, Z.; Parcheta, R.; Łastawiecka, E.; Demchuk, O.M.; Justyniak, I. Efficient Oxidative Resolution of 1-Phenylphosphol-2-ene and Diels-Alder Synthesis of Enantiopure Bicyclic and Tricyclic P-Stereogenic C-P Heterocycles. Symmetry 2020, 12, 346. [Google Scholar] [CrossRef]
  21. Goti, A.; Cicchi, S.; Brandi, A.; Pietrusiewicz, K.M. Nitrone Cycloadditions to 2,3-Dihydro-1-phenyl-1H-phosphole 1-Oxide. Double Asymmetric Reaction and Kinetic Resolution by a Chiral Nitrone. Tetrahedron: Asymmetry 1991, 2, 1371–1378. [Google Scholar] [CrossRef]
  22. Pietrusiewicz, K.M.; Hołody, W.; Koprowski, M.; Cicchi, S.; Goti, A.; Brandi, A. Asymmetric and Doubly Asymmetric 1,3-Dipolar Cycloadditions in the Synthesis of Enantiopure Organophosphorus Compounds. Phosphorus Sulfur Silicon Relat. Elem. 1999, 144, 389–392. [Google Scholar] [CrossRef]
  23. Märkl, G.; Potthast, R. l-Phenyl-phosphol. Tetrahedron Lett. 1968, 9, 1755–1758. [Google Scholar] [CrossRef]
  24. Quin, L.D.; Mesch, K.A.; Bodalski, R.; Pietrusiewicz, K.M. 13C NMR spectra of phosphole–P(IV) dimers; ABX and AA′X 13C–31P coupling in some derived structures. Org. Magn. Reson. 1982, 20, 83–91. [Google Scholar] [CrossRef]
  25. Keglevich, G.; Kovács, J.; Körtvélyesi, T.; Parlagh, G.; Imre, T.; Ludányi, K.; Hegedűs, L.; Hanusz, M.; Simon, K.; Márton, A.; et al. Novel 2-phosphabicyclo [2.2.2]oct-5-ene derivatives and their use in phosphinylations. Heteroat. Chem. 2004, 15, 97–106. [Google Scholar] [CrossRef]
  26. Allen, F.H.; Baalham, C.A.; Lommerse, J.P.M.; Raithby, P.R. Carbonyl–Carbonyl Interactions can be Competitive with Hydrogen Bonds. Acta Cryst. 1998, B54, 320–329. [Google Scholar] [CrossRef]
  27. Breugst, M.; Reissig, H.-U. The Huisgen Reaction: Milestones of the 1,3-Dipolar Cycloaddition. Angew. Chem. Int. Ed. 2020, 59, 12293–12307. [Google Scholar] [CrossRef] [PubMed]
  28. 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]
  29. Houk, K.N.; Liu, F.; Yang, Z.; Seeman, J.I. Evolution of the Diels-Alder Reaction Mechanism since the 1930s: Woodward, Houk with Woodward, and the Influence of Computational Chemistry on Understanding Cycloadditions. Angew. Chem. Int. Ed. 2021, 60, 12660–12681. [Google Scholar] [CrossRef]
  30. 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]
  31. Sustmann, R.; Tappanchai, S.; Bandmann, H. (E)-1-Methoxy-1,3-butadiene and 1,1-Dimethoxy-1,3-butadiene in (4 + 2) Cycloadditions. A Mechanistic Comparison. J. Am. Chem. Soc. 1996, 118, 12555–12561. [Google Scholar] [CrossRef]
  32. CrysAlisPRO Software System; Oxford Diffraction/Agilent Technologies UK Ltd.: Yarnton, UK, 2015.
  33. Dolomanov, O.V.; Bourhis, L.J.; Gildea, R.J.; Howard, J.A.K.; Puschmann, H. OLEX2: A Complete Structure Solution, Refinement and Analysis Program. J. Appl. Crystallogr. 2009, 42, 339–341. [Google Scholar] [CrossRef]
  34. Sheldrick, G.M. SHELXT–Integrated space-group and crystal-structure determination. Acta Cryst. Sect. A Found. Adv. 2015, 71, 3–8. [Google Scholar] [CrossRef] [PubMed]
  35. Sheldrick, G.M. Crystal structure refinement with SHELXL. Acta Crystallogr. Sect. C Struct. Chem. 2015, 71, 3–8. [Google Scholar] [CrossRef]
  36. Spek, A.L. Structure validation in chemical crystallography. Acta Crystallogr. Sect. D Biol. Crystallogr. 2009, 65, 148–155. [Google Scholar] [CrossRef] [PubMed]
  37. Groom, C.R.; Bruno, I.J.; Lightfoot, M.P.; Ward, S.C. The Cambridge Structural Database. Acta Cryst. 2016, B72, 171–179. [Google Scholar] [CrossRef] [PubMed]
  38. Emerson, W.S.; Patrick, R.T.M. The preparation of 2-thiophenealdehyde and some of its derivatives. J. Org. Chem. 1949, 14, 790–797. [Google Scholar] [CrossRef]
  39. Hayat, F.; Salahuddin, A.; Umar, S.; Azam, A. Synthesis, characterization, antiamoebic activity and cytotoxicity of novel series of pyrazoline derivatives bearing quinoline tail. Eur. J. Med. Chem. 2010, 45, 4669–4675. [Google Scholar] [CrossRef]
  40. Urbaniak, K.; Mlostoń, G.; Gulea, M.; Masson, S.; Linden, A.; Heimgartner, H. Thio- and Dithioesters as Dipolarophiles in Reactions with Thiocarbonyl Ylides. Eur. J. Org. Chem. 2005, 1604–1612. [Google Scholar] [CrossRef]
  41. Mlostoń, G.; Urbaniak, K.; Gulea, M.; Masson, S.; Linden, A.; Heimgartner, H. [2+3]-Cycloadditions of Phosphonodithioformate S-Methanides with C=S, N=N, and C=C Dipolarophiles. Helv. Chim. Acta 2005, 88, 2582–2592. [Google Scholar] [CrossRef]
  42. Munasinghe, D.S.; Kasper, M.-A.; Jasiński, R.; Kula, K.; Palusiak, M.; Celeda, M.; Mlostoń, G.; Hackenberger, C.P.R. (3+2)-Cyclization Reactions of Unsaturated Phosphonites with Aldehydes and Thioketones. Chem. Eur. J. 2023, 29, e202300806. [Google Scholar] [CrossRef]
  43. McReynolds, M.D.; Dougherty, J.M.; Hanson, P.R. Synthesis of Phosphorus and Sulfur Heterocycles via Ring-Closing Olefin Metathesis. Chem. Rev. 2004, 104, 2239–2258. [Google Scholar] [CrossRef]
  44. Jun, J.H.; Javed, S.; Ndi, C.N.; Hanson, P.R. Synthesis of P-, S-, Si-, B-, and Se-Heterocycles via Ring-Closing Metathesis. In Synthesis of Heterocycles by Metathesis Reactions. Topics in Heterocyclic Chemistry; Springer: Cham, Switzerland, 2017; Volume 47, pp. 319–379. [Google Scholar] [CrossRef]
  45. Laxmikeshav, K.; Kumari, P.; Shankaraiah, N. Expedition of sulfur-containing heterocyclic derivatives as cytotoxic agents in medicinal chemistry: A decade update. Med. Res. Rev. 2022, 42, 513–575. [Google Scholar] [CrossRef] [PubMed]
  46. Xie, C.; Smaligo, A.J.; Song, X.-R.; Kwon, O. Phosphorus-Based Catalysis. ACS Cent. Sci. 2021, 7, 536–558. [Google Scholar] [CrossRef] [PubMed]
  47. Bouchet, D.; Van Elslande, E.; Masson, G. Isothiourea-Catalyzed Enantioselective Synthesis of trans-3,4-Dihydrothiopyranones: Harnessing Thiochalcones as Original Michael Acceptors. Org. Lett. 2023, 25, 5395–5399. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Aromatic thioketones 1a and 1b and the equilibrium of the monomeric (2) and dimeric (3/3′) structures of thiochalcones in solutions [6,9].
Figure 1. Aromatic thioketones 1a and 1b and the equilibrium of the monomeric (2) and dimeric (3/3′) structures of thiochalcones in solutions [6,9].
Molecules 29 02036 g001
Scheme 1. Recently described hetero-Diels–Alder reactions of thiochalcone 2 with naphthoquinone, leading to polycyclic S-heterocycle 4 [15].
Scheme 1. Recently described hetero-Diels–Alder reactions of thiochalcone 2 with naphthoquinone, leading to polycyclic S-heterocycle 4 [15].
Molecules 29 02036 sch001
Figure 2. The ‘quinone like’ P-heterocycles 5a and 5b, the five-membered phosphole 1-oxide derivatives 6a6c, and dihydrophosphinine 1-oxide 7 as potential or already studied dienophiles to be used in (4+2)- and (3+2)-cycloaddition reactions.
Figure 2. The ‘quinone like’ P-heterocycles 5a and 5b, the five-membered phosphole 1-oxide derivatives 6a6c, and dihydrophosphinine 1-oxide 7 as potential or already studied dienophiles to be used in (4+2)- and (3+2)-cycloaddition reactions.
Molecules 29 02036 g002
Scheme 2. Reported examples of stereoselective Diels–Alder reactions with phospholene 1-oxides 6a and 6b, leading to 15-P-steroid 8 (ref. [19]) and the fused P-heterocycle 9 (ref. [20]), respectively. The corresponding diene framework in cycloadducts 8 and 9 are presented in red.
Scheme 2. Reported examples of stereoselective Diels–Alder reactions with phospholene 1-oxides 6a and 6b, leading to 15-P-steroid 8 (ref. [19]) and the fused P-heterocycle 9 (ref. [20]), respectively. The corresponding diene framework in cycloadducts 8 and 9 are presented in red.
Molecules 29 02036 sch002
Scheme 3. Reported example of a stereoselective (3+2)-cycloaddition of phospholene 1-oxide 6b with a nitrone acting as a 1,3-dipole (ref. [21,22]). The nitrone framework in cycloadduct 10 is presented in red.
Scheme 3. Reported example of a stereoselective (3+2)-cycloaddition of phospholene 1-oxide 6b with a nitrone acting as a 1,3-dipole (ref. [21,22]). The nitrone framework in cycloadduct 10 is presented in red.
Molecules 29 02036 sch003
Scheme 4. Stereoselective dimerization of phospholene 1-oxide 6c and phosphinine 1-oxide 7 leading to polycyclic P-heterocycles 11 and 12, respectively (ref. [23,24] and [25]). The diene molecule 7′ and diene framework in cycloadduct 12 are presented in red.
Scheme 4. Stereoselective dimerization of phospholene 1-oxide 6c and phosphinine 1-oxide 7 leading to polycyclic P-heterocycles 11 and 12, respectively (ref. [23,24] and [25]). The diene molecule 7′ and diene framework in cycloadduct 12 are presented in red.
Molecules 29 02036 sch004
Scheme 5. Diastereoselective hetero-Diels–Alder reactions of 5a with thiochalcones 2a2d leading to fused P,S-heterocycles rac-(4RS,8SR,9SR,10SR)-13a13d as major products and rac-(4SR,8SR,9SR,10SR)-14a14d (tentative assignment of configuration) formed in low yields.
Scheme 5. Diastereoselective hetero-Diels–Alder reactions of 5a with thiochalcones 2a2d leading to fused P,S-heterocycles rac-(4RS,8SR,9SR,10SR)-13a13d as major products and rac-(4SR,8SR,9SR,10SR)-14a14d (tentative assignment of configuration) formed in low yields.
Molecules 29 02036 sch005
Figure 3. Molecular structure of the 4-(thien-2-yl)-substituted cycloadduct 13a. For graphics with atoms represented by thermal ellipsoids (50%), see Figure S14.
Figure 3. Molecular structure of the 4-(thien-2-yl)-substituted cycloadduct 13a. For graphics with atoms represented by thermal ellipsoids (50%), see Figure S14.
Molecules 29 02036 g003
Figure 4. Molecular structure of the 4-phenyl-substituted cycloadduct 13b. For graphics with atoms represented by thermal ellipsoids (50%), see Figure S15.
Figure 4. Molecular structure of the 4-phenyl-substituted cycloadduct 13b. For graphics with atoms represented by thermal ellipsoids (50%), see Figure S15.
Molecules 29 02036 g004
Scheme 6. Proposed structure of the transition states of the (4+2)-cycloaddition of thiochalcone 2 with 1-phenyl-4H-phosphinin-4-one 1-oxide (5a).
Scheme 6. Proposed structure of the transition states of the (4+2)-cycloaddition of thiochalcone 2 with 1-phenyl-4H-phosphinin-4-one 1-oxide (5a).
Molecules 29 02036 sch006
Figure 5. Structure of compound 13a.
Figure 5. Structure of compound 13a.
Molecules 29 02036 g005
Figure 6. Structure of compound 13b.
Figure 6. Structure of compound 13b.
Molecules 29 02036 g006
Figure 7. Structure of compound 13c.
Figure 7. Structure of compound 13c.
Molecules 29 02036 g007
Figure 8. Structure of compound 13d.
Figure 8. Structure of compound 13d.
Molecules 29 02036 g008
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Mlostoń, G.; Urbaniak, K.; Palusiak, M.; Łastawiecka, E.; Frynas, S.; Pietrusiewicz, K.M.; Heimgartner, H. Novel Bicyclic P,S-Heterocycles via Stereoselective hetero-Diels–Alder Reactions of Thiochalcones with 1-Phenyl-4H-phosphinin-4-one 1-Oxide. Molecules 2024, 29, 2036. https://doi.org/10.3390/molecules29092036

AMA Style

Mlostoń G, Urbaniak K, Palusiak M, Łastawiecka E, Frynas S, Pietrusiewicz KM, Heimgartner H. Novel Bicyclic P,S-Heterocycles via Stereoselective hetero-Diels–Alder Reactions of Thiochalcones with 1-Phenyl-4H-phosphinin-4-one 1-Oxide. Molecules. 2024; 29(9):2036. https://doi.org/10.3390/molecules29092036

Chicago/Turabian Style

Mlostoń, Grzegorz, Katarzyna Urbaniak, Marcin Palusiak, Elżbieta Łastawiecka, Sławomir Frynas, Kazimierz Michał Pietrusiewicz, and Heinz Heimgartner. 2024. "Novel Bicyclic P,S-Heterocycles via Stereoselective hetero-Diels–Alder Reactions of Thiochalcones with 1-Phenyl-4H-phosphinin-4-one 1-Oxide" Molecules 29, no. 9: 2036. https://doi.org/10.3390/molecules29092036

Article Metrics

Back to TopTop