Next Article in Journal
Continuous Flow-Mode Synthesis of Aromatic Amines in a 3D-Printed Fixed Bed Reactor Loaded with Amino Sugar-Stabilized Re Apparent Nanoparticles
Next Article in Special Issue
A Novel Method of Coupling In Situ Time-Resolved FTIR and Microwave Irradiation: Application to the Monitoring of Quinoxaline Derivatives Synthesis
Previous Article in Journal
Electrochemical Coagulant Generation via Aluminum-Based Electrocoagulation for Sustainable Greywater Treatment and Reuse: Optimization Through Response Surface Methodology and Kinetic Modelling
Previous Article in Special Issue
Recent Advances in Visible Light-Induced C-H Functionalization of Imidazo[1,2-a]pyridines
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

A Remarkable Selectivity Observed in Hetero-Diels–Alder Reactions of Levoglucosenone (LGO) with Thiochalcones: An Experimental and Computational Study †

by
Grzegorz Mlostoń
1,*,
Katarzyna Urbaniak
1,
Marcin Palusiak
2,
Ernst-Ulrich Würthwein
3,*,
Hans-Ulrich Reissig
4 and
Zbigniew J. Witczak
5
1
Department of Organic and Applied Chemistry, Faculty of Chemistry, University of Lodz, Tamka 12, PL-91-403 Lodz, Poland
2
Department of Physical Chemistry, Faculty of Chemistry, University of Lodz, Pomorska 163/165, PL-90-236 Lodz, Poland
3
Organisch-Chemisches Institut and Center for Multiscale Theory and Computation (CMTC), Universität Münster, Corrensstrasse 40, D-48149 Münster, Germany
4
Institut für Chemie und Biochemie, Freie Universität Berlin, Takustrasse 3, D-14195 Berlin, Germany
5
Department of Pharmaceutical Sciences, Nesbitt School of Pharmacy, Wilkes University, 84 W. South Street, Wilkes-Barre, PA 18766, USA
*
Authors to whom correspondence should be addressed.
Dedicated to the memory of Professor Julian Chojnowski (Łódź) (1935–2025).
Molecules 2025, 30(18), 3783; https://doi.org/10.3390/molecules30183783
Submission received: 8 August 2025 / Revised: 10 September 2025 / Accepted: 13 September 2025 / Published: 17 September 2025
(This article belongs to the Special Issue Heterocyclic Compounds: Synthesis, Application and Theoretical Study)

Abstract

Levoglucosenone (LGO) smoothly undergoes microwave-assisted hetero-Diels–Alder reactions with thiochalcones in THF solution at 60 °C. The studied reactions are completed after 10 min, and the expected tricyclic 2,3-dihydro-4H-thiopyran derivatives are formed in a highly regio- and moderately stereoselective manner via competitive exo- and endo-attacks of the 1-thiadiene moiety onto the activated C=C bond of dienophile LGO. Although eight isomers are possible, only the formation of exo,exo- (major) and exo,endo- (minor) cycloadducts was observed. In most cases, isomeric products were separated by preparative layer chromatography and identified by means of spectroscopic methods. Some of the cycloadducts were obtained as single crystalline solids, and X-ray analyses enabled unambiguous confirmation of their structures. In order to explain the observed selectivity of the studied hetero-Diels–Alder reactions, DFT studies were carried out to determine the thermodynamic and kinetic properties of all regio- and stereoisomers. The results of these calculations predict the preferred formation of the two experimentally observed isomers. In addition, remarkable details on the electronic structure of E-1,3-diphenylprop-2-en-1-thione and on involved and hypothetical transition states could be elucidated.

1. Introduction

(−)-Levoglucosenone (LGO) (1) (systematic names: 1,6-anhydro-3,4-dideoxy-β-D-glycero-hex-3-enopyranos-2-ulose or (1S,5R)-6,8-dioxabicyclo [3.2.1]oct-2-en-4-one) [1,2] is considered a biomass-derived, chiral platform molecule that is accessible by a Brønsted acid-assisted pyrolysis of cellulose. Its first isolation and structure determination was published in 1979 by a Japanese group [3] (Figure 1), and the ‘Organic Syntheses’ procedure was recently elaborated by B. A. Greatrex et al., which allows the preparation of pure 1 (95% purity) on a laboratory scale in ca. 9% yield [4]. Due to the rapidly growing demand for large-scale applications, industrial methods for the fabrication of 1, including biotechnological processes, have also been developed, and they are summarized in a recent review [5].
In recent years, a growing interest in new applications of 1 as a versatile reagent can be observed, and the compound has become an important enantiopure synthetic building block for the preparation of biologically active compounds available via multi-step procedures [6]. In addition, methods for the production of biodegradable polymers starting from levoglucosenone are currently under rapid development [7,8,9].
As shown in Scheme 1, efficient modifications of the LGO (1) skeleton can be achieved via various transformations of the reactive C=O and/or C=C bonds. In the current study, the α,β-unsaturated ketone moiety of 1 plays an important role in cycloadditions. A series of (3+2) cycloadditions (1,3-dipolar cycloadditions, Huisgen reactions) was successfully performed with 1 using nitrones [10,11], in situ-generated nitrile oxides [10], or nitrile imines [10,12] as 1,3-dipoles. Whereas reactions with a nitrile oxide and a C(Ph), N(Ph)-nitrile imine occurred with poor selectivity yielding mixtures of regio- and stereoisomeric isoxazole and pyrazole derivatives [10], (3+2) cycloadditions with fluorinated nitrile imines were highly stereoselective and exo-cycloadducts were formed as exclusive products without oxidation (and concomitant aromatization) of the initially formed 4,5-dihydropyrazole derivatives [12] (Scheme 1, [a] and [b]). The 1,3-dipolar cycloaddition of 1 with ethyl diazoacetate occurred with the formation of a single but unstable cycloadduct, which, under the reaction conditions, underwent further conversions initiated by dinitrogen elimination [13]. Notably, in a recent publication, a ‘higher order cycloaddition’ [(8+2) cycloaddition] of 1 with tropothione was also reported, and this reaction led to an exo-cycloadduct as a single product with complete stereoselectivity [14] (Scheme 1, [c]).
The first study on the (4+2) cycloadditions (Diels–Alder reactions) of LGO (1) with cyclopentadiene and buta-1,3-diene was published in 1981 [15]. In this report, the reaction with cyclopentadiene was claimed to occur with complete exo,exo-stereoselectivity with respect to both components (Scheme 1, [d] and [e]). Interestingly, the less reactive furan did not undergo the (4+2) cycloaddition with 1, but in the presence of the Lewis acid AlCl3, a Michael adduct was isolated, with 1 acting as a Michael acceptor [15]. It is worth noting that in a subsequent publication, the reaction of 1 with cyclopentadiene was reported to deliver an approximate 80:20 mixture (according to isolated amounts of isomeric products) of endo- and exo-cycloadducts, and this result is also presented in Scheme 1 (equation [d]) [16]. In a recent publication, diverse compounds derived from 1 were obtained via manipulation of the initially obtained polycyclic Diels–Alder cycloadducts [17]. The (4+2) cycloadditions of LGO (1) and of 4-cyano-levoglucosenone with four differently substituted dienes—e.g., cyclopentadiene, 2,3-dimethylbuta-1,3-diene, etc.—were carefully studied by means of experimental and computational methods, and the influence of diene substituents on the regioselectivity of the observed reactions was discussed [18]. Moreover, asymmetric Diels–Alder reactions starting with 1 or its derivatives, applied as chiral auxiliaries, have also been described [19,20].
Notably, in spite of the general importance of hetero-Diels–Alder reactions for the synthesis of various six-membered heterocycles [21,22,23], conversions of this type, starting with 1 as an active C=C dienophile, have not yet been reported. On the other side, in our continuing studies on cycloaddition reactions, sulfur-containing functionalized dipolarophiles play an important role. We employed aromatic and cycloaliphatic thioketones [24] as well as 1-thiadienes, such as thiochalcones 2 (α,β-unsaturated thioketones) [25,26,27], as reactive components. In the previous studies we found that thiochalcones efficiently capture acetylenic C≡C dienophiles to give functionalized 4H-thiopyrans in high yields (Scheme 2, left side) [28,29].
Encouraged by the well-documented importance of thiopyrans as attractive bio-isosters applicable in the design of some biologically active compounds and as practically useful building blocks for the preparation of more complex sulfur heterocycles [30], we decided to examine the usefulness of LGO (1) as a dienophile in hetero-Diels–Alder reactions with thiochalcones 2. The main target of the present study was to prove whether novel, polycyclic, and enantiopure tricyclic thiopyrans can be formed in these reactions in the absence of any catalyst. In addition, the scope and limitations of this synthetic approach, as well as the configuration of the anticipated (4+2) cycloadducts 3, should be established. Based on the results of already reported Diels–Alder reactions, formation of four diastereoisomers (exo,exo)-3, (exo,endo)-3, (endo,exo)-3, and (endo,endo)-3 can be anticipated [15,16], whereas the formation of the regioisomeric cycloadducts 4 is less likely. In order to understand the experimental results and to support the suggested reaction mechanism of this thus far unexplored reaction, DFT calculations were performed. Importantly, discussions on mechanistic aspects of cycloaddition reactions (stepwise versus concerted pathways) are in the focus of attention of numerous research groups [31,32,33].

2. Results and Discussion

2.1. Experimental Work

The test experiment was performed in THF solution without any catalyst, starting with equimolar amounts of LGO (1) and thiochalcone 2a (E-1,3-diphenylprop-2-en-1-thione). In the first experiment, the blue colored solution was heated in a thick-walled test tube with screw cap placed in an oil bath heated to 90 °C. After 1 h, the reaction was finished, but the brownish color of the obtained reaction solution indicated the formation of substantial amounts of tarry side products. Therefore, the reaction conditions were modified, and the second experiment was carried out in THF solution at 60 °C with the support of microwave (MW) irradiation. Gratifyingly, in this case, the reaction was complete already after 10 min, and the crude reaction mixture was almost colorless, indicating no formation of the undesired decomposition products. Based on this observation, the second protocol was applied as a standard procedure to study reactions of 1 with thiochalcones 2 (see Experimental). The 1H-NMR spectrum of the crude product mixture revealed two sets of signals, which could eventually be attributed to isomers of the expected cycloadduct derived from 1. Two doublets found at 6.26 (3JH,H = 6.20 Hz) and 6.61 (3JH,H = 3.90 Hz) ppm were of particular diagnostic value as they were located in the characteristic region known for HC(3) atoms of the thiopyran skeleton (for numbering of atoms, see Scheme 3). The ratio of intensities of these signals was established to be ca. 2:1, and it thereby determined the ratio of the anticipated isomeric (4+2) cycloadducts 3a formed in this reaction. Chromatographic separation (silica gel plate chromatographic separation) led to the isolation of two fractions, and the less polar fraction gave the major component found in the crude mixture in 40% yield. After crystallization, this fraction afforded colorless crystals with m.p. 148–149 °C. On the other hand, the more polar fraction separated chromatographically was identified as the minor component of the crude mixture (17% yield based on the isolated material). After crystallization, this cycloadduct formed colorless crystals with m.p. 186 °C (decomp.). The 1H-NMR spectra registered for purified products confirmed the structures of two isomers of the anticipated (4+2) cycloadducts. In the case of the major product, the most characteristic signals, found at 6.26 (d), 4.70 (m), and 4.37 (dd) ppm were attributed to H-atoms located at C(3), C(5), and C(4), respectively. For the minor product, analogous absorptions for the atoms HC(3), HC(5), and H(C4) were found at 6.61 (d), 4.75 (dd), and 4.33 (dd), respectively. In the 13C-NMR spectra of both products, the same number of signals (17 signals) confirmed the anticipated structure of the isomeric (4+2) cycloadducts, and the low-field shifted signals of the C=O group were found at 197.9 ppm for the major product and at 197.7 ppm for the minor component (Scheme 3).
Elemental analysis confirmed for both compounds the molecular formula C21H18O3S related to the structure of the isomeric (4+2) cycloadducts 3a.
Similar results were obtained using thiochalcones 2bf as highly active 1-thiadienes. The ratios of the isomeric (exo,exo)- and (exo,endo)-cycloadducts 3be were similar and in analogy to 3a. The stereochemical structure is based on the registered 1H-NMR spectra of the crude product mixtures, giving ratios of approximately 2:1. An exceptional case was observed for ferrocenyl functionalized thiochalcone 3f, which formed a 5:1 mixture of isomeric (exo,exo)- and (exo,endo)-cycloadducts. As expected, the ferrocenyl moiety exhibits a higher steric repulsion between LGO (1) and the sandwich-type ferrocenyl group, and therefore, the transition state leading to the (exo,endo)-3f is strongly disfavored (Scheme 3; Ar2 = Fc). In analogy to 3a, isomeric (exo,exo)- and (exo,endo)-cycloadducts 3b and 3d could successfully be separated by silica gel plate chromatographic separation (Scheme 3).
In the case of cycloadducts 3b, bearing a 4-BrC6H4 substituent at the C(4) atom, the separated isomers provided after crystallization from petroleum ether/CH2Cl2 mixture single crystals suitable for X-ray measurements which unambiguously confirmed the anticipated (exo,exo)-3b and (exo,endo)-3b structures for the less polar fraction (major isomer) and for the more polar fraction (minor isomer) (Figure 2).
Both isomeric cycloadducts 3b crystallized in the same P 212,121 space group with a single molecule in the unit cell. The enantiomorphic space group confirms the presence of only one enantiomeric form in each X-ray diffraction-identified crystal phase. The absolute configuration of both compounds 3b (exo,exo and exo,endo) is also confirmed by Flack parameter values collected in Table S1 (Supplementary Materials). The molecular geometry of both isomers is essentially comparable, with all bond lengths and angles falling within standard expected values. The key structural difference arises from the distinct arrangement of the 4-BrC6H4 substituent at the carbon atom C(4) in the heterocyclic ring (Figure 2).
Remarkably, attempted chromatographic separation of isomeric cycloadducts 3c with 4-ClC6H4 substituents at the carbon atom C(4) atom led to epimerization of the stereogenic center and the minor cycloadduct (exo,endo)-3c, which could initially be observed in the crude mixture of products, underwent complete conversion into the major one (i.e., (exo,exo)-3c). This effect can be rationalized by the electron-withdrawing effect of the 4-ClC6H4 substituent, which enhances the acidity of the HC(4) atom and thereby induces the observed epimerization by a deprotonation/protonation process (Scheme 4).
Notably, in the case of hetaryl-substituted cycloadducts 3e and 3f, all attempts to isolate pure diastereomers either by crystallization or by chromatography were unsuccessful. It is also worth stressing that no oxidation of compounds 3 was observed, which could lead to the corresponding 4H-thiopyran derivatives. In contrast, the structurally similar 2,3-dihydro-4H-thiopyrans, obtained by hetero-Diels–Alder reactions of quinones with thiochalcones, undergo this oxidation [34].
In order to learn more about the scope and limitations of the presented hetero-Diels–Alder reactions of levoglucosenone (1), we compared the reactivity of rather electron-rich thiochalcones 2 with that of electron-deficient heterodienes such as in situ-generated α-nitrosoalkene 5 and azoalkene 6 (Figure 3), which react efficiently with thioketones, yielding the expected (4+2)-cycloadducts [35,36,37]. However, in both experiments performed under typical conditions [35,36], no formation of the anticipated six-membered cycloadducts—i.e., 1,2-oxazine or pyridazine derivatives, respectively—could be observed.
Before discussing the stereochemical course of the studied hetero-Diels–Alder reactions, the regioselectivity of the cycloaddition should be regarded. The performed experiments demonstrated that only products derived from an attack of the sulfur terminus of thiochalcones 2 on the β-carbon of the α,β-unsaturated ketone 1 are isolated. This observation is in accordance with our earlier study on the reactions of thiochalcones 2 with α-nitrosoalkenes; only cycloadducts resulting from an attack of the sulfur terminus of 2 to the more electrophilic β-carbon of the heterodiene were identified [25]. The formation of mixtures of the isomeric cycloadducts (exo,exo)-3 and (exo,endo)-3 is likely the diastereoselectivity of the cycloaddition, which is influenced by two factors controlling the corresponding transition states in the concerted (4+2) cycloadditions of starting dienophile 1 and heterodiene 2 (Scheme 5).
The obtained experimental results demonstrate that the exclusive reaction pathway via the favored attack of the less hindered α-face (exo-face) of 1 determines the structure of the finally formed (4+2) cycloadducts. The major products (exo,exo)-3 result from exo-attacks with respect to thiochalcones 2, and the minor products (exo,endo)-3 from endo-approaches, respectively.
An alternative explanation of the observed formation of mixtures of isomeric cycloadducts 3 could be based on the assumption of a stepwise mechanism of the studied hetero-Diels–Alder reactions, either via intermediate zwitterionic or diradical species. In order to examine this hypothesis, an experiment was performed starting with 1 and thiochalcone 2a (THF solution, MW irradiation) in the presence of 3 mol equivalents of methanol used as a trapping reagent (for a similar experiment reported by R. Sustmann et al., see e.g., [38]). The obtained mixture of crude products was examined by 1H-NMR spectroscopy, which demonstrated that its composition was identical to that discussed above, performed with 1 and 2a in the absence of methanol. This result can be considered as strong evidence of a concerted and not a stepwise pathway of the hetero-Diels–Alder reactions of 1 with thiochalcones 2. This assumption is supported by the comprehensive DFT calculations (vide infra).

2.2. Mechanistic Investigations by DFT Calculations

The mechanism as well as the regioselectivity and diastereoselectivity of the hetero-Diels–Alder reactions were also analyzed by a systematic quantum chemical DFT study of the eight possible isomers of 3a and 4a (see Scheme 2). In the following, the experimental results will be interpreted based on the calculated Gibbs free energy surfaces (ΔG298 [kcal/mol]), which describe the thermodynamic and kinetic properties.
On the basis of DFT optimizations using the B3LYP/6-31G(d) [39,40] + GD3BJ [41,42] functional, geometry optimizations using PBE1PBE/def2tzvp [43,44,45,46] + GD3BJ [42], including the PCM solvent sphere of tetrahydrofuran [47], were performed. Subsequent frequency calculations (zero-point vibrational energies and free enthalpy contributions) gave the final free Gibbs energies (DG298 [kcal/mol]), which were used for the interpretation of the experimental results. For all calculations, the GAUSSIAN 16 package of programs was used (see Supplementary Information) [48]. In accordance with the reaction conditions (60 °C, microwave irradiation, no light, no catalysts), only closed-shell calculations were performed. Some of the calculated minima were checked by the “stable” option of Gaussian; the wave functions were found to be stable under the perturbations considered.
As shown in Scheme 6 and Scheme 7, thiochalcone 2a (E-1,3-diphenylprop-2-en-1-thione) was employed as the standard hetero-diene for the calculations of the (4+2) cycloadditions with LGO (1) as used in the experiments. For the calculations of the reactions leading to 3a and its hypothetical isomers, the configurations and conformations obtained from the X-ray structures of (exo,exo)-3b (Figure 2 left) and (exo,endo)-3b (Figure 2 right) were taken as starting structures for the complete optimizations. The transition states were localized by stepwise elongation of the respective C–C and C–S bonds (reaction path calculations), followed by transition state geometry optimization including frequency calculations.
Whereas the calculated NBO charges [49,50] of LGO 1 show the expected charge distribution of enones (O: −0.541 e; C(O): 0.488 e; C-α: −0.316 e; C-β: −0.113 e), the charge distribution of thiochalcone 2a differs significantly due to the lower electronegativity of sulfur: S: −0.082 e, C(S): −0.094 e; C-α: −0.286 e; C-β: −0.075 e. Thus, the electrophilic and nucleophilic properties of the enone and of the thiochalcone are expected to be quite different, allowing unusual reactions as experimentally observed. Thus, the almost neutral sulfur atom of 2a is able to attack at the quite nucleophilic α-atom as well as at the neighboring β-atom of the enone LGO (1). Scheme 6 and Scheme 7 show the eight possible pathways to cycloadducts 3a and 4a and the corresponding kinetic and thermodynamic data.
Scheme 6 presents the results of attacks of the almost neutral S terminus of 2a to the β-carbon atom of enone 1 via the α-face (exo) and β-face (endo) (see Figure 1; compare refs. [28,29]). On the right-hand part of Scheme 6, the attacks to the exo-side of LGO (1) are given (O-bridge side, “from below”), on the left-hand part, the attacks to the endo-side (CH2-O-bridge side, “from above”) take place. In the upper line, the thiochalcone attacks with the β-C phenyl group in an exo-orientation, leading to (exo,exo)-3a or to (endo,exo)-3a. In the lower line of Scheme 6, this phenyl group is endo-positioned, giving (exo,endo)-3a and (endo,endo)-3a, respectively (see Scheme 5 for an alternative presentation).
As the calculations show, all four cycloadditions proceed exothermically (−19.2 to −8.7 kcal/mol). However, the calculated kinetic barriers (Gibbs free activation energies) are quite different, with TS-(exo,exo)-3a [14.8 kcal/mol] being the significantly lowest transition state calculated. Interestingly, this transition state leads to the experimentally observed major product (exo,exo)-3a (−19.2 kcal/mol), which is also the thermodynamically best isomer of all. Thus, this reaction is a good example of a kinetically controlled process leading to the experimentally observed major stereoisomer (exo,exo)-3a. The second best isomer, (exo,endo)-3a [−15.1 kcal/mol], is accessible over the second lowest barrier (15.7 kcal/mol); in fact, this compound was isolated as a minor product in the experiments (see above). The difference between (exo,exo)-3a and (exo,endo)-3a is clearly addressable to the sterically slightly more hindered position of the phenyl group at the β-C-position stemming from the thiochalcone. The hypothetical pathways leading to (endo,exo)-3a or (endo,endo)-3a have much higher Gibbs free activation energies (30.4 and 19.8 kcal/mol) due to the severe steric hindrance exhibited by the CH2-O-bridge of LGO (1). Hence, in accord with the experimental approaches of 2a to this β-face of LGO (endo), they are unfavorable.
Similarly, as shown in Scheme 7, the calculated results from the attack of the S terminus of thiochalcone 2a to the α-C of the LGO-enone show a thermodynamically and kinetically preferred attack to the α face (exo) of (LGO) 1, giving (exo,exo)-4a. Interestingly, all four isomers of 4a are thermodynamically in a narrow range (−14.3 to −17.7 kcal/mol), showing that steric interactions are similar for all products; in comparison, higher stability differences were calculated for isomer 3a. However, the attack of the S terminus to the β-C of the enone to give (exo,exo)-3a, as described above, is kinetically preferred. For the hypothetical isomers 4a attack to the β-face from the endo-side is again significantly disfavored compared to other stereochemical pathways. Although the mass balances of the experiments are only moderate to good (see Scheme 3), there is no experimental evidence for the formation of one of the 4a-stereoisomers. From the kinetic point of view, no transition state shown in Scheme 7 is lower in energy compared to the data of Scheme 6. In general, the thermodynamically best isomers result from sulfur attacks towards the β-carbon atom of LGO (1) (see Scheme 6).
Both energy lowest transition states TS-(exo,exo)-3a and TS-(exo,endo)-3a (Figure 4a,b) show quite different atomic distances between the reacting C and S as well as C and C atoms [(a) TS-(exo,exo)-3a: C⋯S: 2.289 Å, C⋯C 2.608 Å, (b) TS-(exo,endo)-3a: C⋯S 2.270 Å, C⋯C 2.539 Å]. It is remarkable to see that the forming C⋯S bond is significantly shorter compared to the C⋯C bond, indicating a slightly asynchronous bond formation with moderate dominance of the C⋯S bond generation. Geometrical differences in the exo,exo- and exo,endo-transition states are only small, as illustrated in Figure 4 (TS-(exo,exo)-3a: C⋯S and C⋯C bonds are slightly larger than for TS-(exo,endo)-3a). As mentioned above, the low energetic difference of 0.9 kcal/mol between these two transition states results from the smaller steric interaction of the CH-phenyl group with the bicycle in TS-(exo,exo)-3a. The NBO-charges for TS-(exo,exo)-3a (LGO part: O: −0.586 e; C(O): 0.457 e; C-α: −0.350 e; C-β: −0.179 e; thiochalkone part: S: 0.093 e; C(S): −0.136 e; C-α: −0.275 e; C-β: −0.060 e) indicate a substantial charge transfer (0.279 e) from the thiochalcogene to the LGO part of the transition state. For both forming bonds, the difference in charges of the reacting atoms is quite similar [C⋯S: 0.093 e − (−0.179 e) = 0.272 e versus C⋯C: (−0.060 e) − (−0.350 e) = 0.290 e].
Out of curiosity, the respective transitions TS-(exo,exo)-4a and TS-(exo,endo)-4a are depicted in Figure 5. In contrast to the 3a-transition states they show shorter C⋯C distances compared to the S⋯C distances [(a) TS-(exo,exo)-4a, distance C⋯S: 2.424 Å, distance C⋯C: 2.259 Å; (b) TS-(exo,endo)-4a, distance C⋯S: 2.394 Å, distance C⋯C: 2.342 Å]. This can be interpreted as a preference for the bond formation to the β-carbon of enone 1, regardless of an attack by a carbon or by a sulfur center of 2a. The NBO charges of the relevant atoms for TS-(exo,exo)-4a (LGO part: O: −0.550 e; C(O): 0.497 e; C-α: −0.316 e; C-β: −0.183 e, thiochalkone part: S: 0.076 e; C(S): −0.169 e; C-α: −0.242 e; C-β: −0.120 e) differ significantly from those of TS-(exo,exo)-3a, being in sum by 0.106 e more positive. Remarkably, in the hypothetical TS-(exo,exo)-4a, the difference in charges of the reacting atoms is quite different (C⋯S: 0.391 e versus C⋯C: 0.062 e) for the two forming bonds.
In summary, the DFT calculations reveal kinetic control of the studied hetero-Diels–Alder reactions and, in accord with the experimental results, the preferred attack of the sulfur atom of thiochalcone 2a to the β-C of enone 1. Not only is the correct regiochemistry proposed, but the observed diastereoselectivity with preferred attacks to the α-face (exo) leading to the thermodynamically best isomer (exo,exo)-3a) and to the second best isomer (exo,endo)-3a. The attack of the S-atom of 2a to the α-C of LGO (1) is kinetically considerably disfavored, and in agreement with this result, isomers 4a are experimentally not observed.

3. Materials and Methods

3.1. Materials

Starting materials: levoglucosenone (LGO) (1) was synthesized following the recently published procedure [4]; thiochalcones 2ad were prepared by treatment of the corresponding chalcones with Lawesson’s reagent as a thionating agent and purified by flash column chromatography according to the published procedures [25,51].

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 reactants were added by using a syringe; subsequent manipulations were conducted in the air. NMR spectra were taken with Bruker AVIII instruments [1H-NMR (600 MHz); 13C-NMR (151 MHz); Bruker, Billerica, MA, USA]. Chemical shifts are given relative to solvent residual peaks, integrals are in accordance with assignments, and coupling constants J are given in Hz. Elemental analyses were obtained with a Vario EL III instrument. Microwave experiments were carried out with CEM-focused Microwave-type (Yokohama-shi, Japan) Discover SPD at 150 W. Optical rotations were measured using an Anton Paar (Graz, Austria) MCP 500 polarimeter at the temperatures indicated. Melting points were determined in capillaries with an Aldrich (Saint Louis, MO, USA) Melt-Temp II apparatus and are uncorrected.

3.3. Quantum Chemical Calculations

Quantum Chemical calculations [PBE1PBE/def2-TZVP+PCM (tetrahydrofuran)+GD3BJ dispersion correction] [41,42,43,44,45,46,47] were performed on the basis of preceding B3LYP/6-31G(d) [39,40,41,42] +GD3BJ-geometry optimizations using the Gaussian 16, Revision B.01 [48], package of programs. Several conformers of each isomer were calculated, often after MM2-conformational analysis, to obtain the lowest energy isomer. The transition state localizations are based on reaction path calculations by elongation of both relevant bonds, starting with the cycloadducts (“retro-hetero-Diels–Alder reaction”) and full optimization of all other parameters. Transition state searches on the basis of the calculated 3D surfaces followed. In several cases, IRC calculations were performed in order to characterize the transition states obtained.

3.4. Synthesis

Reactions of thiochalcones 2ad with (−)-levoglucosenone (LGO) (1)-general procedure:
Molecules 30 03783 i001
A solution of the corresponding thiochalcone 2 (1.1 mmol) and LGO (1) (126 mg, 1 mmol) in 4 mL of dry THF was irradiated in a microwave apparatus (200 W) at 60 °C for 10 min. Then, the reaction mixture was cooled down to room temperature, the solvent was evaporated, and the oily, brownish colored residue was preliminarily purified on a short chromatography column packed with silica gel (ca. 2–3 cm layer) using a CH2Cl2/petroleum ether (1:1) mixture as an eluent. The mixture of the crude products obtained thereafter was analyzed by 1H-NMR spectroscopy, which revealed the presence of two isomeric products. The repeated chromatography on a column (or on preparative plates coated with silica gel) allowed separation of isomeric products 3ad.
Attention: (1) Alternatively, a similar procedure was applied for experiments performed in THF solutions with reaction times estimated for 1h, and in this case, the reaction was carried out at 90 °C in a thick-walled, screwed test tube. Using this method, the formation of substantial amounts of decomposition products was observed, and the yields of isolated cycloadducts 3 were lower than in the case of the above-presented protocol with the MW support. (2) Syntheses of cycloadducts 3e and 3f: attempted separations of the two isomers failed, and the products were characterized as a mixture (their data are presented in the Supplementary Materials.
Molecules 30 03783 i002
(4R,4aS,6R,9R,9aS)-2,4-Diphenyl-4a,8,9,9a-tetrahydro-4H-6,9-epoxythiopyrano [2,3-d]oxepin-5(6H)-one: (exo,exo)-3a (major, isolated as less polar fraction). Yield: 140 mg (40%), colorless crystals, m.p. 148–149 °C (PE/CH2Cl2).
1H NMR (CDCl3): δ = 3.32 (dd, JH,H = 6.1 Hz, JH,H = 2.1 Hz, 1H); 3.83 (dd, JH,H = 6.1 Hz, JH,H = 2.0 Hz, 1H); 4.04–4.08 (m, 2H); 4.37 (dd, JH,H = 6.1 Hz, JH,H = 2.0 Hz, 1H); 4.70–4.71 (m, 1H); 5.32 (s, 1H); 6.26 (d, JH,H = 6.2 Hz, 1H); 7.30–7.32 (m, 3HCarom); 7.35–7.39 (m, 5HCarom); 7.56–7.59 (m, 2HCarom).
13C NMR (CDCl3): δ = 37.9, 44.6, 46.5 (3HC); 67.2 (H2C-O); 75.9, 101.6 (2HC); 118.2, 126.9, 127.2, 128.1, 128.4, 128.5, 129.0 (10HCarom, and HC=); 134.1, 138.9, 142.7 (2Carom, PhC=); 197.9 (C=O).
IR: ν 1741s (C=O), 1490m, 1448m, 1252m, 1110s, 1028m, 976s, 905s, 752s, 693s, cm−1.
EA for C21H18O3S (350.43): calcd. C 71.98, H 5.18, S 9.15; found C 71.78, H 5.15, S 9.37.
α = + 92.5 (CHCl3, c = 0.0020 g/mL).
Molecules 30 03783 i003
(4S,4aS,6R,9R,9aS)-2,4-Diphenyl-4a,8,9,9a-tetrahydro-4H-6,9-epoxythiopyrano [2,3-d]oxepin-5(6H)-one: (exo,endo)-3a (minor, more polar). Yield: 60 mg (17%), colorless crystals, m.p. 184–186 °C (diisopropyl ether/CH2Cl2).
1H NMR (CDCl3): δ = 3.62 (pseudo t, JH,H = 3.3 Hz, 1HC); 3.81 (dd, JH,H = 7.5 Hz, JH,H = 2.8 Hz, 1HC); 4.10 (dd, JH,H = 7.9 Hz, JH,H = 5.6 Hz, 1HC); 4.21 (d, JH,H = 7.9 Hz, 1HC); 4.33 (dd, JH,H = 7.5 Hz, JH,H = 1.6 Hz, 1HC); 4.75 (dd, JH,H = 5.5 Hz, JH,H = 1.5 Hz, 1HC); 5.12 (s, 1HC); 6.61 (d, JH,H = 3.9 Hz, 1HC); 7.30–7.40 (m, 6HCarom); 7.53–7.58 (m, 4HCarom).
13C NMR (CDCl3): δ = 43.1, 46.7, 50.7 (3HC); 67.6 (H2C); 76.7, 102.5 (2HC); 123.5, 126.7, 126.8, 128.3, 128.4, 128.5, 128.8 (10HCarom, and HC=); 134.7, 138.8, 141.0 (2Carom, PhC=); 197.7 (C=O).
IR: ν 1738s (C=O), 1491m, 1446m, 1300m, 1111s, 1029m, 972s, 916s, 749s, 693s, cm−1.
EA for C21H18O3S (350.43): calcd. C 71.98, H 5.18, S 9.15; found C 71.85, H 5.17, S 9.20.
α = + 31.1 (CHCl3, c = 0.0020 g/mL).
Molecules 30 03783 i004
(4R,4aS,6R,9R,9aS)-(4-Bromophenyl)-2-phenyl-4a,8,9,9a-tetrahydro-4H-6,9-epoxythiopyrano [2,3-d]oxepin-5(6H)-one: (exo,exo)-3a (major, less polar). Yield: 170 mg (40%), yellowish crystals, m.p. 168–169 °C (petroleum ether/CH2Cl2).
1H NMR: δ = 3.26 (dd, JH,H = 6.1 Hz, JH,H = 2.1 Hz, 1HC); 3.77 (dd, JH,H = 6.1 Hz, JH,H = 2.1 Hz, 1HC); 4.06 (d, JH,H = 2.8 Hz, 2HC); 4.32 (dd, JH,H = 6.1 Hz, JH,H = 2.1 Hz, 1HC); 4.69–4.71 (m, 1HC); 5.31 (s, 1HC); 6.20 (d, JH,H = 6.2 Hz, 1HC=); 7.16, 7.49 (AB-system, 3JH,H = 8.4 Hz, 4HC+arom); 7.32–7.38 (m, 3HCarom); 7.54–7.57 (m, 2HCarom).
13C NMR: δ = 37.5, 44.6, 46.4 (3HC); 66.9 (H2C-O); 75.7, 101.7 (2HC); 117.6, 126.5, 128.5, 129.8, 128.5 132.1 (9HCarom, and HC=); 121.1, 134.8, 138.7, 141.5 (3Carom, PhC=); 197.5 (C=O).
IR: ν 1744s (C=O); 1483m, 1435m, 1256m, 1103s, 1020m, 976s, 908s, 745s, 708s, 696s.
EA for C21H17BrO3S (429.33) calcd. C 58.75, H 3.99, S 7.47; found C 58.73, H 4.00, S 7.63.
α = +307.3 (CHCl3, c = 0.0018 g/mL).
Molecules 30 03783 i005
(4S,4aS,6R,9R,9aS)-2-Bromophenyl-4-phenyl-4a,8,9,9a-tetrahydro-4H-6,9-epoxythiopyrano [2,3-d]oxepin-5(6H)-one: (exo,endo)-3b (minor, more polar). Yield: 70 mg (16%), yellow crystals, m.p. 155–157 °C (petroleum ether/CH2Cl2).
1H NMR: δ = 3.57 (t, JH,H = 3.4 Hz, 1HC); 3.75 (dd, JH,H = 7.5 Hz, JH,H = 2.8 Hz, 1HC); 4.09 (dd, JH,H = 7.9 Hz, JH,H = 5.1 Hz, 1HC); 4.19 (d, JH,H = 7.9 Hz, 1HC); 4.31 (dd, JH,H = 7.5 Hz, JH,H = 1.6 Hz, 1HC); 4.74 (dd, JH,H = 4.8 Hz, JH,H = 1.2 Hz, 1HC); 5.11 (s, 1HC); 6.51 (d, 3JH,H = 3.9 Hz, 1HC=); 7.32–7.37 (m, 3HCarom); 7.41, 7.49 (AB-system, JH,H = 8.4 Hz, 4HCarom); 7.55–7.57 (2HCarom).
13C NMR: δ = 44.6, 46.6, 50.6 (3HC); 67.5 (H2C); 76.7, 102.4 (2HC); 122.6, 126.7, 127.7, 128.5, 130.6, 131.4 (9HCarom, HC=); 120.8, 135.2, 138.6, 141.0 (3Carom, and PhC=); 197.6 (C=O).
IR: ν 1759s (C=O); 1483m, 1453m, 1300m, 1207m, 1000s, 942s, 905s, 752s, 695s, cm−1.
EA for C21H17BrO3S (429.33) calcd C 58.75, H 3.99, S 7.47; found C 58.75, H 3.97, S 7.54.
α = –7.8 (CHCl3, c = 0.0018 g/mL).
Molecules 30 03783 i006
(4R,4aS,6R,9R,9aS)-2-(4-Chlorophenyl-4-phenyl-4a,8,9,9a-tetrahydro-4H-6,9-epoxythiopyrano [2,3-d]oxepin-5(6H)-one: (exo,exo)-3c (major, less polar). Yield: 145 mg (38%), yellowish crystals, m.p. 164–165 °C (petroleum ether/CH2Cl2).
1H NMR: δ = 3.26 (dd, JH,H = 6.1 Hz, JH,H = 2.3 Hz, 1HC); 3.76 (dd, JH,H = 6.1 Hz, JH,H = 1.9 Hz, 1HC); 4.06 (d, JH,H = 2.9 Hz, H2C); 4.34 (dd, JH,H = 6.2 Hz, JH,H = 2.2 Hz, 1HC); 4.69–4.71 (m, 1HC); 5.30 (s, 1HC); 6.21 (d, JH,H = 6.2 Hz, 1HC=); 7.22, 7.33 (AB-system, JH,H = 8.8 Hz, 4HCarom); 7.35–7.39 (m, 3HCarom); 7.55–7.58 (m, 2HCarom).
13C NMR: δ = 37.4, 44.6, 46.3 (3HC); 67.0 (CH2-O); 75.7, 101.7 (2HC); 117.7, 126.4, 128.5, 128.7, 129.8, 132.0 (for 9HCarom, and HC=); 121.2, 134.9, 138.7, 141.5 (3Carom, PhC=); 197.6 (C=O).
IR: ν 1744s (C=O); 1493m, 1448m, 1259m, 1107s, 1073m, 986s, 905s, 749s, 711s, 697s, cm−1.
EA for C21H17ClO3S (384.87) calcd C 65.54, H 4.45, S 8.33; found C 65.31, H 4.51, S 8.58.
α = + 117.0 (CHCl3, c = 0.0016 g/mL).
Molecules 30 03783 i007
(4S,4aS,6R,9R,9aS)-2-(4-Chlorophenyl-4-phenyl-4a,8,9,9a-tetrahydro-4H-6,9-epoxythiopyrano [2,3-d]oxepin-5(6H)-one: (exo,endo)-3b (minor)—not isolated; underwent isomerization in the course of chromatographic purification on the SiO2 column.
Selected signals taken from the 1H-NMR spectrum of the crude mixture of isomeric products:
1H NMR: δ = 6.51 (d, 1H, JH,H = 3.90 Hz), 5.11 (s, 1H), 4.73–4.75 (m, 1H).
Molecules 30 03783 i008
(4R,4aS,6R,9R,9aS)-4-Phenyl-2-(thiophen-2-yl)-4a,8,9,9a-tetrahydro-4H-6,9-epoxythiopyrano [2,3-d]oxepin-5(6H)-one: (exo,exo)-3d (major, less polar). Yield: 140 mg (39%), colorless crystals, m.p. 170 °C (diisopropyl ether/CH2Cl2).
1H NMR (CDCl3): δ = 3.33 (dd, JH,H = 5.9 Hz, JH,H = 2.2 Hz, 1HC); 3.83 (dd, JH,H = 6.0 Hz, JH,H = 2.0 Hz, 1HC); 4.04–4.09 (m, H2C); 4.34 (dd, JH,H = 6.0 Hz, JH,H = 2.2 Hz, 1HC); 4.71–4.72 (m, 1HC); 5.31 (s, 1HC); 6.34 (d, JH,H = 6.0 Hz, 1HC=); 7.02 (dd, JH,H = 5.0 Hz, JH,H = 3.7 Hz, 1HCarom); 7.24–7.31 (m, 5HCarom); 7.36–7.38 (m, 2HCarom).
13C NMR: (CDCl3): δ = 37.9, 45.1, 47.1 (3HC); 67.0 (H2C); 75.6, 101.7 (2HC); 117.9, 124.1, 124.7, 127.3, 127.4, 128.1, 129.0 (for 8HCarom, and HC=); 127.7, 142.0, 142.7 (2Carom, and PhC=); 197.7 (C=O).
IR: ν 1741s (C=O), 1599m, 1491m, 1449m, 1308m, 1256m, 1107s, 1017m, 976s, 916s, 898m, 771m, 697s, cm−1.
EA for C19H16O3S2 (356.45) calcd C 64.02, H 4.52, S 17.99; found C 63.89, H 4.60, S 17.89.
α = +72.4 (CHCl3, c = 0.0026 g/mL).
Molecules 30 03783 i009
(4R,4aS,6R,9R,9aS)-4-Phenyl-2-(thiophen-2-yl)-4a,8,9,9a-tetrahydro-4H-6,9-epoxythiopyrano [2,3-d]oxepin-5(6H)-one: (exo,endo)-3d (minor, more polar). Yield: 50 mg (14%), beige crystals, m.p. 145 °C (diisopropyl ether/CH2Cl2).
1H NMR (CDCl3): δ = 3.66 (pseudo t, JH,H = 3.4 Hz, 1HC); 3.80 (dd, JH,H = 7.4 Hz, JH,H = 2.8 Hz, 1HC); 4.09 (dd, JH,H = 7.9 Hz, JH,H = 5.0 Hz, 1HC); 4.20 (d, JH,H = 7.9 Hz, 1HC); 4.35 (dd, JH,H = 7.4 Hz, JH,H = 1.7 Hz, 1HC); 4.73 (dd, JH,H = 4.8 Hz, JH,H = 1.2 Hz, 1HC); 5.12 (s, 1HC); 6.67 (d, JH,H = 4.0 Hz, 1HC=); 7.00 (dd, JH,H = 5.1 Hz, JH,H = 3.6 Hz, 1HCarom); 7.21–7.27 (m, 3HCarom); 7.37–7.40 (m, 2HCarom); 7.51–7.52 (m, 2HCarom).
13C NMR (CDCl3): δ = 44.8, 46.7, 50.8 (3HC); 67.5 (H2C); 76.5, 102.4 (2HC); 122.2, 124.2, 124.8, 126.9, 127.3, 128.4, 128.9 (for 8HCarom, and HC=); 128.0, 140.6, 141.8 (2Carom, PhC=); 197.5 (C=O).
IR: ν 1737s (C=O), 1589m, 1494m, 1450m, 1301m, 1226m, 1110s, 1025m, 969s, 902s, 864m, 752m, 693s, cm−1.
EA for C19H16O3S2 (356.45) calcd C 64.02, H 4.52, S 17.99; found C 64.05, H 4.15, S 17.57.
α = +36.4 (CHCl3, c = 0.0022 g/mL).

4. Conclusions

The presented study demonstrates for the first time that (−)-levoglucosenone (LGO) (1) can be successfully applied as a dienophile in the hetero-Diels–Alder reactions starting with thiochalcones 2 as easily available 1-thia-1,4-dienes. Thiopyran derivatives, which are formed as (4+2) cycloadducts, are known as practically useful pharmacophores applied for the preparation of biologically active compounds [52,53], which can act, for example, as enzyme inhibitors [54], antiproliferative [55], anti-inflammatory [56] or antibacterial [57] agents are well documented. Notably, even greater interest concerns biologically active derivatives of structurally similar thiochromane and thiochromene-based sulfur heterocycles [58]. Therefore, the presented method for the modification of levoglucosenone skeleton by annulation of the six-membered sulfur heterocycles can offer an attractive perspective for the search for its new, biologically active derivatives (see in [5,6]). Moreover, the presented method for the synthesis of 2,3-dihydro-4H-thiopyrans supplements the earlier described hetero-Diels–Alder reaction, which was based on the usage of typical 1,3-dienes and some thioaldehydes bearing a carboxamide group as reactive hetero-dienophile. In this study, the obtained (4+2) cycloadducts were evaluated for in vitro and ex vivo ACAT enzyme inhibition [59].
The studied hetero-Diels–Alder reaction occurred regioselectively and stereoselectively. Although eight isomers are possible, it led to only two diastereoisomeric products described as (exo,exo)- and (exo,endo)-cycloadducts, which differ by the orientation of the aryl substituent located at the newly created stereogenic center at C(4) atom of the thiopyran ring. Unexpectedly, the minor cycloadduct (exo,endo)-3c bearing the 4-ClC6H4 substituent at the C(4) atom, underwent slow epimerization in the course of chromatographic purification on the silica gel column. Remarkably, in contrast to electron-rich thiochalcones, electron-deficient heterodienes such as α-nitrosoethylenes and azoethylenes remain unreactive towards LGO (1).
The DFT calculations performed for the reactions of 1 with 2a fully rationalize the experimentally observed selectivities and the preference for the formation of the sterically favored exo,exo-configured product 3a as the major (4+2) cycloadduct. They also revealed a slightly asynchronous bond-forming scenario with shorter C⋯S distances in TS-exo,exo-3a, and TS-exo,endo-3a compared to the newly generated C⋯C bonds. Remarkably, the DFT calculations for the experimentally not observed regioisomers 4a show the opposite effect. In these hypothetical reactions, the bond formations to the β-C atom of the enone system of LGO (1) are more advanced in the transition states than those to the α-C atom.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/molecules30183783/s1. The Supplementary Materials contain some experimental details, the scanned 1H NMR and 13C NMR spectra of the described compounds, and information related to the single crystal X-ray measurements. The X-ray crystallography data of compounds (exo,exo)-3b and (exo,endo)-3b are deposited as CSD Communications under deposition numbers 2471272 and 2471271, respectively. The details of the DFT calculations, including the Gaussian archive entries, are also documented. Refs. [4,25,28,41,42,43,44,45,46,47,48,60,61,62,63,64,65] are cited in the Supplementary Materials.

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 (experimental part), and Supplementary Materials preparation; M.P.—X-ray analysis, graphics preparation, and Supplementary Materials preparation; E.-U.W.—computational work, mechanistic discussion, graphics preparation, main manuscript, and Supplementary Materials preparation; H.-U.R.—reference collection, checking their contents, mechanistic discussion, and manuscript preparation; Z.J.W.—LGO preparation, manuscript preparation, and consultancy of the obtained results. All authors have read and agreed to the published version of the manuscript.

Funding

The experimental work was partially supported by the University of Łódź within the IDUB-Project (2023–2024).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Reported data are available from the authors via e-mail contact.

Acknowledgments

The authors thank Małgorzata Celeda (University of Lodz) for helpful assistance in the course of experimental work performed in the laboratory and Jakub Wręczycki (Lodz University of Technology, Lodz) for his help in preparation of the final version of this manuscript. G.M. thanks Heinz Heimgartner (Zurich) for a fruitful discussion in the course of preparation of this manuscript. We thank Christian Mück-Lichtenfeld (Universität Münster) for very helpful discussions concerning the theoretical part.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Witczak, Z.J. (Ed.) Levoglucosenone and Levoglucosans: Chemistry and Applications; ATL Press: Mount Prospect, IL, USA, 1994. [Google Scholar]
  2. Witczak, Z.J.; Tatsuta, K. (Eds.) Carbohydrate Synthons in Natural Products Chemistry. Synthesis, Functionalization, and Applications; ACS Symposium, Series No. 841; American Chemical Society: Washington, DC, USA, 2003. [Google Scholar]
  3. Tsuchiya, Y.; Sumi, K. Thermal decomposition products of cellulose. J. Appl. Polym. Sci. 1979, 14, 2003–2013. [Google Scholar] [CrossRef]
  4. Klepp, J.; Dillon, W.; Lin, Y.; Feng, P.; Greatrex, B.W. Preparation of (−)-Levoglucosenone from Cellulose Using Sulfuric Acid in Polyethylene Glycol. Org. Synth. 2020, 97, 38–53. [Google Scholar] [CrossRef]
  5. Allais, F. Total syntheses and production pathways of levoglucosenone, a highly valuable chiral chemical platform for the chemical industry. Curr. Opin. Green Sustain. Chem. 2023, 40, 100744. [Google Scholar] [CrossRef]
  6. Camp, J.E.; Greatrex, B.W. Levoglucosenone: Bio-Based Platform for Drug Discovery. Front. Chem. 2022, 10, 902239. [Google Scholar] [CrossRef]
  7. Stanfield, M.K.; Terry, R.S.; Smith, J.A.; Thickett, S.C. Levoglucosan and levoglucosenone as bio-based platforms for polymer synthesis. Polym. Chem. 2023, 14, 4949–4956. [Google Scholar] [CrossRef]
  8. Pollard, B.; Gardiner, M.G.; Banwell, M.G.; Connal, L.A. Polymers from Cellulosic Waste: Direct Polymerization of Levoglucosenone using DBU as a Catalyst. ChemSusChem 2024, 17, e202301165. [Google Scholar] [CrossRef] [PubMed]
  9. Jung, E.; Rizzo, A.; Ryu, H.; Choa, M.; Choi, T.-L. Controlled polymerization of levoglucosenone derived enynes to give bio-based polymers with tunable degradation rates and high glass transition temperatures. Chem. Sci. 2025, 16, 8435–8442. [Google Scholar] [CrossRef] [PubMed]
  10. Blake, A.J.; Cook, T.A.; Forsyth, A.C.; Gould, R.O.; Paton, R.M. 1.3-Dipolar Cycloaddition Reactions of Levoglucosenone. Tetrahedron 1992, 48, 8053–8064. [Google Scholar] [CrossRef]
  11. Cicetti, S.; Spanevello, R.A.; Sarotti, A.M. In DFT We Trust: Exhaustive Exploration of 1,3-Dipolar Cycloadditions Between Nitrones and Levoglucosenone Exposesa Curious Case of Conformational Dynamics. Eur. J. Org. Chem. 2024, 27, e202400433. [Google Scholar] [CrossRef]
  12. Mlostoń, G.; Urbaniak, K.; Palusiak, M.; Witczak, Z.J.; Würthwein, E.-U. (3 + 2)-Cycloadditions of Levoglucosenone (LGO) with Fluorinated Nitrile Imines Derived from Trifluoroacetonitrile; An Experimental and Computational Study. Molecules 2023, 28, 7348. [Google Scholar] [CrossRef]
  13. Novikov, R.A.; Rafikov, R.R.; Shulishov, E.V.; Konyushkin, L.D.; Semenov, V.V.; Tomilov, Y.V. Reactions of levoglucosenone and its derivatives with diazo compounds. Russ. Chem. Bull. Int. Ed. 2009, 58, 327–334. [Google Scholar] [CrossRef]
  14. Mlostoń, G.; Celeda, M.; Palusiak, M. Higher-order (8 + 2) cycloadditions of tropothione with levoglucosenone (LGO) and structurally similar exo-cyclic enones derived from cyrene. Carbohydr. Res. 2023, 529, e108844. [Google Scholar] [CrossRef]
  15. Ward, D.D.; Shafizadeh, F. Cycloaddition “4 + 2” reactions of levoglucosenone. Carbohydr. Res. 1981, 95, 155–176. [Google Scholar] [CrossRef]
  16. Bhate, P.; Horton, D. Stereoselective synthesis of functionalized carbocycles by cycloaddition to levoglucosenone. Carbohydr. Res. 1983, 122, 189–199. [Google Scholar] [CrossRef]
  17. Pollard, B.; Liu, X.; Connal, L.A.; Banwell, M.G.; Gardiner, M.G. The synthesis and manipulation of certain Diels–Alder adducts of levoglucosenone and iso-levoglucosenone. Austr. J. Chem. 2023, 76, 797–811. [Google Scholar] [CrossRef]
  18. Giri, G.F.; Sarotti, A.M.; Spanevello, R.A. Understanding reactivity and regioselectivity in Diels–Alder reactions of a sugar-derived dienophile bearing two competing EWGs. An experimental and computational study. Carbohydr. Res. 2015, 415, 54–59. [Google Scholar] [CrossRef] [PubMed]
  19. Sarotti, A.M.; Spanevello, R.A.; Duhayon, C.; Suárez, A.G. Exploring structural effects of levoglucosenone derived chiral auxiliaries in asymmetric Diels–Alder cycloadditions. Tetrahedron 2007, 63, 241–251. [Google Scholar] [CrossRef]
  20. Comba, M.B.; Tsai, Y.; Sarotti, A.M.; Mangione, M.I.; Suárez, A.G.; Spanevello, R.A. Levoglucosenone and its new applications: Valorization of cellulose residues. Eur. J. Org. Chem. 2018, 2018, 590–604. [Google Scholar] [CrossRef]
  21. Tietze, L.; Kettschau, G. Hetero Diels-Alder reactions in organic chemistry. Top. Curr. Chem. 2015, 189, 1–120. [Google Scholar] [CrossRef]
  22. Eschenbrenner-Lux, V.; Kumar, K.; Waldmann, H. The asymmetric hetero-Diels-Alder reaction in the syntheses of biologically relevant compounds. Angew. Chem. Int. Ed. 2014, 53, 11146–11157. [Google Scholar] [CrossRef]
  23. Blond, G.; Gulea, M.; Mamane, V. Recent Contributions to Hetero Diels-Alder Reactions. Curr. Org. Chem. 2016, 20, 2161–2210. [Google Scholar] [CrossRef]
  24. Mlostoń, G.; Grzelak, P.; Hamera-Fałdyga, 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 Rel. Elem. 2017, 192, 204–211. [Google Scholar] [CrossRef]
  25. 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]
  26. Buday, P.; Seeber, P.; Neumann, C.; Abul-Futouh, H.; Görls, H.; Gräfe, S.; Matczak, P.; Kupfer, S.; Weigand, W.; Mloston, G. Iron(0) mediated stereoselective (3 + 2)-cycloaddition of thiochalcones via a diradical intermediate. Chem. Eur. J. 2020, 26, 11412–11416. [Google Scholar] [CrossRef]
  27. Matczak, P.; Kupfer, S.; Mlostoń, G.; Budday, P.; Görls, H.; Weigand, W. Metal-ligand bonding in tricarbonyliron(0) complexes bearing thiochalcone ligands. New J. Chem. 2022, 46, e12924. [Google Scholar] [CrossRef]
  28. 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][Green Version]
  29. Mlostoń, G.; Hamera-Fałdyga, R.; Heimgartner, H. First synthesis of ferrocenyl substituted thiochalcones and their [4+2]-cycloadditions with acetylenic dienophiles. J. Sulfur Chem. 2018, 39, 322–331. [Google Scholar] [CrossRef]
  30. Mousavi-Ebadia, M.; Safaei-Ghomi, J.; Nejad, M.J. Synthesis of thiopyran derivatives via [4 + 2] cycloaddition reactions. RSC Adv. 2025, 15, 11160. [Google Scholar] [CrossRef]
  31. 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]
  32. 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] [PubMed]
  33. Jasiński, R. On the Question of Stepwise [4 + 2] Cycloaddition Reactions and Their Stereochemical Aspects. Symmetry 2021, 13, 1911. [Google Scholar] [CrossRef]
  34. Mlostoń, G.; Urbaniak, K.; Urbaniak, P.; Marko, 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]
  35. Mlostoń, G.; Urbaniak, K.; Zimmer, R.; Reissig, H.-U.; Heimgartner, H. Hetero-Diels-Alder reactions of conjugated nitrosoalkenes with ferrocenyl, hetaryl and cylcoaliphatic thioketones. ChemistrySelect 2018, 3, 11724–11728. [Google Scholar] [CrossRef]
  36. Mlostoń, G.; Urbaniak, K.; Sobiecka, M.; Heimgartner, H.; Würthwein, E.-U.; Zimmer, R.; Lentz, D.; Reissig, H.-U. Hetero-Diels-Alder reactions of on in situ-generated azoalkenes with thioketones; Experimental and theoretical studies. Molecules 2021, 26, 2544. [Google Scholar] [CrossRef]
  37. Lopes, S.M.M.; Cardoso, A.L.; Lemos, A.; Pinho e Melo, T.M.V.D. Recent Advance in the Chemistry of Conjugated Nitrosoalkenes and Azoalkenes. Chem. Rev. 2018, 118, 11324–11352. [Google Scholar] [CrossRef]
  38. Sustmann, R.; Tappanchai, S.; Bandmann, H. α-(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]
  39. Becke, A.D. Density-functional thermochemistry. III. The role of exact exchange. J. Chem. Phys. 1993, 98, 5648–5652. [Google Scholar] [CrossRef]
  40. 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] [PubMed]
  41. 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]
  42. 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]
  43. Perdew, J.P.; Burke, K.; Ernzerhof, M. Generalized Gradient Approximation Made Simple. Phys. Rev. Lett. 1996, 77, 3865, Erratum in Phys. Rev. Lett. 1997, 78, 1396. https://doi.org/10.1103/PhysRevLett.77.3865. [Google Scholar] [CrossRef]
  44. 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]
  45. Ernzerhof, M.; Scuseria, G.E. Assessment of the Perdew–Burke–Ernzerhof exchange-correlation functional. J. Chem. Phys. 1999, 110, 5029–5036. [Google Scholar] [CrossRef]
  46. 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] [PubMed]
  47. Tomasi, J.; Mennucci, B.; Cammi, R. Quantum Mechanical Continuum Solvation Models. Chem. Rev. 2005, 105, 2999–3093. [Google Scholar] [CrossRef]
  48. Gaussian 16, Revision B.01; Frisch, M.J., Trucks, G.W., Schlegel, H.B., Scuseria, G.E., Robb, M.A., Cheeseman, R., Scalmani, G., Barone, V., Petersson, G.A., Nakatsuji, H., et al., Eds.; Gaussian, Inc.: Wallingford, CT, USA, 2016. [Google Scholar]
  49. Reed, A.E.; Weinstock, R.B.; Weinhold, F. Natural Population Analysis. J. Chem. Phys. 1985, 83, 735–746. [Google Scholar] [CrossRef]
  50. Reed, A.E.; Curtiss, L.A.; Weinhold, F. Intermolecular interactions from a natural bond orbital, donor-acceptor viewpoint. Chem. Rev. 1988, 88, 899–926. [Google Scholar] [CrossRef]
  51. Mlostoń, G.; Grzelak, P.; Utecht, G.; Jasiński, M. First [3 + 2]-cycloadditions of thiochalcones as C=S dipolarophiles in reactions with fluorinated nitrile imines. Synthesis 2017, 49, 2129–2137. [Google Scholar] [CrossRef]
  52. Hegab, M.I. A review on chemical and biological studies of thiopyran derivatives. Phosphorus Sulfur Silicon Relat. Elem. 2025, 200, 1–11. [Google Scholar] [CrossRef]
  53. Deepthi, A.; Leena, S.S.; Krishnan, D. Update on thiopyran-fused heterocycle synthesis (2013–2024). Org. Biomol. Chem. 2024, 22, 5676–5717. [Google Scholar] [CrossRef] [PubMed]
  54. van Vliet, L.A.; Rodenhuis, N.; Dijkstra, D.; Wikstrom, H.; Pugsley, T.A.; Serpa, K.A.; Meltzer, L.T.; Heffner, T.G.; Wise, L.D.; Lajiness, M.E.; et al. Synthesis and Pharmacological Evaluation of Thiopyran Analogues of the Dopamine D3 Receptor-Selective Agonist (4aR,10bR)-(+)-trans-3,4,4a,10b-Tetrahydro-4-n-propyl-2H,5H-[1]-benzopyrano [4,3-b]-1,4-oxazin-9-ol (PD 128907). J. Med. Chem. 2000, 43, 2871–2882. [Google Scholar] [CrossRef]
  55. Baghipour, T.; Khalilzadeh, M.A.; Rajabi, M.; Mehrzad, J. Synthesis and antiproliferative activity of diethyl 5-acetyl-4-methyl-6-(2-fluorophenylimino)-6H-thiopyran-2,3-dicarboxylate (3TM). Afr. J. Biotechnol. 2011, 10, 14155–14159. [Google Scholar] [CrossRef]
  56. Pasha, G.F.; Asghari, S.; Tajbakhsh, M.; Mohseni, M. Synthesis and antimicrobial evaluation of some new bicyclopyrazolone-based thiopyran ring systems. J. Chin. Chem. Soc. 2019, 66, 600–667. [Google Scholar] [CrossRef]
  57. Brown, M.J.; Carter, P.S.; Fenwick, A.E.; Fosberry, A.P.; Hamprecht, D.W.; Hibbs, M.J.; Jarvest, R.L.; Mensah, L.; Milner, P.H.; O’Hanlon, P.J.; et al. The antimicrobial natural product chuangxinmycin and some synthetic analogues are potent and selective inhibitors of bacterial tryptophanyl tRNA synthetase. Bioorg. Med. Chem. Lett. 2002, 12, 3171–3173. [Google Scholar] [CrossRef]
  58. Jatin; Murugappan, S.; Kirad, S.; Ala, C.; Kuthe, P.V.; Kondapalli, C.S.V.G.; Sankaranarayanan, M. Thiochromenes and thiochromanes: A comprehensive review of their diverse biological activities and structure–activity relationship (SAR) insights. RSC Med. Chem. 2025, 16, 1941–1968. [Google Scholar] [CrossRef] [PubMed]
  59. Wilde, R.G.; Billheimer, J.T.; Germain, S.J.; Hausner, E.A.; Meunier, P.C.; Munzer, D.A.; Stoltenborg, J.K.; Gillies, P.J.; Burcham, D.L.; Huang, S.-M.; et al. ACAT Inhibitors derived from hetero-Diels-Alder cycloadducts of thioaldehydes. Bioorg. Med. Chem. 1996, 4, 1493–1513. [Google Scholar] [CrossRef]
  60. CrysAlisPRO Software System, Oxford Diffraction/Agilent Technologies UK Ltd.: Yarnton, UK, 2015.
  61. 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]
  62. 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]
  63. Sheldrick, G.M. Crystal structure refinement with SHELXL. Acta Crystallogr. Sect. C Struct. Chem. 2015, 71, 3–8. [Google Scholar] [CrossRef] [PubMed]
  64. Spek, A.L. Structure validation in chemical crystallography. Acta Crystallogr. Sect. D Biol. Crystallogr. 2009, 65, 148–155. [Google Scholar] [CrossRef]
  65. Groom, C.R.; Bruno, I.J.; Lightfoot, M.P.; Ward, S.C. The Cambridge Structural Database. Acta Cryst. 2016, 72, 171–179. [Google Scholar] [CrossRef] [PubMed]
Figure 1. (−)-Levoglucosenone (LGO) (1) and presentation of two competitive approaches to the activated C=C bond.
Figure 1. (−)-Levoglucosenone (LGO) (1) and presentation of two competitive approaches to the activated C=C bond.
Molecules 30 03783 g001
Scheme 1. Selected examples of 1,3-dipolar cycloadditions, a higher order cycloaddition or Diels–Alder reactions starting with levoglucosenone (LGO) (1), and [a] nitrile oxides [10]; [b] fluorinated nitrile imines [12]; [c] tropothione [14]; [d] cyclopentadiene [15,16]; [e] buta-1,3-diene [15].
Scheme 1. Selected examples of 1,3-dipolar cycloadditions, a higher order cycloaddition or Diels–Alder reactions starting with levoglucosenone (LGO) (1), and [a] nitrile oxides [10]; [b] fluorinated nitrile imines [12]; [c] tropothione [14]; [d] cyclopentadiene [15,16]; [e] buta-1,3-diene [15].
Molecules 30 03783 sch001
Scheme 2. Hetero-Diels–Alder reactions of thiochalcones (2) with acetylenic dienophiles leading to functionalized thiopyrans [28,29] and the possible course of their reactions with levoglucosenone (1), providing stereoisomers of regioisomeric tricyclic products 3 or 4.
Scheme 2. Hetero-Diels–Alder reactions of thiochalcones (2) with acetylenic dienophiles leading to functionalized thiopyrans [28,29] and the possible course of their reactions with levoglucosenone (1), providing stereoisomers of regioisomeric tricyclic products 3 or 4.
Molecules 30 03783 sch002
Scheme 3. Regioselective hetero-Diels–Alder reactions of levoglucosenone (LGO) (1) with thiochalcones 2af leading to the formation of isomeric exo,exo- and exo,endo-cycloadducts 3af (tricyclic 3,4-dihydro-4H-thiopyrans).
Scheme 3. Regioselective hetero-Diels–Alder reactions of levoglucosenone (LGO) (1) with thiochalcones 2af leading to the formation of isomeric exo,exo- and exo,endo-cycloadducts 3af (tricyclic 3,4-dihydro-4H-thiopyrans).
Molecules 30 03783 sch003
Figure 2. Molecular structures of the (4+2) cycloadducts (exo,exo)-3b (major, left) and (exo,endo)-3b (minor, right). Atoms are represented by thermal ellipsoids (50%). For graphics with atom labelling, see Figure S10a,b, respectively (Supplementary Materials).
Figure 2. Molecular structures of the (4+2) cycloadducts (exo,exo)-3b (major, left) and (exo,endo)-3b (minor, right). Atoms are represented by thermal ellipsoids (50%). For graphics with atom labelling, see Figure S10a,b, respectively (Supplementary Materials).
Molecules 30 03783 g002
Scheme 4. Epimerization process of less stable (exo,endo)-3c in the course of the attempted chromatographic purification (SiO2 column), leading to (exo,exo)-3c.
Scheme 4. Epimerization process of less stable (exo,endo)-3c in the course of the attempted chromatographic purification (SiO2 column), leading to (exo,exo)-3c.
Molecules 30 03783 sch004
Figure 3. The in situ-generated electron-deficient heterodienes 5 and 6 examined in reactions with levoglucosenone (1).
Figure 3. The in situ-generated electron-deficient heterodienes 5 and 6 examined in reactions with levoglucosenone (1).
Molecules 30 03783 g003
Scheme 5. Presentation of four possible transition states in the concerted hetero-Diels–Alder reactions of 1 with thiochalcones 2, leading to tricyclic thiopyrans 3.
Scheme 5. Presentation of four possible transition states in the concerted hetero-Diels–Alder reactions of 1 with thiochalcones 2, leading to tricyclic thiopyrans 3.
Molecules 30 03783 sch005
Scheme 6. DFT results of the (4+2) cycloadditions of thiochalcone 2a with levoglucosenone (LGO) (1) (PBE1PBE/def2tzvp + GD3BJ + PCM-tetrahydrofuran) (DG298 [kcal/mol]) leading to the four stereoisomeric tricyclic compounds 3a (attack of the S-terminus of 2a to the β-C atom of 1). Symbol : transition state.
Scheme 6. DFT results of the (4+2) cycloadditions of thiochalcone 2a with levoglucosenone (LGO) (1) (PBE1PBE/def2tzvp + GD3BJ + PCM-tetrahydrofuran) (DG298 [kcal/mol]) leading to the four stereoisomeric tricyclic compounds 3a (attack of the S-terminus of 2a to the β-C atom of 1). Symbol : transition state.
Molecules 30 03783 sch006
Scheme 7. DFT results of the alternative (4+2)-cycloadditions of thiochalcone 2a with levoglucosenone (1) (PBE1PBE/def2tzvp + GD3BJ + PCM-dichloromethane) (DG298 [kcal/mol]) leading to four stereoisomeric tricyclic compounds 4a (attack of the S-terminus of 2a to the α-carbon atom of 1). Symbol : transition state.
Scheme 7. DFT results of the alternative (4+2)-cycloadditions of thiochalcone 2a with levoglucosenone (1) (PBE1PBE/def2tzvp + GD3BJ + PCM-dichloromethane) (DG298 [kcal/mol]) leading to four stereoisomeric tricyclic compounds 4a (attack of the S-terminus of 2a to the α-carbon atom of 1). Symbol : transition state.
Molecules 30 03783 sch007
Figure 4. (a) TS-(exo,exo)-3a, distance C⋯S: 2.289 Å, distance C⋯C: 2.608 Å. (b) TS-(exo,endo)-3a, distance C⋯S: 2.270 Å, distance C⋯C: 2.539 Å.
Figure 4. (a) TS-(exo,exo)-3a, distance C⋯S: 2.289 Å, distance C⋯C: 2.608 Å. (b) TS-(exo,endo)-3a, distance C⋯S: 2.270 Å, distance C⋯C: 2.539 Å.
Molecules 30 03783 g004
Figure 5. (a) TS-(exo,exo)-4a, distance C⋯S: 2.424 Å, distance C⋯C: 2.259 Å. (b) TS-(exo,endo)-4a, distance C⋯S: 2.394 Å, distance C⋯C: 2.342 Å.
Figure 5. (a) TS-(exo,exo)-4a, distance C⋯S: 2.424 Å, distance C⋯C: 2.259 Å. (b) TS-(exo,endo)-4a, distance C⋯S: 2.394 Å, distance C⋯C: 2.342 Å.
Molecules 30 03783 g005
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.; Würthwein, E.-U.; Reissig, H.-U.; Witczak, Z.J. A Remarkable Selectivity Observed in Hetero-Diels–Alder Reactions of Levoglucosenone (LGO) with Thiochalcones: An Experimental and Computational Study. Molecules 2025, 30, 3783. https://doi.org/10.3390/molecules30183783

AMA Style

Mlostoń G, Urbaniak K, Palusiak M, Würthwein E-U, Reissig H-U, Witczak ZJ. A Remarkable Selectivity Observed in Hetero-Diels–Alder Reactions of Levoglucosenone (LGO) with Thiochalcones: An Experimental and Computational Study. Molecules. 2025; 30(18):3783. https://doi.org/10.3390/molecules30183783

Chicago/Turabian Style

Mlostoń, Grzegorz, Katarzyna Urbaniak, Marcin Palusiak, Ernst-Ulrich Würthwein, Hans-Ulrich Reissig, and Zbigniew J. Witczak. 2025. "A Remarkable Selectivity Observed in Hetero-Diels–Alder Reactions of Levoglucosenone (LGO) with Thiochalcones: An Experimental and Computational Study" Molecules 30, no. 18: 3783. https://doi.org/10.3390/molecules30183783

APA Style

Mlostoń, G., Urbaniak, K., Palusiak, M., Würthwein, E.-U., Reissig, H.-U., & Witczak, Z. J. (2025). A Remarkable Selectivity Observed in Hetero-Diels–Alder Reactions of Levoglucosenone (LGO) with Thiochalcones: An Experimental and Computational Study. Molecules, 30(18), 3783. https://doi.org/10.3390/molecules30183783

Article Metrics

Back to TopTop