Next Article in Journal
Hydrolyzed Milk-Derived Peptides Promote Erythropoietin Pathways and Hematologic Recovery: A Cross-Species Analysis
Previous Article in Journal
Vinyl Chloride Degradation Using Ozone-Based Advanced Oxidation Processes: Bridging Groundwater Treatment and Machine Learning for Smarter Solutions
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

On the Question of the Regio-, Stereoselectivity and the Molecular Mechanism of the (3+2) Cycloaddition Reaction Between (Z)-C-Phenyl-N-alkyl(phenyl)nitrones and (E)-3-(Methylsulfonyl)-propenoic Acid Derivatives

by
Martyna Ząbkowska
1,
Karolina Kula
1,*,
Volodymyr Diychuk
2 and
Radomir Jasiński
1,*
1
Cracow University of Technology, Faculty of Chemical Engineering and Technology, Warszawska 24, 31-155 Cracow, Poland
2
Yuriy Fedkovych Chernivtsi National University, Department of Chemistry and Food Examination, Kotsiubynskoho 2, 58-002 Chernivtsi, Ukraine
*
Authors to whom correspondence should be addressed.
Molecules 2025, 30(24), 4738; https://doi.org/10.3390/molecules30244738
Submission received: 27 November 2025 / Revised: 7 December 2025 / Accepted: 9 December 2025 / Published: 11 December 2025

Abstract

In this work, the regio- and stereochemistry as well as the molecular mechanism of the cycloaddition reaction of nitrones with (E)-3-(methylsulfonyl)-propenoic acid derivatives were analyzed based on ωb97xD/6-311G(d,p) quantum chemical calculations. In light of these data, it is possible to propose selectivity of the analyzed processes, which was not clearly determined in light of previous experimental studies. Furthermore, the mechanism of the process was diagnosed. CDFT descriptors indicate that the reaction is triggered by a nucleophilic attack of the nitrone oxygen atom on the electrophilic carbon atom of (E)-3-(methylsulfonyl)-propenoic acid derivatives. In turn, PES analysis shows that, despite the nucleophilic-electrophilic character of the reactants, the corresponding transition states are only weakly polar and highly synchronous. IRC calculations rule out zwitterionic or biradical intermediates, confirming a single-step mechanism. The in silico ADME and PASS predictions indicate that the resulting isoxazolidines possess promising biological profiles, showing potential modulation of the serotonin system through 5-HT2A and 5-HT2C antagonism and stimulation of serotonin release, with structural features compatible with P450-mediated metabolism. Considering this attractive application potential, a detailed mechanistic investigation of their formation becomes essential for understanding and ultimately controlling the reaction pathways leading to these heterocycles.

1. Introduction

Isoxazolidines (tetrahydroisoxazoles) are an important group of building blocks used in modern chemical technology, biotechnology and pharmaceutical research [1,2]. They occur in numerous functional materials and bioactive structures. In particular, isoxazoline scaffolds have attracted growing interest in agrochemical research, where they serve as structural motifs in herbicides, insecticides and fungicides [3,4]. At the same time, isoxazolidine frameworks remain widely exploited for the design of compounds with promising pharmacological potential, such as antitumor agents [5,6], antibacterial substances [7,8], anti-inflammatory agents [9,10] and many other biologically active molecules [11,12,13]. Isoxazolidines containing good leaving groups (LGs) [14] are often prone to ring transformation into β-lactams [15,16,17], which form a well-known molecular core of many clinically significant antibiotics [18,19,20].
The most universal strategy for the synthesis of isoxazolidines relies on (3+2) cycloaddition (32CA) reactions of nitrones [21,22,23]. This approach benefits from the broad availability of three-atom components (TACs) and a wide variety of suitable alkenes [24,25], which together create a highly flexible synthetic platform. The method offers excellent atom economy, reaching 100 percent under ideal conditions [26,27], and provides access to a broad spectrum of heterocycles. A central advantage of nitrone cycloaddition chemistry is the high level of stereocontrol that can be achieved in these reactions. Both the configuration of the nitrone and the geometry of the alkene strongly influence the relative and absolute stereochemistry of the resulting isoxazolidines [28,29,30]. This feature is particularly valuable for applications in medicinal chemistry, where precise spatial arrangement often determines biological activity [31]. Advances in catalytic and asymmetric variants of nitrone cycloadditions have further expanded the stereochemical versatility of this methodology [32,33].
Some years ago, Chanet-Ray and coworkers [34] experimentally investigated the 32CA between (Z)-C-phenyl-N-methylnitrone (1) and the methyl ester of (E)-3-(methylsulfonyl)-propenoic acid (2). They reported that this reaction could proceed under mild conditions, affording a mixture of three out of the four theoretically possible regio- and stereoisomeric 2-methyl-3-phenylisoxazolidines. Similarly, a number of other reactions involving nitrones and (E)-3-(methylsulfonyl)-propenoic acid derivatives (1ac + 2 or 3) were analyzed (Scheme 1).
However, the account of these transformations in their paper is incomplete and includes several omissions and ambiguities. In particular, the following points require further clarification:
(i) Regio- and stereochemistry. The regio- and stereochemical outcome of the reaction remains unclear. The authors only partially separated the post-reaction mixture and inferred the product constitution from partial NMR data recorded on a low-frequency instrument. The only information available indicates that, in each case, the most favored product should be the corresponding 3,4-trans-3-phenyl-4-methylsulfonyl-isoxazolidine (4 or 8). Notably, HPLC and NMR analyses of the crude reaction mixtures, which are standard practices for assessing regio- and stereoselectivity in 32CA reactions [35,36,37], were never performed. Consequently, it is unknown whether other regio- and stereoisomeric adducts were formed, particularly since the yield of isolated products was sometimes as low as a few percent.
(ii) Reaction selectivity. The selectivity of the reaction is not fully interpreted. As a result, no reliable conclusions can be drawn regarding the likely selectivity of analogous reactions involving alkylsulfonyl-substituted alkenes.
(iii) Reaction mechanism. The mechanism of the reaction remains undetermined. Until about 30 years ago, all 32CA reactions were generally assumed to proceed via a single-step “concerted” mechanism [38,39,40]. Current knowledge, however, indicates that some 32CA reactions proceed via a two-step pathway involving either a zwitterionic [41,42,43] or a diradical intermediate [44,45,46]. Moreover, even in single-step reactions, electron density reorganization does not necessarily occur in a purely pericyclic manner; rather, it involves a complex sequence of stages, with multiple directions and modes of electron density transfer between molecular fragments [47,48,49,50].
Considering the issues outlined above, we carried out a comprehensive quantum-chemical investigation of the studied reaction using both the Conceptual Density Functional Theory (CDFT) [51,52,53] and the Molecular Electron Density Theory (MEDT) [24]. Our goal was to obtain a complete, mechanistically grounded picture of the factors controlling its reactivity and selectivity. First, we examined the electronic structures of all reactants, identifying the most electrophilic and nucleophilic reactive centers and clarifying their roles in the initial stage of the cycloaddition. We then simulated all theoretically feasible reaction pathways, which enabled us to characterize their mechanisms, energetic profiles, and the factors governing regio- and stereoselectivity. Finally, we analyzed the reorganization of the electron density along the reaction coordinate of the model process to determine which electronic rearrangements dictate the course of the reaction and the stability of the associated intermediates and products.
Additionally, we performed preliminary in silico studies to evaluate the potential biological activity of the cycloaddition products (411). This approach is particularly important because the biological activity of these isoxazolidines has not yet been studied experimentally. No activity has been reported, and no attempts have been made to identify them as potential APIs. These analyses employed both ADME (Absorption, Distribution, Metabolism, and Excretion) [54] and PASS (Prediction of Activity Spectra for Substances) [55] predictions to assess pharmacokinetic parameters, drug-likeness, and possible activity spectra. The results provide initial but promising evidence for the pharmacological relevance of these compounds and complement the mechanistic investigations presented in this work. Altogether, the combination of detailed mechanistic insights and bioactivity predictions offers a coherent understanding of the reactivity and functional significance of the resulting isoxazolidines. This integrated approach offers a clear basis for understanding the factors that influence the formation and stability of sulfonyl-substituted isoxazolidines in 32CA reactions and helps place their potential biological relevance in a broader synthetic context. Approaches of this kind are increasingly used to relate mechanistic features to functional properties, making them a practical and well-established element of modern studies on cycloaddition processes [56,57]. We believe that this aspect significantly strengthens the study and enhances its potential interest to the research community.

2. Results and Discussion

2.1. Analysis of the Reaction Mechanism

First, we decided to examine the mechanistic aspects of the studied 32CA reaction in a systematic and comprehensive manner. To achieve this, we began with an in-depth analysis of the electronic structures of the reacting substrates within the framework of CDFT. This approach allowed us to identify the most reactive centers, assess polarity patterns, and anticipate the general reactivity profile expected for the cycloaddition. After establishing the electronic characteristics of both partners, we proceeded to investigate the full reaction pathway by exploring the potential energy surface (PES). This analysis made it possible to evaluate the energetic feasibility of the individual pathways, determine the origin of regio- and stereochemical preferences, and clarify the fundamental factors governing the reactivity of this system.

2.1.1. Global and Local Interactions Between Reactants

CDFT is one of the essential approaches for analyzing the reactivity of molecules in polar reactions. This framework links classical chemical descriptors, such as electronic chemical potential (μ) as well as chemical hardness (η), with the electronic structure of molecules. Using these parameters, one can determine global parameters, including electrophilicity (ω) and nucleophilicity (N), which help assign the role of each reactant as either an electrophile or a nucleophile in a given reaction [58,59]. Such information is particularly important for predicting which regioisomers are likely to form [60,61]. Thus, the global reactivity indices of (Z)-C-phenyl-N-alkyl(phenyl)nitrones (1ac) and (E)-3-(methylsulfonyl)-propenoic acid derivatives (2 and 3) are presented in Table 1.
The applicability of CDFT depends strongly on the polarity of the studied processes [62,63], as this approach is valid only for polar reactions. Therefore, a preliminary assessment of the reaction polarity is required before any further analysis. To this end, the electronic chemical potential μ was examined. Thus, the electronic chemical potential μ of the (Z)-C-phenyl-N-alkyl(phenyl)nitrones (1ac) is markedly higher than that of the corresponding (E)-3-(methylsulfonyl)-propenoic acid derivatives (2 and 3) (Table 1). This difference indicates that the electron density will flow from nitrones (1ac) toward ethenes (2 and 3) (Figure 1). Consequently, the reaction between these reagents can be classified as a forward electron density flux (FEDF) process [64]. Additionally, the difference in global electrophilicity between nitrones (1ac) and (E)-3-(methylsulfonyl)-propenoic acid derivatives (2 and 3) suggests that the investigated 32CAs exhibit only weak polar character (Table 1).
Subsequently, reactivity descriptors allowed us to assign the electrophilic and nucleophilic nature of the studied reagents. The computed global electrophilicity (ω) values [58] of the (Z)-C-phenyl-N-alkyl(phenyl)nitrones (1ac) are 0.84 eV for 1a (-Me), 0.80 eV for 1b (-Ph), and 1.05 eV for 1c (-t-Bu) (Table 1). The calculated global nucleophilicity (N) values [59] for these compounds are 3.73 eV for 1a (-Me), 3.79 eV for 1b (-Ph), and 3.75 eV for 1c (-t-Bu) (Table 1). These values clearly indicate that all examined nitrones (1ac) behave as moderate electrophiles while simultaneously acting as strong nucleophilic agents.
In turn, for both (E)-3-(methylsulfonyl)-propenoic acid derivatives (2 and 3), the computed global electrophilicity (ω) values [58] are 1.75 eV and 1.49 eV, respectively (Table 1). These results suggest that both ethenes (2 and 3) behave as strong electrophiles, with a considerably higher electrophilic character compared to the examined nitrones (1ac). The computed global nucleophilicity (N) values [59] are 0.57 eV and 1.16 eV, respectively (Table 1). These results indicate that both ethenes (2 and 3) exhibit only marginal nucleophilic character, with global nucleophilicity values drastically lower than those of the examined nitrones (1ac).
Overall, in all the considered reactions, it can be concluded that the (Z)-C-phenyl-N-alkyl(phenyl)nitrones (1ac) act as nucleophilic components, whereas the (E)-3-(methylsulfonyl)-propenoic acid derivatives (2 and 3) serve as electrophilic agents. This observation provides valuable guidance for predicting local reactivity in these systems. In reactions involving non-symmetric reagents, regioselectivity is typically governed by the interaction between the most electrophilic site of the electrophile and the most nucleophilic site of the nucleophile [65,66]. To identify these reactive centers, the electrophilic Pk+ and nucleophilic Pk Parr functions were computed, allowing the localization of the most electrophilic positions in ethenes (2 and 3) and the most nucleophilic positions in nitrones (1ae) (Figure 2).
The analysis of the nucleophilic Pk Parr functions for the (Z)-C-phenyl-N-alkyl(phenyl)nitrones (1ac) indicates that, in all cases and irrespective of the alkyl substituent, the oxygen atom constitutes the most nucleophilic center of the molecule. The values for PN are 0.57 for 1a (-Me), 0.50 for 1b (-Ph), and 0.57 for 1c (-t-Bu) (Figure 2). In turn, the analysis of the electrophilic Pk+ Parr functions for the (E)-3-(methylsulfonyl)-propenoic acid derivatives (2 and 3) shows that both vinyl carbon atoms display comparable electrophilic characteristics. However, the carbon atom bonded to the methylsulfonyl group is consistently the more electrophilic site. Its Pk+ values amount to 0.35 for 2 (-COOMe) and 0.40 for 3 (-CN) (Figure 2). By contrast, the carbon atom connected to the EWG substituent exhibits lower electrophilicity, with Pk+ values of 0.29 for 2 (-COOMe) and 0.38 for 3 (-CN) (Figure 2).
Analysis of intermolecular interactions, based on the older but still sometimes used Frontier Molecular Orbitals (FMO) theory [38,39], leads to similar conclusions. For all investigated reactions, the energy gap between the HOMO level of all nitrones (1ac) and the LUMO level of the tested alkenes (2 and 3) are markedly smaller than the alternative gap, that is, between these alkenes (2 and 3) HOMO and nitrones (1ac) LUMO (Table 1). This indicates that the discussed processes proceed under so-called normal orbital control [40], with the reaction course governed by electron transfer from the nitrone to the alkene. Examination of the local orbital function distributions shows that the process is consistently initiated by a nucleophilic attack of the oxygen atom within the CNO fragment of the nitrone. This arises from the fact that this atom exhibits a markedly higher HOMO density than the remaining atoms of the nitrone fragment, making it the dominant nucleophilic center (Figure 3). Mentioned nucleophilic attack may occur at either of the electrophilically activated centers of the alkene, as both corresponding LUMOs exhibit comparable shape and amplitude, providing similarly accessible sites for electron transfer (Figure 3).
The presented results correlate well with the molecular electrostatic potential (MEP). In particular, the analysis of the MEP surfaces for all (Z)-C-phenyl-N-alkyl(phenyl)-nitrones (1ac) reveals a region of highly negative electrostatic potential around the oxygen atom (shown in red) (Figure 4). This indicates that all studied nitrones (1ac) act as nucleophiles, with the oxygen atom being the primary site for electron donation in the reaction. On the other hand, the both vinyl carbon atoms in (E)-3-(methylsulfonyl)propenoic acid derivatives (2 and 3) exhibit a slightly negative electrostatic potential (shown in light green) (Figure 4), indicating that these centers are equivalent and can act as electrophilic sites in the cycloaddition reaction.

2.1.2. PES Exploration and Characterization of Critical Structures in the Studied CAs

Having established the general nature of the intermolecular interactions governing the studied processes, we next examined in detail the critical structures that define their reaction pathways. Accordingly, we undertook the systematic localization and characterization of all relevant transition states as well as any potential intermediates. This analysis was carried out using a computational protocol that has been extensively validated in previous theoretical investigations of various cycloaddition reactions [67,68,69], including numerous studies focused on nitrone-based systems [70,71]. The outcomes of these calculations, encompassing the structural features, energetics, and diagnostic reactivity parameters of the key stationary points, are comprehensively summarized in Table 2 and Table 3.
As the primary model, we examined the 1a + 2 reaction. Although the overall character of the reaction profiles for the individual competing pathways (AD) is broadly similar, the quantitative description of the corresponding critical points varies. For each pathway, two critical points were identified between the energy level of the separated reactants and the final product well: a molecular pre-reaction complex (MC) and a transition state (TS). In every case, the first stage of the transformation involves the formation of a molecular pre-reaction complex (MCA for pathway A, MCB for pathway B, MCC for pathway C, and MCD for pathway D). This step is accompanied by a pronounced decrease in the enthalpy of the system (exceeding 10 kcal mol−1) and importantly proceeds without any activation barrier (Table 2).
It should be emphasized that already at this early stage of the reaction, the addend substructures adopt a mutual spatial orientation identical to that later observed in the transition state and ultimately retained in the product. Consequently, the identified intermediates should be regarded as orientational complexes. Similar complexes have previously been located using DFT methods in 32CA pathways involving N-nitrile oxides [72,73], azides [74,75], diazo compounds [75,76,77], ylides [78], and other 4-π-electron components [79,80]. Experimental evidence for the formation of pre-reaction complexes has also been reported for the reactions of ozone with ethene [81] and acetylene [82,83], based on microwave spectrometry data [84,85,86].
Although the reacting centers are already aligned in their target positions, the formation of new bonds has not yet commenced. Both the C3–C4 and C5–O1 interatomic distances remain outside the ranges typical for nascent C–C and C–O bonds in transition states [87,88,89]. This indicates that no transfer of electron density should occur between the substructures. Analysis of the GEDT indices confirms this expectation (Table 3).
The gradual approach of the reactive centers leads to the conversion of the pre-reaction complexes into the corresponding transition states (TSA for pathway A, TSB for pathway B, TSC for pathway C, and TSD for pathway D, respectively). Their character was confirmed by vibrational analysis, which in each case revealed a single imaginary frequency. Formation of these TSs is accompanied by an increase in the reaction-system enthalpy by several kcal/mol (Table 2). Notably, these values are relatively low compared with other 32CAs involving nitrones [90,91].
The kinetically most favorable pathway is pathway A. This conclusion is fully consistent with the experimentally observed selectivity: 3,4-trans-4,5-trans-2-methyl-3-phenyl-4-methylsulfonyl-5-carbomethoxyisoxazolidine 4a was indeed identified in the reaction mixture (Scheme 1) [34]. In addition, two other products were detected, most likely the 3,4-cis and 3,4-trans isomers of 2-methyl-3-phenyl-4-carbomethoxy-5-ethylsulfonylisoxazolidines 6a and 7a [34]. Pathway B, which would lead to 3,4-cis-4,5-trans-2-methyl-3-phenyl-4-methylsulfonyl-5-carbomethoxyisoxazolidine 5a, should be regarded as kinetically inaccessible under the examined competitive conditions. Considering the activation barriers, the reaction pathways can be ordered as A < C < D < B (Table 2).
All localized transition states adopt a biplanar geometry characteristic of 32CA reactions involving allylic TACs. In each case, two new bonds essential for forming the isoxazolidine ring, C3–C4 and C5–O1, are established simultaneously (Figure 5). These structures display a relatively high degree of synchronicity, which is expected because both reaction centers on the alkene are electrophilically activated by EWG substituents (Table 3).
Given the nucleophilic character of the nitrone and the electrophilic character of the alkene, a highly polar transition state could be anticipated in this type of reaction, as noted previously for related nitrone 32CAs [92]. However, the transition structures identified here do not display high polarity. The GEDT values for all localized TSs remain below 0.15 e (Table 3), which confirms their low-polar character. This finding clarifies the apparent inconsistency between the predictions based on global and local CDFT reactivity indices and the actual properties of the transition states. Although the CDFT descriptors correctly classify nitrones as nucleophiles and alkenes as electrophiles, the mutual activation of both reacting centers leads to synchronous bond formation with only limited charge separation at the transition state.
IRC analyses show that, for each pathway, the localized TS is the only critical point connecting the pre-reaction complex valley with the product (Figure 5). Attempts to optimize potential acyclic intermediates, whether biradical or zwitterionic, were unsuccessful. These results indicate that the reaction proceeds through a concerted, single-step mechanism characterized by low polarity and a high degree of synchronicity.
Quantum-chemical calculations carried out for the remaining reactions in the series provided analogous results. The conclusions regarding both selectivity and the molecular mechanism are consistent across all systems examined.

2.2. Predicted Biological Activity of the Isoxazolidine Derivatives Formed in the Studied 32CAs

Isoxazolidines and their derivatives constitute a versatile class of five-membered heterocycles whose relevance in medicinal chemistry continues to grow due to their diverse biological and pharmacological properties [93,94]. Their saturated N,O-heterocyclic framework combines conformational rigidity with favorable hydrogen-bonding capabilities, which makes isoxazolidines valuable structural motifs in the design of bioactive molecules and enzyme inhibitors, comparable to other privileged scaffolds such as morpholines and piperazines [95,96]. These compounds exhibit a broad spectrum of biological activities, including anticancer, antibacterial, antifungal, antioxidant, and anti-inflammatory effects [97,98]. Certain conjugates have demonstrated potent cytotoxicity against human cancer cell lines with low-micromolar IC50 values, while others showed significant antibacterial and antifungal activity in biochemical assays [99,100]. These characteristics make isoxazolidine derivatives promising candidates for further exploration in drug discovery and molecular design.
In view of the diverse applications documented for related isoxazolidine scaffolds, we decided to examine the prospective application potential of the heterocycles characterized in this work. The fundamental physicochemical characteristics, together with selected pharmacokinetic properties, were evaluated using the SwissADME online platform [101]. These results were subsequently analyzed in light of the commonly accepted drug-likeness criteria proposed by Lipinski et al. [102], Ghose et al. [103], Veber et al. [104], Egan et al. [105], as well as Muegge et al. [106]. In addition, the probable biological activities and potential application routes of the investigated isoxazolidine derivatives (411) were predicted through the PASS 2024 software [107].
It should be underline that, because both SwissADME and PASS rely on two-dimensional structural descriptors generated from canonical SMILES, which do not reliably differentiate individual stereoisomers, all computational predictions in this study were performed using a single representative configuration for each isoxazolidine derivative. This approach ensures methodological consistency and prevents the introduction of artifacts resulting from incomplete stereochemical encoding in the input structure. It is also consistent with current cheminformatics practice, as tools operating on SMILES-based fingerprints or similarity metrics typically provide identical or highly similar output for molecules that differ only in stereochemistry. Consequently, the ADME parameters, drug-likeness assessment and predicted biological activity profiles reported in this work correspond to the same defined structural representation for each compound [108,109,110]. These analyses constitute a foundational evaluation intended to deliver initial understanding of the ADME and predicted biological profiles of the investigated isoxazolidine derivatives (411) and to inform subsequent in-depth studies.

2.2.1. Computational Evaluation of ADME and Pharmacokinetic Profiles

ADME analyses are crucial in the early stages of drug discovery because obtaining experimental pharmacokinetic data is often time-consuming and expensive. As a result, computational methods are increasingly used to predict ADME profiles and improve the efficiency of preliminary screening. In this study the key physicochemical properties of the investigated isoxazolidine derivatives (411), including lipophilicity, solubility, predicted pharmacokinetic behavior and overall medicinal chemistry suitability, were evaluated. The drug-likeness of these molecules was assessed using the SwissADME platform [111], which provides reliable in silico predictions that support the identification and optimization of potential drug candidates [112,113]. The calculated parameters for compounds (411) are summarized in Table 4, whereas their bioavailability radars are shown in Figure 6.
Analysis of the in silico physicochemical parameters and predicted ADME profiles summarized in Table 4 shows that all investigated isoxazolidine derivatives (411) comply with the major drug-likeness criteria applied in this study. In particular, every compound in this series satisfies the accepted limits for molecular weight, number of rotatable bonds, and hydrogen-bond donor/acceptor counts, and displays favorable Topological Polar Surface Area (TPSA) values (Table 4). TPSA, which represents the total surface area of polar atoms (typically nitrogen, oxygen, and sulfur), is a key indicator of membrane permeability and overall drug bioavailability. In general, TPSA values below 140 Å2 are associated with good oral bioavailability, whereas values below 90 Å2 tend to support blood–brain barrier permeation [114,115]. For all considered compounds (411), the calculated TPSA values are close to 80 Å2 (Table 4), which is consistent with good membrane permeability and overall favorable oral bioavailability, according to these computational predictions.
Molar refractivity (MR) is another highly relevant parameter because it describes both the polarizability and the steric volume of a molecule, which together determine how effectively the compound interacts with biological targets. According to the Ghose filter, the optimal MR range spans 40–130 cm3/mol, which corresponds to molecular dimensions and electronic responsiveness that typically support efficient and selective binding within protein active sites [103]. Compounds with MR values below this threshold are often small and only weakly polarizable, which may limit the strength of their noncovalent interactions and lead to rapid metabolic clearance, ultimately reducing pharmacological potency. In contrast, excessively high MR values may indicate bulky, rigid structures with reduced conformational adaptability. Such molecules frequently suffer from impaired membrane permeability, diminished solubility, and an increased propensity for nonspecific interactions that can negatively impact bioavailability and safety [116,117]. All isoxazolidine derivatives (411) display MR values in the range of 69–96 cm3/mol (Table 4), which, according to these in silico estimates, fall well within the optimal interval, confirming that the compounds possess balanced molecular size and polarizability, consistent with favorable interaction potential toward biological targets and promising overall computationally predicted pharmacokinetic behavior.
The evaluation of lipophilicity and aqueous solubility indicates that all isoxazolidines (411) display favorable lipophilic profiles and sufficient water solubility. These properties play a central role in absorption and metabolic processing because lipophilicity facilitates membrane permeation, while aqueous solubility enables adequate compound distribution in physiological fluids. Both features contribute directly to effective gastrointestinal uptake [118,119], and lipophilicity additionally modulates susceptibility to metabolic enzymes, thereby influencing molecular stability and clearance [120,121]. The consensus Log Po/w values [122,123] fall within the optimal range for drug-like molecules and vary depending on the substituent attached to the nitrogen atom of the isoxazolidine ring. In particular, compounds (411) bearing a methyl group show values close to 1.00, compounds with a tert-butyl substituent reach approximately 1.70, and analogs containing a phenyl ring exceed 2.00 (Table 4). This trend reflects the increasing hydrophobic character introduced by progressively bulkier substituents and suggests correspondingly enhanced membrane affinity, which may improve cellular uptake while simultaneously requiring careful consideration of metabolic stability and distribution. A similar substituent-dependent trend is observed for the aqueous solubility of the isoxazolidines (411). Their solubility varies substantially and is strongly influenced by the substituent attached to the nitrogen atom of the isoxazolidine ring. Compounds bearing a methyl group exhibit solubility values of approximately 1–2 mg/mL, derivatives containing a tert-butyl substituent reach around 0.3 mg/mL, and analogs with a phenyl group display markedly lower solubility of about 0.01–0.05 mg/mL (Table 4). This decrease in solubility correlates with the increasing hydrophobic surface area introduced by bulkier substituents, which promotes stronger interactions with lipid environments and reduces water compatibility, a factor relevant to absorption, distribution, and overall computationally predicted pharmacokinetic performance [111].
SwissADME predictions indicate that all isoxazolidines (411) display favorable gastrointestinal (GI) absorption (Table 4). High GI absorption suggests that the molecules can effectively traverse the intestinal epithelium and reach systemic circulation, which is crucial for achieving adequate oral bioavailability and therapeutic action [124,125]. The pharmacokinetic parameters summarized in Table 4 also show that each of the tested compounds (411) has low permeability across the blood–brain barrier (BBB). Substances that penetrate the BBB can influence central nervous system function, which is relevant for CNS-targeted therapies. However, limited BBB permeability is advantageous for molecules intended to act peripherally, as it minimizes the likelihood of unintended neurological effects [126]. Finally, the SwissADME predictions show that all isoxazolidines (411) display differentiated inhibitory patterns toward cytochrome P450 isoforms (Table 4).
The extent of CYP inhibition varies depending on the structural features of each compound, indicating that small changes in substituent type or steric environment can significantly influence interactions with metabolic enzymes.
The presented information is fully consistent with the bioavailability radars, and no issues arise apart from a moderate increase in unsaturation for the compounds bearing two phenyl substituents (411b), which still remain within the acceptable range (Figure 6). Moreover, it is worth noting that the predicted SwissADME properties of the resulting regioisomeric products remain essentially unchanged, with only minor differences observed in lipophilicity (Table 4). All of these observations should be understood as in silico predictions, providing guidance for future experimental studies rather than confirmed pharmacological outcomes.

2.2.2. PASS Computational Bioactivity Analysis

Due to the encouraging biological profiles of the isoxazolidine derivatives (411), their potential activity spectrum was explored using the PASS tool [55], accessed via the Way2Drug platform [107]. PASS employs a computational approach that predicts drug-like behavior, possible mechanisms of action, pharmacological effects, toxicity risks, and other features relevant at the early stages of drug development [127]. Although the method relies solely on 2D structural descriptors, it offers fast and reliable estimations that can support both further in silico analyses and future experimental work [128,129]. The output is presented as the probability of activity (Pa) relative to inactivity (Pi), with a compound considered potentially active when Pa exceeds Pi [55]. A Pa value above 0.700 denotes a strong likelihood of activity and makes the compound a good candidate for laboratory validation, whereas values between 0.500 and 0.700 indicate moderate confidence. Predictions below 0.500 are regarded as low-probability outcomes [55]. For all examined derivatives (411), the biological activities that fulfilled the Pa > Pi and Pa > 0.700 criteria are collected in Table 5.
The PASS prediction shows that the main biological potential of the studied isoxazolidines (411) is associated with modulation of the serotonin system, including antagonism of 5-HT2A and 5-HT2C receptors and stimulation of serotonin release, while the predicted CYP2H substrate activity reflects structural features relevant to metabolism but is of limited significance for human pharmacology (Table 5). In general, 5-HT2 antagonists can modulate neurotransmission in the central nervous system, influencing mood, cognition, and cardiovascular as well as metabolic processes [130,131]. Specifically, 5-HT2A antagonism is often associated with effects on perception and psychotropic activity [132], while 5-HT2C antagonism may regulate appetite, energy balance, and serotonergic signaling [133]. Stimulation of serotonin release suggests an additional mechanism to enhance serotonergic neurotransmission, potentially affecting mood and other physiological responses [134]. In contrast, the predicted CYP2H substrate activity indicates that these compounds possess structural features compatible with metabolism by P450 enzymes; however, as CYP2H is an avian isoform, this prediction has limited relevance for human pharmacology [135].
Based on the PASS prediction, two important in silico observations can be drawn. First, the predicted biological activity of the investigated isoxazolidine derivatives (411) is strongly influenced by the EWG present in the ethenyl moiety. In particular, for compounds where the EWG is a methoxycarbonyl group, the main predicted activity of these compounds (47) is predicted as substrates of CYP2H, an avian cytochrome P450 isoform (Table 5). In contrast, when the EWG is a cyano group, the predicted activities of derivatives (811) are primarily related to 5-HT2 antagonism or serotonin release stimulation (Table 5).
It should be emphasized that both the PASS and ADME analyses presented here are purely in silico predictions. They are intended to provide preliminary insights and suggest potential directions for future experimental studies rather than to assert confirmed biological activities. These results complement the mechanistic investigations and highlight possible functional trends, but their pharmacological relevance must be validated through comprehensive laboratory testing.

3. Materials and Methods

All computational work was performed with the GAUSSIAN 16 [136] program suite on the Ares cluster hosted by the AGH CYFRONET regional computing center in Cracow, Poland. The DFT computations were conducted with the hybrid ωb97xD functional [137], which combines long-range exchange treatment with an additional semiclassical London-dispersion correction. For all atoms, the 6-311G(d,p) basis set [138,139] was employed, providing d-polarization functions for second-row elements and p-polarization functions for hydrogen. The MEDT approach was employed as the general theoretical tool for examining the mechanisms of reactions involving organic and inorganic systems [140,141,142], with demonstrated applicability to cycloaddition processes [143,144]. All relevant stationary points were optimized in the gas phase at 298 K and 1 atm and subsequently verified by vibrational frequency calculations. The analyses showed that the initial reactants and the products represent true minima, as their Hessian matrices exhibit no imaginary frequencies.
The electronic characteristics of nitrones (1ac) and the ethenes (2 and 3) were examined using Conceptual Density Functional Theory (CDFT), following the guidelines proposed by Domingo [145,146,147]. The electrophilic Parr functions (Pk+) and nucleophilic Parr functions (Pk) were derived from the Atomic Spin Density (ASD) values of the corresponding radical ionic species [148,149].
Optimizations of the transition states (TSs) were carried out using the Berny method [150,151]. Frequency analyses confirmed the nature of the TSs by showing only one imaginary mode. To trace the connection between TSs and minimum-energy structures, intrinsic reaction coordinate (IRC) paths were computed [152] using the second-order González–Schlegel integration procedure [153,154]. Solvent effects of benzene were incorporated through full reoptimization of the gas-phase geometries at the same computational level, employing the polarizable continuum model (PCM) [155] within the self-consistent reaction field (SCRF) formalism [156,157].
The Global Electron Density Transfer (GEDT) [64] values were estimated by the Natural Population Analysis (NPA) [158,159] using the equation GEDT (f) = charge qf were q are the atoms of a framework (f) at the TSs. Indices of single bond development (l) were calculated according to the formula [160,161]:
l X Y = 1 r   X Y TS r   X Y P r   X Y P
where rTSX−Y is the distance between the reaction centers X and Y in the transition structure and rPX−Y is the same distance in the corresponding product.
The physicochemical and pharmacokinetic characteristics of the studied compounds were predicted using the SwissADME web platform [101]. This tool allowed for the assessment of absorption, distribution, metabolism, and excretion (ADME) parameters, offering a thorough overview of molecular behavior in a biological context. To evaluate the drug-likeness of the designed molecules, several widely recognized predictive models were applied, including the criteria formulated by Lipinski et al. [102], Ghose et al. [103], Veber et al. [104], Egan et al. [105], and Muegge et al. [106]. Together, these guidelines enabled a systematic analysis of molecular features that affect oral bioavailability and therapeutic potential. Additionally, the PASS online tool [107] was employed to estimate the likelihood of specific biological activities. Based on structure-activity relationship (SAR) modeling, this approach provided complementary insights into the potential pharmacological profiles of the isoxazolidine derivatives under investigation.
GaussView 6.0 software [162] was used to visualize molecular geometries of all the systems, as well as 3D representations of the HOMO and LUMO orbitals, MEP surfaces, and the radical anion and radical cation species. In turn, the bioavailability radar plots were generated using the SwissADME online platform [101].

4. Conclusions and Future Perspective

The ωB97X-D/6-311G(d,p) (PCM) quantum-chemical calculations allow for a comprehensive assessment of both the regio- and stereochemistry, as well as the molecular mechanism, of the (3+2) cycloaddition reactions between (Z)-C-phenyl-N-alkyl(phenyl)nitrones and (E)-3-(methylsulfonyl)-propenoic acid derivatives. The results indicate that the reactions proceed through a single-step mechanism involving highly synchronous formation of the C3–C4 and C5–O1 bonds. Pre-reaction orientational complexes are formed without an activation barrier and already adopt the spatial arrangement characteristic of the final products, emphasizing the role of molecular alignment in determining reaction outcomes. The reaction proceeds through nucleophilic attack of the nitrone oxygen center on the vinyl carbon of (E)-3-(methylsulfonyl)-propenoic acid derivatives. Both vinyl carbons of the ethene exhibit similar electrophilic character, which makes the prediction of regioselectivity challenging. Consequently, according to CDFT analysis, the regioselectivity of the process cannot be unambiguously determined, highlighting the subtle interplay of electronic factors in directing the cycloaddition outcome. Despite the expected polar character based on global and local CDFT descriptors, the transition states exhibit low polarity, as confirmed by GEDT values below 0.15 e, highlighting limited charge separation during bond formation.
Among the examined 32CA pathways, route A is predicted to be the kinetically most favorable pathway across all the reactions analyzed. The remaining pathways display higher activation barriers, indicating lower kinetic preference. Experimental results show partial correspondence with the computational predictions. The main limitations of the experimental data are due to the fragmentary nature of the reported results, the lack of consistent yield information, and the incomplete description of the remaining possible products, which prevents a clear assessment of their relative contributions. While the major product is consistent with route A, minor products likely corresponding to pathways C and D are also observed. This indicates that although route A is kinetically favored, contributions from alternative routes cannot be completely excluded, reflecting the subtle electronic and steric factors that complicate the regio- and stereoselectivity of these 32CA reactions. The comparison with the available experimental data underlines the relevance of our computational analysis while also highlighting the limitations inherent to the fragmentary experimental information.
As a future perspective, the analysis of the physicochemical parameters and predicted ADME profiles indicates that all isoxazolidine products formed in the studied reactions exhibit favorable drug-like characteristics. These compounds display properties consistent with good molecular stability, adequate solubility, and suitable permeability, suggesting their potential for effective absorption and distribution in a biological context. What is more, PASS predictions further indicate that the isoxazolidine derivatives exhibit pharmacological potential, particularly in modulation of the serotonin system, including antagonism of 5-HT2A and 5-HT2C receptors and stimulation of serotonin release. The nature of the EWG substituent significantly influences predicted biological activity, with methoxycarbonyl-substituted derivatives showing CYP2H substrate activity, whereas cyano-substituted compounds predominantly display 5-HT2 antagonism or serotonin release stimulation. These findings collectively underscore the interplay between structural features and biological potential, confirming the designed isoxazolidine derivatives as promising candidates for further pharmacological investigation. We hope that the inclusion of these analyses will make our manuscript a useful reference for researchers interested in both the synthetic and potential biological aspects of these compounds.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/molecules30244738/s1, Tables S1–S29: Thermochemistry and cartesian coordinates of molecules.

Author Contributions

Conceptualization, K.K. and R.J.; methodology, K.K. and R.J.; software, M.Z., K.K., V.D. and R.J.; validation, M.Z., K.K., V.D. and R.J.; formal analysis, M.Z., K.K., V.D. and R.J.; writing—original draft preparation, K.K. and R.J.; writing—review and editing, M.Z., K.K., V.D. and R.J. visualization, M.Z. and K.K.; supervision, R.J.; project administration, K.K. and R.J.; funding acquisition, K.K. and R.J. All authors have read and agreed to the published version of the manuscript.

Funding

We gratefully acknowledge Polish high-performance computing infrastructure PLGrid (HPC Center: ACK Cyfronet AGH) for providing computer facilities and support within computational grant no. PLG/2025/018201.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The original contributions presented in this study are included in the article/Supplementary Materials. Further inquiries can be directed to the corresponding authors.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Wang, X.; Hu, Q.; Tang, H.; Pan, X. Isoxazole/Isoxazoline Skeleton in the Structural Modification of Natural Products: A Review. Pharmaceuticals 2023, 16, 228. [Google Scholar] [CrossRef]
  2. Moncada-Basualto, M.; Saavedra-Olavarría, J.; Rivero-Jerez, P.S.; Rojas, C.; Maya, J.D.; Liempi, A.; Zúñiga-Bustos, M.; Olea-Azar, C.; Lapier, M.; Pérez, E.G.; et al. Assessment of the Activity of Nitroisoxazole Derivatives against Trypanosoma cruzi. Molecules 2024, 29, 2762. [Google Scholar] [CrossRef]
  3. Xiong, L.; Zhang, R.; Zhang, K.L.; Qi, B. Isoxazoline: A Privileged Scaffold for Agrochemical Discovery. J. Agric. Food Chem. 2025, 73, 14115–14128. [Google Scholar] [CrossRef]
  4. Kong, L.J.; Cao, X.Y.; Sun, N.B.; Min, L.J.; Duke, S.O.; Wu, H.K.; Zhang, L.Q.; Liu, X.H. Isoxazoline: An Emerging Scaffold in Pesticide Discovery. J. Agric. Food Chem. 2025, 73, 8678–8693. [Google Scholar] [CrossRef]
  5. Eneama, W.A.; Salman, H.H.; Mousa, M.N. Synthesis of a New Isoxazolidine and Evaluation Anticancer Activity against MCF 7 Breast Cancer Cell Line. Radiother. Oncol. 2023, 17, 1–8. [Google Scholar]
  6. Łysakowska, M.; Głowacka, I.E.; Honkisz-Orzechowska, E.; Handzlik, J.; Piotrowska, D.G. New 3-(Dibenzyloxyphosphoryl)isoxazolidine Conjugates of N1-Benzylated Quinazoline-2,4-diones as Potential Cytotoxic Agents against Cancer Cell Lines. Molecules 2024, 29, 3050. [Google Scholar] [CrossRef]
  7. Alshamari, A.; Al-Qudah, M.; Hamadeh, F.; Al-Momani, L.; Abu-Orabi, S. Synthesis, Antimicrobial and Antioxidant Activities of 2-Isoxazoline Derivatives. Molecules 2020, 25, 4271. [Google Scholar] [CrossRef]
  8. Sebbar, N.K.; Taha, M.L.; Ellouz, M.; Essassi, E.M.; Zerzouf, A.; Karrouchi, K.; Ouzidan, Y.; Zakaria, M.; Mague, J.T. Synthesis, DFT Study and Antibacterial Activity of some Isoxazoline Derivatives Containing 1,4-benzothiazin-3-one Nucleus Obtained Using 1,3-dipolar Cycloaddition Reaction. Iran. J. Chem. Chem. Eng. 2020, 39, 53–67. [Google Scholar] [CrossRef]
  9. Chiacchio, M.A.; Giofrè, S.V.; Romeo, R.; Romeo, G.; Chiacchio, U. Isoxazolidines as Biologically Active Compounds. Curr. Org. Synth. 2016, 13, 726–749. [Google Scholar] [CrossRef]
  10. Ouzounthanasis, K.A.; Rizos, S.R.; Koumbis, A.E. A Convenient Synthesis of Novel Isoxazolidine and Isoxazole Isoquinolinones Fused Hybrids. Molecules 2024, 29, 91. [Google Scholar] [CrossRef]
  11. Frederickson, M. Optically active isoxazolidines via asymmetric cycloaddition reactions of nitrones with alkenes: Applications in organic synthesis. Tetrahedron 1997, 53, 403–425. [Google Scholar] [CrossRef]
  12. Berthet, M.; Cheviet, T.; Dujardin, G.; Parrot, I.; Martinez, J. Isoxazolidine: A Privileged Scaffold for Organic and Medicinal Chemistry. Chem. Rev. 2016, 116, 15235–15283. [Google Scholar] [CrossRef]
  13. Kathiravan, M.K.; Salake, A.B.; Chothe, A.S.; Dudhe, P.B.; Watode, R.P.; Mukta, M.S.; Gadhwe, S. The Biology and Chemistry of Antifungal Agents: A Review. Bioorg. Med. Chem. 2012, 20, 5678–5698. [Google Scholar] [CrossRef]
  14. Stirling, C.J. Leaving groups and nucleofugality in elimination and other organic reactions. Acc. Chem. Res. 1979, 12, 198–203. [Google Scholar] [CrossRef]
  15. Kras, J.; Wróblewska, A.; Kącka-Zych, A. Unusual Regioselectivity in [3 + 2] Cycloaddition Reactions between (E)-3-Nitroacrylic Acid Derivatives and (Z)-C,N-Diphenylimine N-Oxide. Sci. Rad. 2023, 2, 112–117. [Google Scholar] [CrossRef]
  16. Jasiński, R. A new mechanistic insight on β-lactam systems formation from 5-nitroisoxazolidines. RSC Adv. 2015, 5, 50070–50072. [Google Scholar] [CrossRef]
  17. Padwa, A.; Koehler, K.F.; Rodriguez, A. New synthesis of β-lactams based on nitrone cycloaddition to nitroalkenes. J. Org. Chem. 1984, 49, 282–288. [Google Scholar] [CrossRef]
  18. Katsarou, A.; Stathopoulos, P.; Tzvetanova, I.D.; Asimotou, C.-M.; Falagas, M.E. β-Lactam/β-Lactamase Inhibitor Combination Antibiotics Under Development. Pathogens 2025, 14, 168. [Google Scholar] [CrossRef]
  19. Feng, Y.; De Vos, A.L.; Khan, S.; St. John, M.; Hasan, T. Quantitative Insights Into β-Lactamase Inhibitor’s Contribution in the Treatment of Carbapenemase-Producing Organisms with β-Lactams. Front. Microbiol. 2021, 12, 756410. [Google Scholar] [CrossRef]
  20. Zheng, T.; Nolan, E.M. Enterobactin-mediated delivery of beta-lactam antibiotics enhances antibacterial activity against pathogenic Escherichia coli. J. Am. Chem. Soc. 2014, 136, 9677–9691. [Google Scholar] [CrossRef]
  21. Jasiński, R. Recent Progress in the Synthesis of Nitroisoxazoles and Their Hydrogenated Analogs via [3 + 2] Cycloaddition Reactions (Microreview). Chem. Heterocycl. Compd. 2023, 59, 730–732. [Google Scholar] [CrossRef]
  22. Thakur, S.; Das, A.; Das, T. 1,3-Dipolar cycloaddition of nitrones: Synthesis of multisubstituted, diverse range of heterocyclic compounds. New J. Chem. 2021, 45, 11420–11456. [Google Scholar] [CrossRef]
  23. Black, D.S.C.; Crozier, R.F.; Davis, V.C. 1,3-Dipolar cycloaddition reactions of nitrones. Synthesis 1975, 1975, 205–221. [Google Scholar] [CrossRef]
  24. Domingo, L.R. Molecular Electron Density Theory: A Modern View of Reactivity in Organic Chemistry. Molecules 2016, 21, 1319. [Google Scholar] [CrossRef]
  25. Domingo, L.R.; Emamian, S.R. Understanding the mechanisms of [3 + 2] cycloaddition reactions. The pseudoradical versus the zwitterionic mechanism. Tetrahedron 2014, 70, 1267–1273. [Google Scholar] [CrossRef]
  26. Trost, B.M. The atom economy—A search for synthetic efficiency. Science 1991, 254, 1471–1477. [Google Scholar] [CrossRef] [PubMed]
  27. Soares, M.I.L.; Cardoso, A.L.; Pinho e Melo, T.M.V.D. Diels–Alder Cycloaddition Reactions in Sustainable Media. Molecules 2022, 27, 1304. [Google Scholar] [CrossRef]
  28. Jasiński, R.; Żmigrodzka, M.; Dresler, E.; Kula, K. A full regio- and stereoselective synthesis of 4-nitroisoxazolidines via stepwise [3 + 2] cycloaddition reactions between (Z)-C-(9-anthryl)-N-arylnitrones and (E)-3,3,3-trichloro-1-nitroprop-1-ene: Comprehensive experimental and theoretical study. J. Heterocycl. Chem. 2017, 54, 3314–3320. [Google Scholar] [CrossRef]
  29. Fryźlewicz, A.; Łapczuk-Krygier, A.; Kula, K.; Demchuk, O.M.; Dresler, E.; Jasiński, R. Regio- and Stereoselective Synthesis of Nitrofunctionalized 1,2-Oxazolidine Analogs of Nicotine. Chem. Heterocycl. Compd. 2020, 56, 120–122. [Google Scholar] [CrossRef]
  30. Kula, K.; Sadowski, M. Regio- and stereoselectivity of [3 + 2] cycloaddition reactions Between (Z)-C-(9-anthryl)-N-methylnitrone and analogues of trans-β-nitrostyrene in the light of MEDT computational study. Chem. Heterocycl. Compd. 2023, 59, 138–144. [Google Scholar] [CrossRef]
  31. Kowalska, E.; Sieroń, L.; Albrecht, A. Enantioselective, Decarboxylative (3 + 2)-Cycloaddition of Azomethine Ylides and Chromone-3-Carboxylic Acids. Molecules 2022, 27, 6809. [Google Scholar] [CrossRef]
  32. Hashimoto, T.; Maruoka, K. Recent Advances of Catalytic Asymmetric 1,3-Dipolar Cycloadditions. Chem. Rev. 2015, 115, 5366–5412. [Google Scholar] [CrossRef]
  33. Yang, X.; Cheng, F.; Kou, Y.-D.; Pang, S.; Shen, Y.-C.; Huang, Y.-Y.; Shibata, N. Catalytic Asymmetric 1,3-Dipolar Cycloaddition of β-Fluoroalkylated α,β-Unsaturated 2-Pyridylsulfones with Nitrones for Chiral Fluoroalkylated Isoxazolidines and γ-Amino Alcohols. Angew. Chem. Int. Ed. 2017, 56, 1510–1514. [Google Scholar] [CrossRef]
  34. Charmier, M.O.J.; Moussalli, N.; Chanet-Ray, J.; Chou, S. 1,3-dipolar cycloaddition reactions of nitrones with unsaturated methylsulfones and substituted crotonic esters. J. Chem. Res. Synop. 1991, 9, 566–567. [Google Scholar] [CrossRef]
  35. Woliński, P.; Kącka-Zych, A.; Wróblewska, A.; Wielgus, E.; Dolot, R.; Jasiński, R. Fully Selective Synthesis of Spirocyclic-1,2-oxazine N-Oxides via Non-Catalysed Hetero Diels-Alder Reactions with the Participation of Cyanofunctionalysed Conjugated Nitroalkenes. Molecules 2023, 28, 4586. [Google Scholar] [CrossRef] [PubMed]
  36. Woliński, P.; Kącka-Zych, A.; Demchuk, O.M.; Łapczuk-Krygier, A.; Mirosław, B.; Jasiński, R. Clean and molecularly programmable protocol for preparation of bis-heterobiarylic systems via a domino pseudocyclic reaction as a valuable alternative for TM-catalyzed cross-couplings. J. Clean. Prod. 2020, 275, 122086. [Google Scholar] [CrossRef]
  37. Zawadzińska-Wrochniak, K.; Kula, K.; Ríos-Gutiérrez, M.; Gostyński, B.; Krawczyk, T.; Jasiński, R. A Comprehensive Study of the Synthesis, Spectral Characteristics, Quantum–Chemical Molecular Electron Density Theory, and In Silico Future Perspective of Novel CBr3-Functionalyzed Nitro-2-Isoxazolines Obtained via (3 + 2) Cycloaddition of (E)-3,3,3-Tribromo-1-Nitroprop-1-ene. Molecules 2025, 30, 2149. [Google Scholar] [CrossRef]
  38. Huisgen, R. 1,3-Dipolar cycloadditions. 76. Concerted nature of 1,3-dipolar cycloadditions and the question of diradical intermediates. J. Org. Chem. 1976, 41, 403–419. [Google Scholar] [CrossRef]
  39. Huisgen, R. Kinetics and Mechanism of 1,3-Dipolar Cycloadditions. Angew. Chem. Int. Ed. 1963, 2, 633–696. [Google Scholar] [CrossRef]
  40. Huisgen, R. 1,3-dipolar cycloadditions. Past and future. Angew. Chem. Int. Ed. 1963, 2, 565–598. [Google Scholar] [CrossRef]
  41. Huisgen, R.; Pöchlauer, P.; Mlostoń, G.; Polborn, K. Reactions of Di(tert-butyl) diazomethane with Acceptor-Substituted Ethylenes. Helv. Chim. Acta 2007, 90, 983–998. [Google Scholar] [CrossRef]
  42. Wróblewska, A.; Sadowski, M.; Jasiński, R. Selectivity and molecular mechanism of the Au(III)-catalyzed [3 + 2] cycloaddition reaction between (Z)-C,N-diphenylnitrone and nitroethene in the light of the molecular electron density theory computational study. Chem. Heterocycl. Compd. 2024, 60, 639–645. [Google Scholar] [CrossRef]
  43. Dresler, E.; Woliński, P.; Wróblewska, A.; Jasiński, R. On the Question of Zwitterionic Intermediates in the [3 + 2] Cycloaddition Reactions between Aryl Azides and Ethyl Propiolate. Molecules 2023, 28, 8152. [Google Scholar] [CrossRef]
  44. 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]
  45. Mlostoń, G.; Urbaniak, K.; Linden, A.; Heimgartner, H. Selenophen-2-yl-substituted thiocarbonyl ylides—At the borderline of dipolar and biradical reactivity. Helv. Chim. Acta 2015, 98, 453–461. [Google Scholar] [CrossRef]
  46. Jasiński, R. A new insight on the molecular mechanism of the reaction between (Z)-C,N-diphenylnitrone and 1,2-bismethylene-3,3,4,4,5,5-hexamethylcyclopentane. J. Mol. Graph. Model. 2020, 94, 107461. [Google Scholar] [CrossRef]
  47. Fryźlewicz, A.; Kącka-Zych, A.; Demchuk, O.M.; Mirosław, B.; Woliński, P.; Jasiński, R. Green synthesis of nitrocyclopropane-type precursors of inhibitors for the maturation of fruits and vegetables via domino reactions of diazoalkanes with 2-nitroprop-1-ene. J. Clean. Prod. 2021, 292, 126079. [Google Scholar] [CrossRef]
  48. Ryachi, K.; Mohammad-Salim, H.; Al-Sadoon, M.K.; Zeroual, A.; de Julián-Ortiz, J.V.; El Idrissi, M.; Tounsi, A. Quantum study of the [3 + 2] cycloaddition of nitrile oxide and carvone oxime: Insights into toxicity, pharmacokinetics, and mechanism. Chem. Heterocycl. Compd. 2024, 60, 646–654. [Google Scholar] [CrossRef]
  49. Aitouna, A.O.; Syed, A.; Alfagham, A.T.; Mazoir, N.; de Julián-Ortiz, J.V.; Elgorban, A.M.; El Idrissi, M.; Wong, L.S.; Zeroual, A. Investigating the chemical reactivity and molecular docking of 2-diazo-3,3,3-trifluoro-1-nitropropane with phenyl methacrylate using computational methods. Chem. Heterocycl. Compd. 2024, 60, 592–599. [Google Scholar] [CrossRef]
  50. Abdoul-Hakim, M.; Kenzy, C.; Subramaniam, M.; Zeroual, A.; Syed, A.; Bahkali, A.H.; Verma, M.; Wang, S.; Garmes, H. Elucidating Chemoselectivity and Unraveling the Mechanism of 1,3-Dipolar Cycloaddition between Diphenyl Nitrilimine and (Isoxazol-3-yl)methylbenzimidazole through Molecular Electron Density Theory. Chem. Heterocycl. Compd. 2024, 60, 617–626. [Google Scholar] [CrossRef]
  51. Chattaraj, P.K.; Roy, D.R. Update 1 of: Electrophilicity Index. Chem. Rev. 2007, 107, PR46–PR74. [Google Scholar] [CrossRef]
  52. Domingo, L.R. 1999–2024, a quarter century of the Parr’s electrophilicity ω index. Sci. Radices 2024, 3, 157–186. [Google Scholar] [CrossRef]
  53. Domingo, L.R.; Ríos-Gutiérrez, M.; Pérez, P. Electrophilicity w and Nucleophilicity N Scales for Cationic and Anionic Species. Sci. Radices 2025, 4, 1–17. [Google Scholar] [CrossRef]
  54. Di, L.; Kerns, E. Drug-like Properties: Concepts, Structure Design and Methods from ADME to Toxicity Optimization; Academic Press: Cambridge, MA, USA, 2015. [Google Scholar]
  55. Filimonov, D.A.; Lagunin, A.A.; Gloriozova, T.A.; Rudik, A.V.; Druzhilovskii, D.S.; Pogodin, P.V.; Poroikov, V.V. Prediction of the biological activity spectra of organic compounds using the PASS online web resource. Chem. Heterocycl. Compd. 2014, 50, 444–457. [Google Scholar] [CrossRef]
  56. Sadowski, M.; Kula, K. Unexpected Course of Reaction Between (1E,3E)-1,4-Dinitro-1,3-butadiene and N-Methyl Azomethine Ylide—A Comprehensive Experimental and Quantum-Chemical Study. Molecules 2024, 29, 5066. [Google Scholar] [CrossRef] [PubMed]
  57. Kula, K.; Jasiński, R. On the Question of the Application Potential and the Molecular Mechanism of the Formation of 1,3-Diaryl-5-Nitropyrazoles from Trichloromethylated Diarylnitropyrazolines. Molecules 2025, 30, 4306. [Google Scholar] [CrossRef]
  58. Parr, R.G.; von Szentpaly, L.; Liu, S. Electrophilicity Index. J. Am. Chem. Soc. 1999, 121, 1922–1924. [Google Scholar] [CrossRef]
  59. Domingo, L.R.; Pérez, P. The Nucleophilicity N Index in Organic Chemistry. Org. Biomol. Chem. 2011, 9, 7168–7175. [Google Scholar] [CrossRef]
  60. Aurell, M.J.; Domingo, L.R.; Pérez, P.; Contreras, R. A Theoretical Study on the Regioselectivity Of 1,3-Dipolar Cycloadditions Using DFT-based Reactivity Indexes. Tetrahedron 2004, 60, 11503–11509. [Google Scholar] [CrossRef]
  61. Fryźlewicz, A.; Olszewska, A.; Zawadzińska, K.; Woliński, P.; Kula, K.; Kącka-Zych, A.; Łapczuk-Krygier, A.; Jasiński, R. On the Mechanism of the Synthesis of Nitrofunctionalised Δ2-Pyrazolines via [3 + 2] Cycloaddition Reactions between α-EWG-Activated Nitroethenes and Nitrylimine TAC Systems. Organics 2022, 3, 59–76. [Google Scholar] [CrossRef]
  62. Domingo, L.R.; Aurell, M.J.; Pérez, P.; Contreras, R. Quantitative characterization of the global electrophilicity power of common diene/dienophile pairs in Diels–Alder reactions. Tetrahedron 2002, 58, 4417–4423. [Google Scholar] [CrossRef]
  63. Kula, K.; Jasiński, R. Synthesis of bis(het)aryl systems via domino reaction involving (2E,4E)-2,5-dinitrohexa-2,4-diene: DFT mechanistic considerations. Chem. Heterocycl. Comp. 2024, 60, 600–610. [Google Scholar] [CrossRef]
  64. Domingo, L.R.; Ríos-Gutiérrez, M. A Useful Classification of Organic Reactions Based on the Flux of the Electron Density. Sci. Radices 2023, 2, 1–24. [Google Scholar] [CrossRef]
  65. Domingo, L.R.; Pérez, P.; Sáez, J.A. Understanding the Local Reactivity in Polar Organic Reactions Through Electrophilic and Nucleophilic Parr Functions. RSC Adv. 2013, 3, 1486–1494. [Google Scholar] [CrossRef]
  66. Zawadzińska, K.; Kula, K. Application of β-phosphorylated nitroethenes in [3 + 2] cycloaddition reactions involving benzonitrile N-oxide in the light of DFT computational study. Organics 2021, 2, 26–37. [Google Scholar] [CrossRef]
  67. Messaadia, S.; Nacereddine, A.K.; Djerourou, A. Exploring the factors controlling the mechanism and the high stereoselectivity of the polar [3 + 2] cycloaddition reaction of the N,N’-cyclic azomethine imine with 3-nitro-2-phenyl-2H-chromene. A Molecular Electron Density Theory study. Chem. Heterocycl. Compd. 2023, 59, 128–137. [Google Scholar] [CrossRef]
  68. Kula, K.; Kącka-Zych, A.; Łapczuk-Krygier, A.; Jasiński, R. Analysis of the Possibility and Molecular Mechanism of Carbon Dioxide Consumption in the Diels-Alder Processes. Pure Appl. Chem. 2021, 93, 427–446. [Google Scholar] [CrossRef]
  69. Kula, K.; Kącka-Zych, A.; Łapczuk-Krygier, A.; Wzorek, Z.; Nowak, A.K.; Jasiński, R. Experimental and Theoretical Mechanistic Study on the Thermal Decomposition of 3,3-Diphenyl-4-(Trichloromethyl)-5-Nitropyrazoline. Molecules 2021, 26, 1364. [Google Scholar] [CrossRef]
  70. Chafaa, F.; Nacereddine, A.K. A molecular electron density theory study of mechanism and selectivity of the intramolecular [3 + 2] cycloaddition reaction of a nitrone–vinylphosphonate adduct. Chem. Heterocycl. Compd. 2023, 59, 171–178. [Google Scholar] [CrossRef]
  71. Sadowski, M.; Kula, K.; Jasiński, R. On the question of the transformation of amino acids into aldimine N-oxides (nitrones): Molecular Electron Density Theory (MEDT) considerations. Monatsh. Chem. 2025, in press. [Google Scholar] [CrossRef]
  72. Dresler, E.; Wróblewska, A.; Jasiński, R. Energetic Aspects and Molecular Mechanism of 3-Nitro-substituted 2-Isoxazolines Formation via Nitrile N-Oxide [3 + 2] Cycloaddition: An MEDT Computational Study. Molecules 2024, 29, 3042. [Google Scholar] [CrossRef]
  73. Mondal, A.; Acharjee, N. Unveiling the exclusive stereo and site selectivity in [3 + 2] cycloaddition reactions of a tricyclic strained alkene with nitrile oxides from the molecular electron density theory perspective. Chem. Heterocycl. Compd. 2023, 59, 145–154. [Google Scholar] [CrossRef]
  74. Yousfi, Y.; Benchouk, W.; Mekelleche, S.M. Prediction of the regioselectivity of the ruthenium-catalyzed [3 + 2] cycloadditions of benzyl azide with internal alkynes using conceptual DFT indices of reactivity. Chem. Heterocycl. Compd. 2023, 59, 118–127. [Google Scholar] [CrossRef]
  75. Jasiński, R. Nitroacetylene as dipolarophile in [2 + 3] cycloaddition reactions with allenyl-type three-atom components: DFT computational study. Monatshefte Für Chem. 2015, 146, 591–599. [Google Scholar] [CrossRef]
  76. Kula, K.; Dobosz, J.; Jasiński, R.; Kącka-Zych, A.; Łapczuk-Krygier, A.; Mirosław, B.; Demchuk, O.M. [3 + 2] Cycloaddition of diaryldiazomethanes with (E)-3,3,3-trichloro-1-nitroprop-1-ene: An experimental, theoretical and structural study. J. Mol. Struct. 2020, 1203, 127473. [Google Scholar] [CrossRef]
  77. Zaki, H.; Ouabane, M.; Aissaoui, S.; Alaqarbeh, M.; Bouachrine, M. 1,2,4-triazole-chalcone and derivatives as antiproliferative agents: Quantum chemical studies, molecular docking, ADME-Tox and MD simulation. Curr. Chem. Lett. 2024, 14, 867–884. [Google Scholar] [CrossRef]
  78. Mohtat, B.; Siadati, S.A.; Khalilzadeh, M.A. Understanding the mechanism of the 1,3-dipolar cycloaddition reaction between a thioformaldehyde S-oxide and cyclobutadiene: Competition between the stepwise and concerted routes. Prog. React. Kinet. Mec. 2019, 44, 213–221. [Google Scholar] [CrossRef]
  79. Kula, K.; Łapczuk, A.; Sadowski, M.; Kras, J.; Zawadzińska, K.; Demchuk, O.M.; Gaurav, G.K.; Wróblewska, A.; Jasiński, R. On the Question of the Formation of Nitro-Functionalized 2,4-Pyrazole Analogs on the Basis of Nitrylimine Molecular Systems and 3,3,3-Trichloro-1-Nitroprop-1-Ene. Molecules 2022, 27, 8409. [Google Scholar] [CrossRef]
  80. Kącka-Zych, A. The Molecular Mechanism of the Formation of Four-Membered Cyclic Nitronates and Their Retro (3 + 2) Cycloaddition: A DFT Mechanistic Study. Molecules 2021, 26, 4786. [Google Scholar] [CrossRef]
  81. Gillies, C.W.; Gillies, J.Z.; Suenram, D.R.; Lovas, F.J.; Kraka, E.; Cremer, D. Van der Waals complexes in 1,3-dipolar cycloaddition reactions: Ozone-ethylene. J. Am. Chem. Soc. 1991, 113, 2412–2421. [Google Scholar] [CrossRef]
  82. Gillies, J.Z.; Gillies, C.W.; Lovas, F.J.; Matsumura, K.; Suenram, R.D.; Kraka, E.; Cremer, D. Van der Waals complexes of chemically reactive gases: Ozone-acetylene. J. Am. Chem. Soc. 1991, 113, 6408–6415. [Google Scholar] [CrossRef]
  83. Cremer, D.; Kraka, E.; Crehuet, R.; Anglada, J.; Gräfenstein, J. The Ozone–Acetylene Reaction: Concerted Or Non-Concerted Reaction Mechanism? A Quantum Chemical Investigation. Chem. Phys. Lett. 2001, 347, 268–276. [Google Scholar] [CrossRef]
  84. Jodłowski, P.J.; Kurowski, G.; Dymek, K.; Oszajca, M.; Piskorz, W.; Hyjek, K.; Wach, A.; Pajdak, A.; Mazur, M.; Rainrer, D.N.; et al. From crystal phase mixture to pure metal-organic frameworks–Tuning pore and structure properties. Ultrason. Sonochem. 2023, 95, 106377. [Google Scholar] [CrossRef]
  85. Ronduda, H.; Zybert, M.; Patkowski, W.; Ostrowski, A.; Jodłowski, P.; Szymański, D.; Kępiński, L.; Raróg-Pilecka, W. A high performance barium-promoted cobalt catalyst supported on magnesium–lanthanum mixed oxide for ammonia synthesis. RSC Adv. 2021, 11, 14218–14228. [Google Scholar] [CrossRef]
  86. Jodlowski, P.; Chlebda, D.; Piwowarczyk, E.; Chrzan, M.; Jędrzejczyk, R.; Sitarz, M.; Węgrzynowicz, A.; Kołodziej, A.; Łojewska, J. In situ and operando spectroscopic studies of sonically aided catalysts for biogas exhaust abatement. J. Mol. Struct. 2016, 1126, 132–140. [Google Scholar] [CrossRef]
  87. Karaś, A.; Łapczuk, A. Computational Model of the Formation of Novel Nitronorbornene Analogs via Diels–Alder Process. React. Kinet. Mech. Catal. 2025, 138, 2671–2689. [Google Scholar] [CrossRef]
  88. Kącka-Zych, A.; Zeroual, A.; Syed, A.; Bahkali, A.H. Docking Survey, ADME, Toxicological Insights, and Mechanistic Exploration of the Diels-Alder Reaction between Hexachlorocyclopentadiene and Dichloroethylene. J. Comput. Chem. 2025, 46, 70092. [Google Scholar] [CrossRef]
  89. Fałowska, A.; Grzybowski, S.; Kapuściński, D.; Sambora, K.; Łapczuk, A. Modeling of the General Trends of Reactivity and Regioselectivity in Cyclopentadiene–Nitroalkene Diels–Alder Reactions. Molecules 2025, 30, 2467. [Google Scholar] [CrossRef]
  90. Salih, S.A.M.; Basheer, H.A.; de Julián-Ortiz, J.V.; Mohammad-Salim, H.A. Unveiling the Stereoselectivity and Regioselectivity of the [3 + 2] Cycloaddition Reaction between N-methyl-C-4-methylphenyl-nitrone and 2-Propynamide from a MEDT Perspective. Int. J. Mol. Sci. 2023, 24, 9102. [Google Scholar] [CrossRef]
  91. Mohammad-Salim, H.A.; Acharjee, N.; Abdallah, H. Insights into the mechanism and regioselectivity of the [3 + 2] cycloaddition reactions of cyclic nitrone to nitrile functions with a molecular electron density theory perspective. Theor. Chem. Acc. 2021, 140, 1. [Google Scholar] [CrossRef]
  92. Domingo, L.R.; Chamorro, E.; Pérez, P. Understanding the mechanism of non-polar Diels-Alder reactions. A comparative ELF analysis of concerted and stepwise diradical mechanisms. Org. Biomol. Chem. 2010, 8, 5495–5504. [Google Scholar] [CrossRef] [PubMed]
  93. Cao, X.; You, J.; Wang, Y.; Yu, Y.; Wu, W.; Liang, Y. Synthesis of New Isoxazolidine Derivatives Utilizing the Functionality of N-Carbonylpyrazol-Linked Isoxazolidines. Molecules 2024, 29, 3454. [Google Scholar] [CrossRef]
  94. Moi, D.; Vittorio, S.; Angeli, A.; Balboni, G.; Supuran, C.T.; Onnis, V. Investigation on Hydrazonobenzenesulfonamides as Human Carbonic Anhydrase I, II, IX and XII Inhibitors. Molecules 2023, 28, 91. [Google Scholar] [CrossRef]
  95. Łapczuk, A. Zwitterionic Pathway in the Diels–Alder Reaction: Solvent and Substituent Effects from ωB97XD/6-311G(d) Calculations. Molecules 2025, 30, 4710. [Google Scholar] [CrossRef]
  96. Alhawday, F.; Alminderej, F.; Ghannay, S.; Hammami, B.; Albadri, A.E.A.E.; Kadri, A.; Aouadi, K. In Silico Design, Synthesis, and Evaluation of Novel Enantiopure Isoxazolidines as Promising Dual Inhibitors of α-Amylase and α-Glucosidase. Molecules 2024, 29, 305. [Google Scholar] [CrossRef]
  97. Alminderej, F.; Ghannay, S.; Elsamani, M.O.; Alhawday, F.; Albadri, A.E.A.E.; Elbehairi, S.E.I.; Alfaifi, M.Y.; Kadri, A.; Aouadi, K. In Vitro and In Silico Evaluation of Antiproliferative Activity of New Isoxazolidine Derivatives Targeting EGFR: Design, Synthesis, Cell Cycle Analysis, and Apoptotic Inducers. Pharmaceuticals 2023, 16, 1025. [Google Scholar] [CrossRef]
  98. Karcev, D.D.; Efremova, M.M.; Molchanov, A.P.; Rostovskii, N.V.; Kryukova, M.A.; Bunev, A.S.; Khochenkov, D.A. Selective and Reversible 1,3-Dipolar Cycloaddition of 2-(2-Oxoindoline-3-ylidene)acetates with Nitrones in the Synthesis of Functionalized Spiroisoxazolidines. Int. J. Mol. Sci. 2022, 23, 12639. [Google Scholar] [CrossRef]
  99. Ravi Kumar, K.R.; Mallesha, H.; Rangappa, K.S. Synthesis of Novel Isoxazolidine Derivatives and Their Antifungal and Antibacterial Properties. Arch. Pharm. 2003, 336, 159–164. [Google Scholar] [CrossRef]
  100. Jasiński, R. Recent progress in the synthesis of nitropyrazoles and their hydrogenated analogs via non-catalysed [3+2] cycloaddition reactions of conjugated nitroalkenes. Chem. Heterocyclic Compd. 2025, 61, 286–289. [Google Scholar] [CrossRef]
  101. SwissADME. Swiss Institute of Bioinformatics. Available online: http://www.swissadme.ch/ (accessed on 12 November 2025).
  102. Lipinski, C.A.; Lombardo, F.; Dominy, B.W.; Feeney, P.J. Experimental and computational approaches to estimate solubility and permeability in drug discovery and development settings. Adv. Drug Deliv. Rev. 2001, 46, 3–26. [Google Scholar] [CrossRef] [PubMed]
  103. Ghose, A.K.; Viswanadhan, V.N.; Wendoloski, J.J. A knowledge-based approach in designing combinatorial or medicinal chemistry libraries for drug discovery. 1. A qualitative and quantitative characterization of known drug databases. J. Comb. Chem. 1999, 1, 55–68. [Google Scholar] [CrossRef]
  104. Veber, D.F.; Johnson, S.R.; Cheng, H.-Y.; Smith, B.R.; Ward, K.W.; Kopple, K.D. Molecular properties that influence the oral bioavailability of drug candidates. J. Med. Chem. 2002, 45, 2615–2623. [Google Scholar] [CrossRef] [PubMed]
  105. Egan, W.J.; Merz, K.M.; Baldwin, J.J. Prediction of drug absorption using multivariate statistics. J. Med. Chem. 2000, 43, 3867–3877. [Google Scholar] [CrossRef]
  106. Muegge, I.; Heald, S.L.; Brittelli, D. Simple selection criteria for drug-like chemical matter. J. Med. Chem. 2001, 44, 1841–1846. [Google Scholar] [CrossRef]
  107. Way2Drug, PASS Online. Available online: http://www.way2drug.com/passonline/ (accessed on 16 November 2025).
  108. Fourches, D.; Muratov, E.; Tropsha, A. Trust, but verify: On the importance of chemical structure curation in cheminformatics and QSAR modeling research. J. Chem. Inf. Model. 2010, 50, 1189–1204. [Google Scholar] [CrossRef]
  109. Bosc, N.; Atkinson, F.; Felix, E.; Gaulton, A.; Hersey, A.; Leach, A.R. Large scale comparison of QSAR and conformal prediction methods and their applications in drug discovery. J. Cheminform. 2019, 11, 4. [Google Scholar] [CrossRef]
  110. Allam, T.; Balderston, D.E.; Chahal, M.K.; Hilton, K.L.F.; Hind, C.K.; Keers, O.B.; Lilley, R.J.; Manwani, C.; Overton, A.; Popoola, P.I.A.; et al. Tools to Enable the Study and Translation of Supramolecular Amphiphiles. Chem. Soc. Rev. 2023, 52, 6892–6917. [Google Scholar] [CrossRef]
  111. Daina, A.; Michielin, O.; Zoete, V. SwissADME: A Free Web Tool to Evaluate Pharmacokinetics, Drug-Likeness and Medicinal Chemistry Friendliness of Small Molecules. Sci. Rep. 2017, 7, 42717. [Google Scholar] [CrossRef]
  112. Dulsat, J.; López-Nieto, B.; Estrada-Tejedor, R.; Borrell, J.I. Evaluation of Free Online ADMET Tools for Academic or Small Biotech Environments. Molecules 2023, 28, 776. [Google Scholar] [CrossRef] [PubMed]
  113. Sumontri, S.; Eiamart, W.; Tadtong, S.; Samee, W. Utilizing ADMET Analysis and Molecular Docking to Elucidate the Neuroprotective Mechanisms of a Cannabis-Containing Herbal Remedy (Suk-Saiyasna) in Inhibiting Acetylcholinesterase. Int. J. Mol. Sci. 2025, 26, 3189. [Google Scholar] [CrossRef] [PubMed]
  114. Ertl, P.; Rohde, B.; Selzer, P. Fast Calculation of Molecular Polar Surface Area as a Sum of Fragment-Based Contributions and Its Application to the Prediction of Drug Transport Properties. J. Med. Chem. 2000, 43, 3714–3717. [Google Scholar] [CrossRef] [PubMed]
  115. Mauri, A.; Bertola, M. Alvascience: A New Software Suite for the QSAR Workflow Applied to the Blood–Brain Barrier Permeability. Int. J. Mol. Sci. 2022, 23, 12882. [Google Scholar] [CrossRef]
  116. Lopez-Mercado, S.; Enríquez, C.; Valderrama, J.A.; Pino-Rios, R.; Ruiz-Vásquez, L.; Ruiz Mesia, L.; Vargas-Arana, G.; Buc Calderon, P.; Benites, J. Exploring the Antibacterial and Antiparasitic Activity of Phenylaminonaphthoquinones—Green Synthesis, Biological Evaluation and Computational Study. Int. J. Mol. Sci. 2024, 25, 10670. [Google Scholar] [CrossRef]
  117. Hoekman, D. Exploring QSAR Fundamentals and Applications in Chemistry and Biology, Volume 1. Hydrophobic, Electronic and Steric Constants. J. Am. Chem. Soc. 1996, 118, 10678. [Google Scholar] [CrossRef]
  118. Delaney, J.S. ESOL: Estimating Aqueous Solubility Directly from Molecular Structure. J. Chem. Inf. Comput. Sci. 2004, 44, 1000–1005. [Google Scholar] [CrossRef]
  119. Ali, J.; Camilleri, P.; Brown, M.B.; Hutt, A.J.; Kirton, S.B. In Silico Prediction of Aqueous Solubility Using Simple QSPR Models: The Importance of Phenol and Phenol-like Moieties. J. Chem. Inf. Model. 2012, 52, 2950–2957. [Google Scholar] [CrossRef]
  120. Arnott, J.A.; Planey, S.L. The influence of lipophilicity in drug discovery and design. Expert Opin. Drug Discov. 2012, 7, 863–875. [Google Scholar] [CrossRef]
  121. Savjani, K.T.; Gajjar, A.K.; Savjani, J.K. Drug solubility: Importance and enhancement techniques. ISRN Pharm. 2012, 2012, 195727. [Google Scholar] [CrossRef]
  122. Mannhold, R.; Poda, G.I.; Ostermann, C.; Tetko, I.V. Calculation of molecular lipophilicity: State-of-the-art and comparison of log P methods on more than 96,000 compounds. J. Pharm. Sci. 2009, 98, 861–893. [Google Scholar] [CrossRef] [PubMed]
  123. He, Z.; Yang, D.; Fan, X.; Zhang, M.; Li, Y.; Gu, X.; Yang, M. The Roles and Mechanisms of lncRNAs in Liver Fibrosis. Int. J. Mol. Sci. 2020, 21, 1482. [Google Scholar] [CrossRef] [PubMed]
  124. Nicze, M.; Borówka, M.; Dec, A.; Niemiec, A.; Bułdak, Ł.; Okopień, B. The Current and Promising Oral Delivery Methods for Protein- and Peptide-Based Drugs. Int. J. Mol. Sci. 2024, 25, 815. [Google Scholar] [CrossRef]
  125. Wu, K.; Kwon, S.H.; Zhou, X.; Fuller, C.; Wang, X.; Vadgama, J.; Wu, Y. Overcoming Challenges in Small-Molecule Drug Bioavailability: A Review of Key Factors and Approaches. Int. J. Mol. Sci. 2024, 25, 13121. [Google Scholar] [CrossRef]
  126. Muller, T.; Demizieux, L.; Troy-Fioramonti, S.; Buch, C.; Leemput, J.; Belloir, C.; Pais de Barros, J.-P.; Jourdan, T.; Passilly-Degrace, P.; Fioramonti, X.; et al. Chemical Synthesis, Pharmacokinetic Properties and Biological Effects of JM-00266, a Putative Non-Brain Penetrant Cannabinoid Receptor 1 Inverse Agonist. Int. J. Mol. Sci. 2022, 23, 2923. [Google Scholar] [CrossRef] [PubMed]
  127. Filimonov, D.A.; Rudik, A.V.; Dmitriev, A.V.; Poroikov, V.V. Computer-Aided Estimation of Biological Activity Profiles of Drug-Like Compounds Taking into Account Their Metabolism in Human Body. Int. J. Mol. Sci. 2020, 21, 7492. [Google Scholar] [CrossRef] [PubMed]
  128. Kula, K.; Kuś, E. In Silico Study About Substituent Effects, Electronic Properties, and the Biological Potential of 1,3-Butadiene Analogues. Int. J. Mol. Sci. 2025, 26, 8983. [Google Scholar] [CrossRef]
  129. Sadowski, M.; Synkiewicz-Musialska, B.; Kula, K. (1E,3E)-1,4-Dinitro-1,3-butadiene—Synthesis, Spectral Characteristics and Computational Study Based on MEDT, ADME and PASS Simulation. Molecules 2024, 29, 542. [Google Scholar] [CrossRef]
  130. Furge, L.L.; Guengerich, F.P. Cytochrome P450 Enzymes in Drug Metabolism and Chemical Toxicology: An Introduction. Biochem. Mol. Biol. Educ. 2006, 34, 66–74. [Google Scholar] [CrossRef]
  131. Guengerich, F.P. Cytochrome P450 and Chemical Toxicology. Chem. Res. Toxicol. 2008, 21, 70–83. [Google Scholar] [CrossRef]
  132. Zhang, G.; Stackman, R. The Role of Serotonin 5-HT 2A Receptors in Memory and Cognition. Front. Pharmacol. 2015, 6, 225. [Google Scholar] [CrossRef]
  133. Halford, J.C.; Harrold, J.A. 5-HT(2C) receptor agonists and the control of appetite. Handb. Exp. Pharmacol. 2012, 209, 349–356. [Google Scholar] [CrossRef]
  134. Ciranna, L.; Costa, L. Therapeutic Effects of Pharmacological Modulation of Serotonin Brain System in Human Patients and Animal Models of Fragile X Syndrome. Int. J. Mol. Sci. 2025, 26, 2495. [Google Scholar] [CrossRef]
  135. Rendic, S. Summary of information on human CYP enzymes: Human P450 metabolism data. Drug Metab. Rev. 2002, 34, 83–448. [Google Scholar] [CrossRef] [PubMed]
  136. Frisch, M.J.; Trucks, G.W.; Schlegel, H.B.; Scuseria, G.E.; Robb, M.A.; Cheeseman, J.R.; Scalmani, G.; Barone, V.; Petersson, G.A.; Nakatsuji, H.; et al. Gaussian 16, Revision C.01; Gaussian, Inc.: Wallingford, CT, USA, 2016. [Google Scholar]
  137. Chai, J.-D.; Head-Gordon, M. Long-range corrected hybrid density functionals with damped atom-atom dispersion corrections. Phys. Chem. Chem. Phys. 2008, 10, 6615–6620. [Google Scholar] [CrossRef] [PubMed]
  138. Hehre, W.J.; Radom, L.; Schleyer, P.v.R.; Pople, J.A. Ab Initio Molecular Orbital Theory; Wiley: New York, NY, USA, 1986. [Google Scholar] [CrossRef]
  139. Krishnan, R.; Binkley, J.S.; Seeger, R.; Pople, J.A. Self-consistent molecular orbital methods. XX. A basis set for correlated wave functions. J. Chem. Phys. 1980, 72, 650–654. [Google Scholar] [CrossRef]
  140. Łapczuk-Krygier, A.; Kazimierczuk, K.; Pikies, J.; Ríos-Gutiérrez, M. A Comprehensive Experimental and Theoretical Study on the [{(η5-C5H5)2Zr[P(µ-PNEt2)2P(NEt2)2P]}]2O Crystalline System. Molecules 2021, 26, 7282. [Google Scholar] [CrossRef]
  141. Abbasi, M.; Farnia, S.M.F.; Tahghighi, A. The possibility of applying some heteroatom-decorated g-C3N4 heterocyclic nanosheets for delivering 5-aminosalicylic acid anti-inflammatory agent. Chem. Heterocycl. Comp. 2024, 60, 655–662. [Google Scholar] [CrossRef]
  142. Demchuk, O.M.; Jasiński, R.; Strzelecka, D.; Dziuba, K.; Kula, K.; Chrzanowski, J.; Krasowska, D. A clean and simple method for deprotection of phosphines from borane complexes. Pure Appl. Chem. 2018, 90, 49–62. [Google Scholar] [CrossRef]
  143. Kącka-Zych, A. Understanding of the stability of acyclic nitronic acids in the light of molecular electron density theory. J. Mol. Graph. 2024, 129, 108754. [Google Scholar] [CrossRef]
  144. Kapuściński, D.; Łapczuk, A. Bicyclic 1, 2-oxazine 2-oxides as the discrete intermediates in the reactions between regioisomeric trimetylsilylcyclopentadienes and (2E)-2-nitro-3-phenylprop-2-enenitrile. Chem. Heterocycl. Comp. 2025, 61, 372–380. [Google Scholar] [CrossRef]
  145. Domingo, L.R.; Ríos-Gutiérrez, M.; Pérez, P. Applications of the Conceptual Density Functional Theory Indices to Organic Chemistry Reactivity. Molecules 2016, 21, 748. [Google Scholar] [CrossRef]
  146. Parr, R.G.; Gadre, S.R.; Bartolotti, L.J. Local Density Functional Theory of Atoms and Molecules. Proc. Natl. Acad. Sci. USA 1979, 76, 2522–2526. [Google Scholar] [CrossRef]
  147. Domingo, L.R. A new C–C bond formation model based on the quantum chemical topology of electron density. RSC Adv. 2014, 4, 32415–32428. [Google Scholar] [CrossRef]
  148. Chamorro, E.; Pérez, P.; Domingo, L.R. On the nature of Parr functions to predict the most reactive sites along organic polar reactions. Chem. Phys. Lett. 2013, 582, 141–143. [Google Scholar] [CrossRef]
  149. Chattaraj, P.K.; Duley, S.; Domingo, L.R. Understanding local electrophilicity/nucleophilicity activation through a single reactivity difference index. Org. Biomol. Chem. 2012, 10, 2855–2861. [Google Scholar] [CrossRef]
  150. Schlegel, H.B. Optimization of equilibrium geometries and transition structures. J. Comput. Chem. 1982, 3, 214–218. [Google Scholar] [CrossRef]
  151. Schlegel, H.B. Geometry Optimization on Potential Energy Surfaces. In Modern Electronic Structure Theory; Yarkony, D.R., Ed.; World Scientific Publishing: Singapore, 1994; pp. 459–501. [Google Scholar] [CrossRef]
  152. Fukui, K. Formulation of the reaction coordinate. J. Phys. Chem. 1970, 74, 4161–4163. [Google Scholar] [CrossRef]
  153. Gonzalez, C.; Schlegel, H.B. Reaction path following in mass-weighted internal coordinates. J. Phys. Chem. 1990, 94, 5523–5527. [Google Scholar] [CrossRef]
  154. Gonzalez, C.; Schlegel, H.B. Improved algorithms for reaction path following: Higher order implicit algorithms. J. Chem. Phys. 1991, 95, 5853–5860. [Google Scholar] [CrossRef]
  155. Tomasi, J.; Persico, M. Molecular Interactions in Solution: An Overview of Methods Based on Continuous Distributions of the Solvent. Chem. Rev. 1994, 94, 2027–2094. [Google Scholar] [CrossRef]
  156. Cossi, M.; Barone, V.; Cammi, R.; Tomasi, J. Ab initio study of solvated molecules: A new implementation of the polarizable continuum model. Chem. Phys. Chem. 1996, 225, 327–335. [Google Scholar] [CrossRef]
  157. Barone, V.; Cossi, M.; Tomasi, J. Geometry optimization of molecular structures in solution by the polarizable continuum model. J. Comput. Chem. 1998, 19, 404–417. [Google Scholar] [CrossRef]
  158. Reed, A.E.; Weinstock, R.B.; Weinhold, F. Natural population analysis. J. Chem. Phys. 1985, 83, 735–746. [Google Scholar] [CrossRef]
  159. 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]
  160. Jasiński, R. A stepwise, zwitterionic mechanism for the 1,3-dipolar cycloaddition between (Z)-C-4-methoxyphenyl-N-phenylnitrone and gem-chloronitroethene catalysed by 1-butyl-3-methylimidazolium ionic liquid cations. Tetrahedron Lett. 2015, 56, 532–535. [Google Scholar] [CrossRef]
  161. Dresler, E.; Allnajar, R.; Jasiński, R. Sterical index: A novel, simple tool for the interpretation of organic reaction mechanisms. Sci. Radices 2023, 2, 69–74. [Google Scholar] [CrossRef]
  162. Dennington, R.; Keith, T.A.; Millam, J.M. GaussView, Version 6.0; Semichem Inc.: Shawnee, KS, USA, 2016. [Google Scholar]
Scheme 1. Theoretically possible pathways of 32CA reactions of (Z)-C-phenyl-N-alkyl(phenyl)-nitrones (1ac) with (E)-3-(methylsulfonyl)-propenoic acid derivatives (2 and 3).
Scheme 1. Theoretically possible pathways of 32CA reactions of (Z)-C-phenyl-N-alkyl(phenyl)-nitrones (1ac) with (E)-3-(methylsulfonyl)-propenoic acid derivatives (2 and 3).
Molecules 30 04738 sch001
Figure 1. Visualization of electron density flux from nitrones (1ac) to (E)-3-(methylsulfonyl)-propenoic acid derivatives (23) for tested 32CAs.
Figure 1. Visualization of electron density flux from nitrones (1ac) to (E)-3-(methylsulfonyl)-propenoic acid derivatives (23) for tested 32CAs.
Molecules 30 04738 g001
Figure 2. ωb97xD/6-311G(d,p) local electronic properties presented as three-dimensional atomic-spin representations of Mulliken densities for radical cations of nitrones (1ac) and the radical anion of ethenes (23), together with the nucleophilic Pk and the electrophilic Pk+ Parr functions’ values.
Figure 2. ωb97xD/6-311G(d,p) local electronic properties presented as three-dimensional atomic-spin representations of Mulliken densities for radical cations of nitrones (1ac) and the radical anion of ethenes (23), together with the nucleophilic Pk and the electrophilic Pk+ Parr functions’ values.
Molecules 30 04738 g002
Figure 3. ωb97xD/6-311G(d,p) FMO visualization, showing the HOMO orbitals for nitrones (1ac) and the LUMO orbitals for ethenes (23).
Figure 3. ωb97xD/6-311G(d,p) FMO visualization, showing the HOMO orbitals for nitrones (1ac) and the LUMO orbitals for ethenes (23).
Molecules 30 04738 g003
Figure 4. ωb97xD/6-311G(d,p) MEP maps for nitrones (1ac) and ethenes (23). Regions of most negative (electron-rich) electrostatic potential are shown in red, whereas regions of most positive (electron-deficient) potential are shown in blue.
Figure 4. ωb97xD/6-311G(d,p) MEP maps for nitrones (1ac) and ethenes (23). Regions of most negative (electron-rich) electrostatic potential are shown in red, whereas regions of most positive (electron-deficient) potential are shown in blue.
Molecules 30 04738 g004
Figure 5. Views of critical structures for 32CA reaction of (Z)-C-phenyl-N-methylnitrone (1a) with (E)-3-(methylsulfonyl)-propenoic acid (2) according to ωb97xD/6-311G(d,p) calculations in benzene solution (PCM).
Figure 5. Views of critical structures for 32CA reaction of (Z)-C-phenyl-N-methylnitrone (1a) with (E)-3-(methylsulfonyl)-propenoic acid (2) according to ωb97xD/6-311G(d,p) calculations in benzene solution (PCM).
Molecules 30 04738 g005
Figure 6. The SwissADME [101] bioavailability radars generated for isoxazolidine (411), illustrate their overall drug-likeness, where the pink region represents the recommended intervals for lipophilicity, molecular size, polarity, aqueous solubility, saturation, and molecular flexibility.
Figure 6. The SwissADME [101] bioavailability radars generated for isoxazolidine (411), illustrate their overall drug-likeness, where the pink region represents the recommended intervals for lipophilicity, molecular size, polarity, aqueous solubility, saturation, and molecular flexibility.
Molecules 30 04738 g006aMolecules 30 04738 g006b
Table 1. ωb97xD/6-311G(d,p) global reactivity indices of nitrones (1ac) and (E)-3-(methylsulfonyl)-propenoic acid derivatives (23), presented in gas phase in electronvolts [eV].
Table 1. ωb97xD/6-311G(d,p) global reactivity indices of nitrones (1ac) and (E)-3-(methylsulfonyl)-propenoic acid derivatives (23), presented in gas phase in electronvolts [eV].
1a1b1c23
HOMO energy−7.67−7.60−7.64−10.83−10.23
LUMO energy0.340.44−0.25−0.93−0.52
Electronic chemical potential, μ−3.67−3.58−3.95−5.88−5.38
Chemical hardness, η8.018.057.399.909.72
Global electrophilicity, ω0.840.801.051.751.49
Global nucleophilicity, N3.733.793.750.571.16
Table 2. Energetical parameters of 32CA reactions of nitrones (1ac) with (E)-3-(methylsulfonyl)-propenoic acid derivatives (23) according to ωb97xD/6-311G(d,p) calculations in benzene (PCM) (ΔH, ΔG are in kcal/mol; ΔS are in cal/mol·K).
Table 2. Energetical parameters of 32CA reactions of nitrones (1ac) with (E)-3-(methylsulfonyl)-propenoic acid derivatives (23) according to ωb97xD/6-311G(d,p) calculations in benzene (PCM) (ΔH, ΔG are in kcal/mol; ΔS are in cal/mol·K).
ReactionPathTransitionΔHΔSΔGΔHΔSΔGTransitionPathReaction
1a+2A1a+2MCA−11.9−40.90.3−11.3−39.10.41a+3MCAA1a+3
1a+2TSA3.3−50.318.33.8−48.518.31a+3TSA
1a+24a−28.4−48.5−13.9−25.6−49.6−10.81a+38a
B1a+2MCB−12.5−43.20.4−10.3−38.21.11a+3MCBB
1a+2TSB8.4−51.923.911.2−50.526.21a+3TSB
1a+25a−26.3−53.2−10.4−22.9−53.4−7.01a+39a
C1a+2MCC−13.6−42.8−0.8−10.7−39.71.11a+3MCCC
1a+2TSC4.4−52.620.06.1−49.420.81a+3TSC
1a+26a−30.3−52.7−14.6−27.1−50.7−11.91a+310a
D1a+2MCD−11.5−38.3−0.1−8.9−41.23.31a+3MCDD
1a+2TSD5.5−49.020.16.5−48.721.01a+3TSD
1a+27a−28.1−50.1−13.2−26.2−54.1−10.11a+311a
1b+2A1b+2MCA−13.1−32.0−3.6−11.6−33.1−2.01b+3MCAA1b+3
1b+2TSA1.1−45.214.61.8−43.521.51b+3TSA
1b+24b−32.1−44.7−18.8−28.9−44.2−12.41b+38b
B1b+2MCB−14.0−39.7−2.1−11.8−32.8−2.01b+3MCBB
1b+2TSB6.6−47.420.77.5−46.921.51b+3TSB
1b+25b−30.4−50.7−15.3−26.8−48.1−12.41b+39b
C1b+2MCC−14.7−35.9−4.0−11.6−34.3−3.51b+3MCCC
1b+2TSC3.8−47.317.95.8−45.419.31b+3TSC
1b+26b−32.3−45.2−18.8−29.5−45.3−16.01b+310b
D1b+2MCD−12.1−34.4−1.8−11.6−33.1−1.71b+3MCDD
1b+2TSD2.7−44.816.14.9−43.117.81b+3TSD
1b+27b−32.7−45.6−19.2−27.3−42.3−14.71b+311b
1c+2A1c+2MCA−8.1−42.34.5−11.8−40.50.31c+3MCAA1c+3
1c+2TSA4.6−51.019.85.2−50.220.21c+3TSA
1c+24c−26.3−49.4−11.5−24.0−49.8−9.11c+38c
B1c+2MCB−12.7−42.1−0.1−8.8−35.31.71c+3MCBB
1c+2TSB8.3−54.724.611.0−52.426.61c+3TSB
1c+25c−25.2−54.3−9.0−22.1−52.2−6.51c+39c
C1c+2MCC−13.3−42.8−0.6−11.9−40.60.21c+3MCCC
1c+2TSC6.0−55.522.57.5−52.923.21c+3TSC
1c+26c−28.1−52.2−12.5−25.2−50.3−10.21c+310c
D1c+2MCD−13.9−47.40.3−9.7−33.70.41c+3MCDD
1c+2TSD6.0−49.920.96.5−50.021.41c+3TSD
1c+27c−29.0−50.5−13.9−24.9−46.4−11.01c+311c
Table 3. Key parameters of critical structures for 32CA reactions of nitrones (1ac) with alkenes (23) according to ωb97xD/6-311G(d,p) calculations in benzene solution (PCM).
Table 3. Key parameters of critical structures for 32CA reactions of nitrones (1ac) with alkenes (23) according to ωb97xD/6-311G(d,p) calculations in benzene solution (PCM).
ReactionPathStructureInteratomic Distances [Å]lΔlGEDT *
O1–N2N2–C3C3–C4C4–C5C5–O1C3–C4C5–O1 [e]
1a+2AMCA1.2741.3033.2031.3252.788
TSA1.2881.3442.1191.3922.0080.6400.5910.0490.147
4a1.4191.4611.5581.5301.425
BMCB1.2741.3033.4101.3243.019
TSB1.2791.3482.0781.3902.1220.6590.5170.1420.146
5a1.4321.4621.5501.5261.431
CMCC1.2731.3033.5321.3223.332
TSC1.3061.3322.2911.3931.8670.5360.6730.1370.141
6a1.4371.4611.5651.5201.407
DMCD1.2711.3043.2561.3233.108
TSD1.2991.3372.2111.3921.9160.5880.6460.0580.117
7a1.4251.4601.5661.5241.414
1b+2AMCA1.2761.3073.1731.3233.176
TSA1.2861.3482.1231.3872.0520.6370.5640.0730.125
4b1.4041.4641.5581.5261.429
BMCB1.2751.3073.4311.3243.031
TSB1.2721.3562.0351.3862.2440.6830.4190.2640.128
5b1.4311.4691.5451.5261.419
CMCC1.2811.3054.2931.3224.788
TSC1.3071.3352.2781.3921.8920.5390.6590.1200.127
6b1.4251.4611.5591.5151.411
DMCD1.2771.3074.0251.3223.294
TSD1.2941.3442.1711.3851.9960.6050.5950.0100.097
7b1.4141.4621.5561.5231.420
1c+2AMCA1.2741.3023.2231.3242.973
TSCA1.2831.3472.0601.3952.0270.6710.5720.0990.138
4c1.4231.4671.5501.5231.420
BMCB1.2741.3013.7301.3232.982
TSB1.2631.3551.9961.3912.2450.7070.4060.3010.161
5c1.4491.4661.5441.5251.409
CMCC1.2781.3004.0701.3224.779
TSC1.3051.3312.2731.3941.8560.5440.6740.1300.129
6c1.4441.4601.5611.5151.400
DMCD1.2731.3023.3291.3233.454
TSD1.2991.3362.1701.3951.9000.5990.6230.0240.094
7c1.4531.4771.5491.5141.379
1a+3AMCA1.2741.3043.2291.3292.760
TSA1.2841.3462.0971.4012.0030.6580.5950.0630.195
8a1.4191.4611.5621.5381.426
BMCB1.2831.2995.1741.3293.092
TSB1.2781.3472.1101.3962.0990.6460.5280.1180.199
9a1.4251.4621.5591.5411.426
CMCC1.2751.3033.3121.3292.979
TSC1.3141.3282.3591.4061.7900.4940.7250.2310.201
10a1.4381.4571.5661.5341.404
DMCD1.2771.3024.6821.3253.211
TSD1.3011.3352.2221.3981.8990.5860.6540.0680.049
11a1.4251.4561.5711.5401.411
1b+3AMCA1.2811.3055.5861.3262.975
TSA1.2811.3512.0951.3952.0590.6580.5590.0990.175
8b1.4051.4641.5611.5341.428
BMCB1.2781.3034.7541.3283.063
TSB1.2701.3552.0591.3922.2260.6730.4480.2250.179
9b1.4161.4641.5511.5341.434
CMCC1.2841.3053.8791.3273.150
TSC1.3141.3302.3471.4041.8120.4960.7130.2170.189
10b1.4321.4621.5611.5261.408
DMCD1.2811.3054.7781.3263.457
TSD1.3011.3402.2561.3961.8820.5560.6590.1030.213
11b1.4131.4761.5631.5281.404
1c+3AMCA1.2751.3033.2411.3282.856
TSCA1.2781.3512.0341.4032.0380.6880.5600.1280.194
8c1.4271.4721.5501.5321.415
BMCB1.2781.3015.4921.3272.898
TSB1.2681.3512.0621.3962.1560.6720.4850.1870.200
9c1.4281.4601.5541.5351.423
CMCC1.2781.2994.4111.3264.861
TSC1.3161.3252.3671.4071.7640.4830.7360.2530.194
10c1.4491.4601.5611.5261.396
DMCD1.2791.3005.0921.3263.339
TSD1.3031.3342.1921.4021.8750.6030.6610.0580.163
11c1.4291.4661.5691.5351.400
* GEDT(f) = charge qf.
Table 4. Drug-likeness parameters for isoxazolidines (411), calculated using the SwissADME [101].
Table 4. Drug-likeness parameters for isoxazolidines (411), calculated using the SwissADME [101].
4,5a6,7a4,5b6,7b4,5c6,7c8,9a10,11a8,9b10,11b8,9c10,11c
Physchem. propertiesMW [g/mol]299299361361341341266266328328308308
#heavy atoms202025252323181823232121
#arom. heavy atoms6612126666121266
#rotatable bonds445555223333
#H-bond acceptors665566554455
#H-bond donors000000000000
Molar refractivity76.2876.2896.5296.5290.7490.7469.9469.9490.1890.1884.4084.40
TPSA [Å2]81.2981.2981.2981.2981.2981.2978.7878.7878.7878.7878.7878.78
LipophilicityLog Po/w (iLOGP)1.912.212.202.582.382.621.371.621.752.131.852.04
Log Po/w (XLOGP3)1.011.012.792.791.991.990.880.882.662.661.861.86
Log Po/w (WLOGP)0.941.142.512.722.102.311.291.492.862.972.452.66
Log Po/w (MLOGP)0.820.822.102.101.591.590.550.551.871.871.341.34
Log Po/w (SILICOS-IT)0.130.131.221.220.950.950.190.191.281.280.990.99
Consensus Log Po/w0.961.062.172.281.801.890.850.942.082.201.701.78
Water solubilityLog S (ESOL)−2.29−2.29−3.86−3.86−3.07−3.07−2.16−2.16−3.74−3.74−2.94−2.94
solubility [mg/mL]1.531.530.0490.0490.2880.2881.841.840.0590.0590.3560.356
Log S (Ali)−2.31−2.31−4.15−4.15−3.32−3.32−2.12−2.12−3.97−3.97−3.14−3.14
solubility [mg/mL]1.481.480.0250.0250.1620.1622.032.030.0350.0350.2260.226
Log S (SILICOS-IT)−2.38−2.38−4.47−4.47−3.18−3.18−2.40−2.40−4.50−4.50−3.21−3.21
solubility [mg/mL]1.251.250.0120.0120.2270.2271.061.060.0100.0100.1910.191
PharmacokineticsCYP1A2 inhibitorNoNoYesYesNoNoNoNoNoNoNoNo
CYP2C19 inhibitorNoNoYesYesNoNoNoNoYesYesNoNo
CYP2C9 inhibitorNoNoNoNoNoNoNoNoNoNoNoNo
CYP2D6 inhibitorNoNoNoNoNoNoNoNoNoNoNoNo
CYP3A4 inhibitorNoNoYesYesNoNoNoNoYesYesNoNo
IG absorptionHighHighHighHighHighHighHighHighHighHighHighHigh
BBB permeantNoNoNoNoNoNoNoNoNoNoNoNo
DruglikenessLipinski et al. (Pfizer) [102]
Ghose et al. (Amgen) [103]
Veber et al. (GSK) [104]
Egan et al. (Pharmacia) [105]
Muegge et al. (Bayer) [106]
Table 5. The PASS prediction [107] for isoxazolidine (411) indicates their main potential activities, with results expressed as the probability of activity (Pa) values above 0.700.
Table 5. The PASS prediction [107] for isoxazolidine (411) indicates their main potential activities, with results expressed as the probability of activity (Pa) values above 0.700.
4,5a6,7a4,5b6,7b4,5c6,7c8,9a10,11a8,9b10,11b8,9c10,11c
CYP2H substrate0.7880.8820.7180.8530.760.870------
5 Hydroxytryptamine 2C antagonist--0.759-----0.877---
5 Hydroxytryptamine 2A antagonist--0.724-----0.820---
5 Hydroxytryptamine 2 antagonist--------0.733---
5 Hydroxytryptamine release stimulant-----------0.719
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

Ząbkowska, M.; Kula, K.; Diychuk, V.; Jasiński, R. On the Question of the Regio-, Stereoselectivity and the Molecular Mechanism of the (3+2) Cycloaddition Reaction Between (Z)-C-Phenyl-N-alkyl(phenyl)nitrones and (E)-3-(Methylsulfonyl)-propenoic Acid Derivatives. Molecules 2025, 30, 4738. https://doi.org/10.3390/molecules30244738

AMA Style

Ząbkowska M, Kula K, Diychuk V, Jasiński R. On the Question of the Regio-, Stereoselectivity and the Molecular Mechanism of the (3+2) Cycloaddition Reaction Between (Z)-C-Phenyl-N-alkyl(phenyl)nitrones and (E)-3-(Methylsulfonyl)-propenoic Acid Derivatives. Molecules. 2025; 30(24):4738. https://doi.org/10.3390/molecules30244738

Chicago/Turabian Style

Ząbkowska, Martyna, Karolina Kula, Volodymyr Diychuk, and Radomir Jasiński. 2025. "On the Question of the Regio-, Stereoselectivity and the Molecular Mechanism of the (3+2) Cycloaddition Reaction Between (Z)-C-Phenyl-N-alkyl(phenyl)nitrones and (E)-3-(Methylsulfonyl)-propenoic Acid Derivatives" Molecules 30, no. 24: 4738. https://doi.org/10.3390/molecules30244738

APA Style

Ząbkowska, M., Kula, K., Diychuk, V., & Jasiński, R. (2025). On the Question of the Regio-, Stereoselectivity and the Molecular Mechanism of the (3+2) Cycloaddition Reaction Between (Z)-C-Phenyl-N-alkyl(phenyl)nitrones and (E)-3-(Methylsulfonyl)-propenoic Acid Derivatives. Molecules, 30(24), 4738. https://doi.org/10.3390/molecules30244738

Article Metrics

Back to TopTop