Next Article in Journal
Effect of Pinoresinol and Vanillic Acid Isolated from Catalpa bignonioides on Mouse Myoblast Proliferation via the Akt/mTOR Signaling Pathway
Next Article in Special Issue
Multilayer Graphtriyne Membranes for Separation and Storage of CO2: Molecular Dynamics Simulations of Post-Combustion Model Mixtures
Previous Article in Journal
Molecular Dereplication and In Vitro and In Silico Pharmacological Evaluation of Coriandrum sativum against Neuroblastoma Cells
Previous Article in Special Issue
Sensing the ortho Positions in C6Cl6 and C6H4Cl2 from Cl2 Formation upon Molecular Reduction
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Chemoselective Oxidation of Isoxazolidines with Ruthenium Tetroxide: A Successful Intertwining of Combined Theoretical and Experimental Data

by
Laura Legnani
1,
Salvatore V. Giofré
2,
Daniela Iannazzo
3,
Consuelo Celesti
3,4,
Lucia Veltri
5 and
Maria Assunta Chiacchio
6,*
1
Dipartimento di Biotecnologie e Bioscienze, Università di Milano-Bicocca, Piazza della Scienza 2, 20126 Milano, Italy
2
Dipartimento di Scienze Chimiche, Biologiche, Farmaceutiche ed Ambientali, Università di Messina, Viale F. Stagno D’Alcontres, 98166 Messina, Italy
3
Dipartimento di Ingegneria, Università di Messina, Contrada di Dio, 98166 Messina, Italy
4
Dipartimento di Medicina Clinica e Sperimentale, Università di Messina, Via Consolare Valeria, 98125 Messina, Italy
5
Dipartimento di Chimica e Tecnologie Chimiche, Università della Calabria, Via Pietro Bucci 12/C, 87036 Aracavacata di Rende, Italy
6
Dipartimento di Scienze del Farmaco e della Salute, Università di Catania, Viale A. Doria 6, 95125 Catania, Italy
*
Author to whom correspondence should be addressed.
Molecules 2022, 27(17), 5390; https://doi.org/10.3390/molecules27175390
Submission received: 20 July 2022 / Revised: 20 August 2022 / Accepted: 22 August 2022 / Published: 24 August 2022

Abstract

:
The direct oxidation reaction of isoxazolidines plays an important role in organic chemistry, leading to the synthesis of biologically active compounds. In this paper, we report a computational mechanistic study of RuO4-catalyzed oxidation of differently N-substituted isoxazolidines 1ac. Attention was focused on the endo/exo oxidation selectivity. For all the investigated compounds, the exo attack is preferred to the endo one, showing exo percentages growing in parallel with the stability order of transient carbocations found along the reaction pathway. The study has been supported by experimental data that nicely confirm the modeling results.

1. Introduction

Heterocyclic chemistry [1,2] represents one of the most complex and fascinating branches of organic chemistry of equal interest for its theoretical implications [3,4], involving also almost all aspects of modern organic chemistry. The synthesis and functionalization of heterocycles hold a pivotal role in medicinal chemistry, showing a wide range of pharmaceutical and biological properties. Moreover, heterocyclic compounds are key elements in vitamins, hormones, alkaloids, herbicides, dyes, and other products of industrial importance [5,6,7]. Among the different classes of heterocyclic compounds, some isoxazolidine derivatives, analogs of natural nucleosides and nucleotides, have shown great interest for their anticancer and antiviral properties [8,9,10,11]. The synthetic strategies towards these five-membered heterocyclic rings mostly exploited are the classical 1,3-dipolar cycloadditions of nitrones with differently substituted dipolarophiles [12]. The so-formed cycloadducts can be furtherly functionalized to give the 3-isoxazolidinone nucleus, a cyclic Weinreb amide, whose reduction [13] and nucleosidation lead to reverse transcriptase inhibitors [14]. The first example of direct oxidation of isoxazolidines to the 3-isoxazolidinones was reported in the literature in 2007 [15]. This transformation carried out using RuO2/NaIO4, under ethyl acetate/water biphasic conditions, proved to be highly regioselective, giving only 3-isoxazolidinone derivatives as exclusive compounds.
Based on our expertise in the field of computational mechanistic studies [16,17], we have performed a preliminary study of ruthenium tetroxide-mediated oxidation of some cyclic and heterocyclic compounds. In these studies, DFT and topological methods highlighted that, on these substrates, the rate-limiting step of the reaction takes place through a highly asynchronous (3 + 2) concerted cycloaddition [18]. More recently, we reported a complete computational mechanistic study concerning the oxidation reaction of 2-methylisoxazolidine with RuO4, taking into consideration the different sites where the oxidation could take place [19]. In fact, all the hydrogen atoms of the isoxazolidine system (C-1′, C-3, C-5, and C-4) could be transferred by oxidation with RuO4 and reactions appeared to be competitive. However, the corresponding barriers for oxidation resulted to be correlated to the stability of the transient carbocation forming along the reaction pathway. So, the N-methylisoxazolidin-3-one was detected as the preferred product.
In this paper, with the support of experimental data, we have extended our computational study to isoxazolidines 1ac (Figure 1), bearing a methyl-carboxylate group at C-5 of the isoxazolidine ring, to avoid competition with oxidation in this position [19]. The results of this study will correlate with the endo/exo oxidation selectivity and follow the carbocations stability order.

2. Results and Discussion

2.1. Computational Investigation

All calculations were performed using the Gaussian16 program package [20]. After a preliminary screening considering different levels of calculation already tested on analogous systems, as reported in the literature [18,19], optimizations were performed using the B3LYP functional [21,22] in conjunction with Grimme’s dispersion correction [23,24] (henceforth referred to as B3LYP-d3bj) chosen referring to similar systems’ studies [18,19]. The standard basis set Def2SVP was employed [25,26]. Solvent effects water, using the C-PCM method [27,28]) were taken into consideration. As reported [19], the highly asynchronous [3 + 2] one-step oxidation mechanism [29,30,31,32,33,34] presents two different stages: (a) activation of the R-CH bond, coordinated to the Ru(VI); (b) transfer of the second hydrogen with the obtainment of the oxygenated compound and Ru(IV) (Scheme 1). The first stage is the regioselectivity-determining one and was firstly computationally investigated.
In all the cases, the transition states (TS1), leading to P1 for both the endo and exo pathways (Scheme 1), were located and their 3D plots are reported in Figure 2. For the exo pathway, due to the presence of the stereogenic center at position 5 of the isoxazolidine, and the formation of a second stereogenic center for compounds 1a and 1c, the two possible pro-R and pro-S transition states have been considered. Conversely, for the endo route, it was not possible to locate the pro-S transition state, due to the steric hindrance caused by the presence on C-5 of the methylcarboxylate moiety. For the N-cyclohexyl derivative 1b, the two possible 1C4 and 4C1 chair conformations of the six-membered ring, have been examined. The percentages of the compounds derived from the TSs at 298 K, were calculated and the corresponding values are given in Table 1. As expected, contrary to the endo preference detected for the N-methylisoxazolidine [19], in all the studied compounds, the exo adduct resulted in being favored by different and, from 1a to 1c, growing percentages (63%, 87%, 99%, respectively), related to the stability of the transient carbocation, which is generated during the reaction for the different oxidized compounds.
IRC analyses were performed on all the located TSs1. In the case of the forward direction, limited to the exo attack, for 1ac, a species very similar to an ion pair [19], with a partial character of the double bond between C-3 and N and the O-Ru-oxygen negative charged, was obtained. Additionally, in this case, for compounds 1a and 1c, the two possible diastereomeric routes were investigated. In Figure 3, the three-dimensional plots of representative ion pair structures are reported. The nature of ion pairs IP was confirmed through natural bond order (NBO) analysis (see ESI). When the transfer of the hydrogen atom to the oxygen of the ruthenium occurs, it shows a negative charge (−0.670, −0.694, −0.677 for 1a, 1b, and 1c, respectively), while the corresponding carbon is positively charged (0.205, 0.440 and 0.200 for 1a, 1b, and 1c, respectively). In all ion pairs (IP) large dipole values are detected (IP: 14.3 D; 15.5 D; 15.2 D for 1a, 1b and 1c, respectively). Starting from the ion pair, passing through a very low barrier TSI of about 2 kcal/mol (Figure 3), the products P1 of the first step are obtained. Considering the reaction progression, the second step easily occurred with a new H transfer, with higher barriers (ΔΔG(TS2_endo) = 12.88, 18.63, 8.26 kcal/mol, for 1a, 1b, and 1c, respectively, ΔΔG(TS2_exo_proR) = 23.46, 22.37 kcal/mol; ΔΔG(TS2_exo_proS) = 12.24, 11.90 kcal/mol for 1a and 1c) and so, the second step can be defined as the rate-determining one. For compound 1b, the second step of reaction takes place only in the endo position since there is no second hydrogen in the exo one. The corresponding three-dimensional plots of transition states TS2 are reported in Figure 4.
When the exo attack occurs on isoxazolidines 1ac, the possibility of an alternative route in which a second hydrogen is removed from the endo position, has been evaluated. The obtained products could be the methyl 4,5-dihydroisoxazole-5-carboxylate 4 together with benzaldehyde, cyclohexanone, and 4-metoxybenzaldehyde starting from 1a, 1b, and 1c, respectively (Scheme 2).
The corresponding transition states have been located, and the 3D plots of those related to N-benzyl 1a are shown in Figure 5, as an example. However, the corresponding barriers were found to be greater than 25 kcal/mol and therefore this possibility must be excluded.
Moreover, for compound 1b the second hydrogen transfer can also occur, involving one hydrogen from the β-carbon of the cyclohexyl ring, leading to H2 RuO4 and the corresponding enamine 5 which then evolves to cyclohexanone and methyl-isoxazolidine-5-carboxylate 6 (Scheme 3). However, compound 6 and cyclohexanone could be formed by hydrolysis of P1_exo_cy.
We examined this route, considering the possible transfer of both equatorial and axial hydrogens, locating the corresponding transition states. Nevertheless, the calculated barriers are too high (values exceeding 30 kcal/mol) and the reaction, for the second step, cannot proceed through this pathway.
Finally, for compounds 2a–c, we also explore the possibility of a further attack of RuO4 on P2-endo in position exo, considering the two different diastereomeric routes, as shown in Scheme 4. The process is not energetically demanding, with surmountable barriers that make possible the first step of the reaction (ΔΔG(TS1 b_proR) = 13.83, 11.87 kcal/mol; ΔΔG(TS1 b_proS) = 19.09, 16.17 kcal/mol for 2a and 2c, respectively, and ΔΔG(TS1 b) = 16.80, 14.34 kcal/mol for 2b considering conformation 1C4 and 4C1, respectively), that then evolve to 7 and the corresponding carbonyl compounds.
Once obtained P1 b-exo, we also considered the possibility of the second hydrogen extraction at C-1′ with the obtainment of derivatives 8 (Scheme 5). The corresponding calculated barriers are not so high (ΔΔG(TS2 b_exo_proR) = 19.55, 16.15 kcal/mol; ΔΔG(TS2 b_exo_proS) = 13.17, 10.80 for 2a and 2c) and the oxidation might proceed through the second step. In Figure 6, the 3D plots of the TSs1 b of the two steps and the final products P1 b obtained at the end of the reaction are reported.
Compound 8 can be also obtained by an oxidation process at the endo position of C-3, starting from P2_exo (3a,c) (Scheme 5). The corresponding calculated barriers are low enough to be surmountable ΔΔG(TS1 b_endo) = 17.13, 15.96 kcal/mol; ΔΔG(TS2 b_endo) = 21.12, 13.23 for 3a and 3c).
IRC analysis in the forward direction for TS1 b_exo and TS1 b_endo showed a shoulder, corresponding to the ion pair, as determined for the main route. However, when optimized, the ion pairs fell into the energy holes corresponding to P1 b_exo and P1 b_endo.

2.2. Experimental Investigation

To verify the computational outcomes, the first example we have taken into consideration was the oxidation reaction of methyl 2-benzylisoxazolidine-5-carboxylate 1a. The reaction, using 0.25 equivalent of RuO2 and 1 equivalent of NaIO4, in a biphasic ethyl acetate/water system, occurs in 90 min and provides a mixture of 2a, 3a, and 7 in a ratio of 22:60:18, together with a fair amount of benzaldehyde (15% yield) (Scheme 6).
Compound 2a, according to computational data, can be easily rationalized by an endo attack of RuO4 on the C-3 carbon of the isoxazolidine ring. On the contrary, compound 3a is formed through a C-1′(exo) attack of RuO4 to the benzylic position of isoxazolidine 1a. Instead, the methyl 3-oxoisoxazolidine-5-carboxylate 7 derives from the further oxidation reaction of compound 2a that undergoes a debenzylation process leading also to benzaldehyde (see Scheme 1 and Scheme 4). Thus, the oxidation reaction of 1a proceeds with an exo/endo selectivity of 3:2, showing that this selectivity, according to in silico studies, is controlled by carbocation stability (benzyl vs. secondary carbocation).
Then, following the computational results reported in Table 1, we explored the oxidation reaction of methyl 2-cyclohexyl isoxazolidine-5-carboxylate 1b, where it is expected that the formation of a tertiary carbocation intermediate is able to increase the exo selectivity. The reaction, performed in the same conditions reported in Scheme 6, gave a mixture of compounds 2b, 6, and 7 in a 14:80:6 ratio together with cyclohexanone (Scheme 7). As expected, isoxazolidin-3-one 2b is produced by an endo attack of RuO4 on 1b following the route reported in Scheme 1. Compound 6 is probably formed by hydrolysis of P1_exo_cy (Scheme 3) after exo oxidation of compound 1b, while compound 7 is produced from 2b by elimination of cyclohexyl group (Scheme 4). Additionally, for this reaction, the experimental results agree with computational outcomes affording a 4:1 exo/endo selectivity (tertiary vs. secondary carbocation).
The regioselectivity of RuO4 oxidation favoring the exo attack was also proven on methyl 2-(4-methoxybenzyl) isoxazolidine-5-carboxylate 1c, which contains an electron donor group in the para position of benzyl group (Scheme 8). The reaction, performed following the same synthetic protocol, afforded compound 3c in 74.8% yield, 4-methoxybenzaldehyde, and a low amount of 2c and 7 (5.95% and 4.25 % yield, respectively). The formation of 2c, 3c, and 7 are amenable following the routes described in Scheme 1 and Scheme 4. For this reaction, the exo/endo ratio was found to be about 9:1, in good agreement with the computational data (4-methoxybenzyl vs. secondary carbocation) (See Supplementary Materials).

3. Materials and Methods

3.1. Computational Methods

All calculations were performed using the Gaussian16 program package [20]. Optimizations were performed using the B3 LYP functional [21,22] in conjunction with Grimme’s dispersion correction [23,24] (henceforth referred to as B3 LYP-d3 bj) chosen referring to similar systems’ studies [18,19]. The standard basis set Def2 SVP was employed [25,26]. Solvent effects on water, using the C-PCM method [27,28] were taken into consideration. The reaction pathways were confirmed by IRC analyses performed at the same level as above. Vibrational frequencies were computed at the same level of theory to define the optimized structures as minima or transition states, which present an imaginary frequency corresponding to the forming bonds. Thermodynamics at 298.15 K allowed Gibb’s free energies to be calculated.

3.2. General

Solvents and reagents were used as received from commercial sources. NMR spectra (1H-NMR recorded at 500 MHz, 13C-NMR recorded at 125 MHz) were obtained in CDCl3 solution on a Varian instrument (Agilent Technologies, Palo Alto, CA, USA), and data are reported in ppm relative to TMS as an internal standard. Elemental analyses were performed with a Perkin Elmer elemental analyzer (PerkinElmer, Waltham, MA, USA). MW-assisted reactions were performed on a CEM Discover instrument equipped with electromagnetic stirring and an IR probe used for external temperature control (CEM Corporation, NC, USA). Thin-layer chromatographic separations were carried out on Merck silica gel 60-F254 precoated aluminum plates (Merck, Darmstadt, Germany). Preparative separations were carried out using a Büchi C-601 MPLC instrument (BUCHI Italia S.r.l., Milano, Italy) using Merck silica gel 0.040–0.063 mm, and the eluting solvents were delivered by a pump at the flow rate of 3.5–7.0 mL/min. All solvents were dried according to methods in the literature. Isoxazolidines 1ac have been synthesized according to standard procedures [8,35].

3.3. General Procedure for RuO2/NaIO4 Oxidation

To a solution of NaIO4 (1 mmol) in water (30 mL) was added RuO2 (0.25 mmol) under nitrogen. The resulting green–yellow solution was stirred for 30 min and was followed by addition of isoxazolidine 1ac (0.90 mmol.) in EtOAc (30 mL) in one portion. The solution remained yellowish during the reaction. After 90 min of stirring at room temperature, the mixture was diluted with EtOAc and filtered through a pad of Celite. The organic layer was washed with saturated NaHSO3, which resulted in precipitation of black Ru. The precipitate was filtered off through a pad of Celite. The EtOAc layer was washed with brine and dried with anhydrous Na2 SO4; the solvent was removed by evaporation in a rotary evaporator to obtain the crude product. All products were purified by MPLC chromatography. From 1a we obtained 2a, 3a, and 7 (total yield 85 %), from 1b we obtained 2b, 6, and 7 (total yield 80 %), from 1c we obtained 2c, 3c, and 7 (total yield 85%).
Methyl 2-benzyl-3-oxoisoxazolidine-5-carboxylate (2a): pale yellow oil, yield 18.70%. 1H NMR (500 MHz, CDCl3) δ 7.50–7.21 (m, 5 H), 4.86 (dd, J = 9.7, 5.2 Hz, 1 H), 4.76 (d, J = 15.7 Hz, 1 H), 4.67 (d, J = 15.7 Hz, 1 H), 3.72 (s, 3 H), 3.12 (dd, J = 16.8, 9.7 Hz, 1 H), 2.99 (dd, J = 16.8, 5.2 Hz, 1 H) ppm. 13C NMR (126 MHz, CDCl3) δ 169.57, 167.04, 134.82, 129.21, 128.63, 128.24, 74.29, 52.83, 49.05, 35.89 ppm. Anal. Calcd for C12H13NO4:C, 61.27; H, 5.57; N, 5.95; found C, 61.23; H, 5.55; N, 5.91.
Methyl 2-cyclohexyl-3-oxoisoxazolidine-5-carboxylate (2b): yellow oil, 11.20% yield. 1H NMR (500 MHz, CDCl3) δ 4.83 (t, J = 4.7 Hz, 1 H), 3.82–3.66 (m, 4 H), 3.16–2.95 (m, 2 H), 1.92–1.75 (m, 4 H), 1.66–1.36 (m, 6 H) ppm. 13C NMR (126 MHz, CDCl3) δ 171.45, 171.11, 73.18, 57.02, 52.37, 36.60, 29.93, 26.25, 24.42. Anal. Calcd for C11H17NO4:C, 58.14; H, 7.54; N, 6.16; found C, 58.15; H, 7.51; N, 6.12.
Methyl 2-(4-methoxybenzyl)-3-oxoisoxazolidine-5-carboxylate (2c): pale yellow oil, 5.95% yield. 1H NMR (500 MHz, CDCl3) δ 7.18 (d, J = 8.6 Hz, 2 H), 6.78 (d, J = 8.6 Hz, 2 H), 4.78 (dd, J = 9.6, 5.2 Hz, 1 H), 4.64 (d, J = 15.5 Hz, 1 H), 4.52 (d, J = 15.5 Hz, 1 H), 3.70 (s, 3 H), 3.65 (s, 3 H), 3.03 (dd, J = 16.8, 9.6 Hz, 1 H), 2.90 (dd, J = 16.7, 5.2 Hz, 1 H) ppm. 13C NMR (126 MHz, CDCl3) δ 169.83, 169.58, 159.40, 129.68, 126.85, 113.96, 74.22, 55.28, 52.80, 44.36, 35.91 ppm. Anal. Calcd for C13H15NO5:C, 58.86; H, 5.70; N, 5.28; found C, 58.83; H, 5.68; N, 5.25.
Methyl 2-benzoylisoxazolidine-5-carboxylate (3a): pale yellow oil, 51%. 1 H NMR (500 MHz, CDCl3) δ 7.85–7.80 (m, 2 H), 7.50–7.44 (m, 1 H), 7.43–7.38 (m, 2 H), 4.68–4.64 (m, 1 H), 4.12–4.03 (m, 1 H), 3.86–3.82 (m, 1 H), 3.64 (s, 3 H), 2.66–2.50 (m, 2 H) ppm. 13C NMR (126 MHz, CDCl3) δ 170.31, 167.06, 145.39, 133.44, 128.10, 127.94, 77.25, 52.87, 49.09, 31.26 ppm. Anal. Calcd for C12H13NO4:C, 61.27; H, 5.57; N, 5.95; found C, 61.28; H, 5.56; N, 5.94.
Methyl 2-(4-methoxybenzoyl)isoxazolidine-5-carboxylate (3c): pale yellow oil, 74.8% yield. 1H NMR (500 MHz, CDCl3) δ 7.78 (d, J = 8.9 Hz, 2 H), 6.82 (d, J = 8.9 Hz, 2 H), 4.58 (dd, J = 8.5, 4.1 Hz, 1 H), 3.95 (ddd, J = 10.8, 8.7, 7.1 Hz, 1 H), 3.75 (s, 3 H), 3.79–3.71 (m, 1 H), 3.56 (s, 3 H), 2.57–2.39 (m, 2 H) ppm. 13C NMR (126 MHz, CDCl3) δ 170.78, 170.27, 162.07, 131.34, 125.38, 113.10, 77.14, 55.29, 52.36, 48.41, 31.09 ppm. Anal. Calcd for C13H15NO5:C, 58.86; H, 5.70; N, 5.28; found C, 58.84; H, 5.68; N, 5.29.
Methyl isoxazolidine-5-carboxylate (6): white sticky oil, yield 64%. 1H NMR (500 MHz, CDCl3) δ 4.96 (bs, 1 H), 4.59 (t, J = 3.9 Hz, 1 H), 3.73 (s, 3 H), 3.38–3.26 (m, 1 H), 3.27–3.13 (m, 1 H), 2.23–2.08 (m, 2 H) ppm. 13C NMR (126 MHz, CDCl3) δ 172.25, 77.32, 52.34, 47.82, 29.52 ppm. Anal. Calcd for C5H9NO3:C, 45.80; H, 6.92; N, 10.68; found C, 45.77; H, 6.91; N, 10.62.
Methyl 3-oxoisoxazolidine-5-carboxylate (7): white sticky oil (from 1 a, 15.30% yield; from 1 b, 4.8% yield; from 1 c, 4.25% yield). 1H NMR (500 MHz, CDCl3) δ 8.25 (bs, 1 H), 4.46 (dd, J = 6.1, 5.0 Hz, 1 H), 3.85 (s, 3 H), 2.87 (dd, J = 16.8, 5.0 Hz, 1 H), 2.79 (dd, J = 16.8, 6.1 Hz, 1 H) ppm. 13C NMR (126 MHz, CDCl3) δ 171.91, 170.32, 66.65, 53.27, 23.39 ppm. Anal. Calcd for C5H7NO4:C, 41.38; H, 4.86; N, 9.65; found C, 41.35; H, 4.85; N, 9.66.

4. Conclusions

In this paper, we performed a computational mechanistic study of RuO4-catalyzed oxidation of differently N-substituted isoxazolidines 1ac. Based on our previous results, we first considered the C-3(endo)/C-1′(exo) selectivity, following the supposed [3 + 2] one-step, but asynchronous reaction mechanism with a double hydrogen transfer. In analogy with previously reported results [19], the energy barrier of the second transfer, for compounds 1a and 1c, is higher than that of the first one and so the second transfer can be defined as rate-determining. On the other hand, the first one rules the regioselectivity of the reaction and is considered the regioselectivity-determining step.
From the first hydrogen transfer, passing a low energy barrier, P1_endo and P1_exo are obtained. These intermediates evolve to 2ac and 3a,c, after a second hydrogen extraction, resulting in the final compounds oxidized in C-3 or C-1′, respectively. For all compounds 1ac the exo attack is preferred to the endo one, showing for the corresponding oxidated product percentages growing from 1a to 1c. The computationally determined selectivity in parallel reflects the stability order of transient carbocations involved in the ion pairs. These carbocations were found along the reaction pathway and confirmed by NBO analysis. Once obtained, P1_exo can undergo hydrolysis, affording methyl 4,5-dihydroisoxazole-5-carboxylate 4 together with benzaldehyde, cyclohexanone, or 4-methoxybenzaldehyde.
Based on theoretical results, products 2a, c, and 3a, c can still react with RuO4, giving a new [3 + 2] one-step process with the obtainment of the dicarbonyl derivative 8. Nevertheless, the product of the first step P1 b can be hydrolyzed before the second oxidation, generating the methyl 3-isoxazolidinone-5-carboxylate 7.
For the second reaction step, the possibility of hydrogen extraction in alternative positions was examined, but the energy barrier results were too high.
Finally, the computational outcomes have been experimentally confirmed. For all the investigated reactions, the exo attack is preferred to the endo one, confirming that the oxidation selectivity is strictly related to the stability order of transient carbocations found along the reaction pathway.

Supplementary Materials

The following are available online at https://www.mdpi.com/article/10.3390/molecules27175390/s1, Figure S1: 1H NMR of 2a; Figure S2: 13C NMR of 2a; Figure S3: 1H NMR of 2b; Figure S4: 13C NMR of 2b; Figure S5: 1H NMR of 2c; Figure S6: 13C NMR of 2c; Figure S7: 1H NMR of 3a; Figure S8: 13C NMR of 3a; Figure S9: 1H NMR of 3c; Figure S10: 13C NMR of 3c; Figure S11: 1H NMR of 6; Figure S12: 13C NMR of 6; Figure S13: 1H NMR of 7; Figure S14: 13C NMR of 7; Figure S15: NBO analysis of ion pair IP located along the reaction pathway of compounds 1ac; B3LYP/def2svp/emp=gd3bj/int=ultrafine/solvent=water cartesian coordinates; Table S1: Free energies and imaginary frequencies for transition states to the oxidation reaction of 1a; Table S2: Free energies and imaginary frequencies for transition states to the oxidation reaction of 1b; Table S3: Free energies and imaginary frequencies for transition states to the oxidation reaction of 1b.

Author Contributions

Conceptualization, L.L. and M.A.C.; methodology, L.L., M.A.C., S.V.G., L.V., C.C. and D.I.; software, L.L.; validation, M.A.C.; investigation, S.V.G., L.V. and D.I.; writing—original draft preparation, L.L. and M.A.C.; writing—review and editing, M.A.C. and D.I.; supervision, M.A.C. All authors have read and agreed to the published version of the manuscript.

Funding

The authors thank the Universities of Catania (progetto PIACERI 2020–2022 grant number 57722172121), Messina, and Milano-Bicocca for partial financial support.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

The authors are very grateful to Professor Pedro Merino and Professor Ugo Chiacchio for helpful discussion and to Professor Alessio Porta for providing machine time.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Nakamura, I.; Yamamoto, Y. Transition-Metal-Catalyzed Reactions in Heterocyclic Synthesis. Chem. Rev. 2004, 104, 2127–2198. [Google Scholar] [CrossRef] [PubMed]
  2. Scriven, E.; Ramsden, C.A. Advances in Heterocyclic Chemistry, 1st ed.; Elsevier Inc.: Amsterdam, The Netherlands, 2021. [Google Scholar]
  3. Ormachea, C.M.; Kneeteman, M.N.; Mancini, P.M.E. Diels–Alder Polar Reactions of Azaheterocycles: A Theoretical and Experimental Study. Organics 2022, 3, 102–110. [Google Scholar] [CrossRef]
  4. Pappalardo, A.; Ballistreri, F.P.; Toscano, R.M.; Chiacchio, M.A.; Legnani, L.; Grazioso, G.; Veltri, L.; Trusso Sfrazzetto, G. Alkene epoxidations mediated by mn-salen macrocyclic catalysts. Catalysts 2021, 11, 465. [Google Scholar] [CrossRef]
  5. Jiang, B.; Rajale, T.; Wever, W.; Tu, S.-J.; Li, G. Multicomponent Reactions for the Synthesis of Heterocycles. Chem. Asian J. 2010, 5, 2318–2335. [Google Scholar] [CrossRef] [PubMed]
  6. Martins, M.A.P.; Frizzo, C.P.; Moreira, D.N.; Buriol, L.; Machado, P. Solvent-Free Heterocyclic Synthesis. Chem. Rev. 2009, 109, 4140–4182. [Google Scholar] [CrossRef] [PubMed]
  7. Chiacchio, M.A.; Giofrè, S.V.; Romeo, R.; Romeo, G.; Chiacchio, U. Isoxazolidines as biologically active compounds. Curr. Org. Chem. 2016, 13, 726–749. [Google Scholar] [CrossRef] [Green Version]
  8. Giofrè, S.V.; Romeo, R.; Carnovale, C.; Mancuso, R.; Cirmi, S.; Navarra, M.; Garozzo, A.; Chiacchio, M.A. Synthesis and biological properties of 5-(1H-1,2,3-triazol-4-yl)isoxazolidines: A new class of C-nucleosides. Molecules 2015, 20, 5260–5275. [Google Scholar] [CrossRef]
  9. Merino, P.; Greco, G.; Tejero, T.; Hurtado-Guerrero, R.; Matute, R.; Chiacchio, U.; Corsaro, A.; Pistarà, V.; Romeo, R. Stereoselective 1,3-dipolar cycloadditions of nitrones derived from amino acids. Asymmetric synthesis of N-(alkoxycarbonylmethyl)-3-hydroxypyrrolidin-2-ones. Tetrahedron 2013, 69, 9381–9390. [Google Scholar] [CrossRef]
  10. Chiacchio, M.A.; Legnani, L.; Chiacchio, U. Recent Advances in the Synthesis of Isoxazolidines. In Synthetic Approaches to Nonaromatic Nitrogen Heterocycles; Phillips, A.M.F., Ed.; John Wiley & Sons Ltd.: Hoboken, NJ, USA, 2020; Volume 1, pp. 161–177. [Google Scholar]
  11. Romeo, G.; Chiacchio, U.; Corsaro, A.; Merino, P. Chemical synthesis of heterocyclic sugar nucleoside analogues. Chem. Rev. 2010, 110, 3337–3370. [Google Scholar] [CrossRef]
  12. Breugst, M.; Reissig, H.-U. The Huisgen Reaction: Milestones of the 1,3-Dipolar Cycloaddition. Angew. Chem. Int. Ed. 2020, 59, 12293–12307. [Google Scholar] [CrossRef] [Green Version]
  13. Lanza, G.; Chiacchio, M.A.; Giofrè, S.V.; Romeo, R.; Merino, P. The high selectivity of the Cp2 ZrHCl reducing agent for imides: A combined experimental and theoretical study on γ-lactam and isoxazolidinone derivatives. Eur. J. Org. Chem. 2012, 2013, 95–104. [Google Scholar] [CrossRef]
  14. Romeo, R.; Giofré, S.V.; Macchi, B.; Balestrieri, M.; Mastino, A.; Merino, P.; Carnovale, C.; Romeo, G.; Chiacchio, U. Truncated Reverse Isoxazolidinyl Nucleosides: A New Class of Allosteric HIV-1 Reverse Transcriptase Inhibitors. ChemMedChem 2012, 7, 565–569. [Google Scholar] [CrossRef]
  15. Piperno, A.; Chiacchio, U.; Iannazzo, D.; Giofré, S.V.; Romeo, G.; Romeo, R. First Example of Direct RuO4-Catalyzed Oxidation of Isoxazolidines to 3-Isoxazolidones. J. Org. Chem. 2007, 72, 3958–3960. [Google Scholar] [CrossRef] [PubMed]
  16. Legnani, L.; Puglisi, R.; Pappalardo, A.; Chiacchio, M.A.; Sfrazzetto, G.T. Supramolecular recognition of phosphocholine by an enzyme-like cavitand receptor. Chem. Commun. 2020, 56, 539–542. [Google Scholar] [CrossRef] [PubMed]
  17. Chiacchio, M.A.; Legnani, L.; Caramella, P.; Tejero, T.; Merino, P. Revealing carbocations in highly asynchronous concerted reactions: The ene-type reaction between dithiocarboxylic acids and alkenes. Tetrahedron 2018, 74, 5627–5634. [Google Scholar] [CrossRef]
  18. Pedrón, M.; Legnani, L.; Chiacchio, M.A.; Caramella, P.; Tejero, T.; Merino, P. Transient and intermediate carbocations in ruthenium tetroxide oxidation of saturated rings. Beilstein J. Org. Chem. 2019, 15, 1552–1562. [Google Scholar] [CrossRef] [PubMed]
  19. Chiacchio, M.A.; Iannazzo, D.; Giofré, S.V.; Romeo, R.; Legnani, L. Ruthenium Tetroxide Oxidation of N-Methyl-Isoxazolidine: Computational Mechanistic Study. Arab. J. Chem. 2022, 15, 104063. [Google Scholar] [CrossRef]
  20. Gaussian 16, Revision C.01; Gaussian Inc.: Wallingford, CT, USA, 2016.
  21. Becke, A.D. Density-functional thermochemistry. III. The role of exact exchange. J. Chem. Phys. 1993, 98, 5648–5652. [Google Scholar] [CrossRef] [Green Version]
  22. Lee, C.; Yang, W.; Parr, R.G. Development of the Colle-Salvetti correlation-energy formula into a functional of the electron density. Phys. Rev. B 1988, 37, 785–789. [Google Scholar] [CrossRef] [Green Version]
  23. Grimme, S.; Ehrlich, S.; Goerigk, L. Effect of the damping function in dispersion corrected density functional theory. J. Comput. Chem. 2011, 32, 1456–1465. [Google Scholar] [CrossRef]
  24. Grimme, S.; Antony, J.; Ehrlich, S.; Krieg, H. A consistent and accurate ab initio parametrization of density functional dispersion correction (DFT-D) for the 94 elements H-Pu. J. Chem. Phys. 2010, 132, 154104. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  25. Weigend, F. Accurate Coulomb-fitting basis sets for H to Rn. Phys. Chem. Chem. Phys. 2006, 8, 1057–1065. [Google Scholar] [CrossRef] [PubMed]
  26. Weigend, F.; Ahlrichs, R. Balanced basis sets of split valence, triple zeta valence and quadruple zeta valence quality for H to Rn: Design and assessment of accuracy. Phys. Chem. Chem. Phys. 2005, 7, 3297–3305. [Google Scholar] [CrossRef]
  27. Barone, V.; Cossi, M.J. Quantum Calculation of Molecular Energies and Energy Gradients in Solution by a Conductor Solvent Model. Phys. Chem. A 1998, 102, 1995–2001. [Google Scholar] [CrossRef]
  28. Cossi, M.; Rega, N.; Scalmani, G.; Barone, V. Energies, structures, and electronic properties of molecules in solution with the C-PCM solvation model. J. Comput. Chem. 2003, 24, 669–681. [Google Scholar] [CrossRef]
  29. Mitka, K.; Fela, K.; Olszewska, A.; Jasiński, R. On the Question of Zwitterionic Intermediates in the [3 + 2] Cycloaddition Reactions between C-arylnitrones and Perfluoro 2-Methylpent-2-ene. Molecules 2021, 26, 7147. [Google Scholar] [CrossRef]
  30. Dresler, E.; Kącka-Zych, A.; Kwiatkowska, M.; Jasiński, R. Regioselectivity, stereoselectivity, and molecular mechanism of [3  +  2] cycloaddition reactions between 2-methyl-1-nitroprop-1-ene and (Z)-C-aryl-N-phenylnitrones: A DFT computational study. J. Mol. Model. 2018, 24, 329. [Google Scholar] [CrossRef] [Green Version]
  31. Jasiński, R. Competition between one-step and two-step mechanism in polar [3 + 2] cycloadditions of (Z)-C-(3,4,5-trimethoxyphenyl)-N-methyl-nitrone with (Z)-2-EWG-1-bromo-1-nitroethenes. Comput. Theor. Chem. 2018, 1125, 77–85. [Google Scholar] [CrossRef]
  32. Kącka-Zych, A.; Domingo, L.R.; Jasiński, R. Does a fluorinated Lewis acid catalyst change the molecular mechanism of the decomposition process of nitroethyl carboxylates? Res. Chem. Intermed. 2018, 44, 325–337. [Google Scholar] [CrossRef]
  33. Kącka-Zych, A.; Domingo, L.R.; Ríos-Gutiérrez, M.; Jasiński, R. Understanding the mechanism of the decomposition reaction of nitroethyl benzoate through the Molecular Electron Density Theory. Theor. Chem. Acc. 2017, 136, 129. [Google Scholar] [CrossRef] [Green Version]
  34. Kącka, A.B.; Jasiński, R.A. A density functional theory mechanistic study of thermal decomposition reactions of nitroethyl carboxylates: Undermine of “pericyclic” insight. Heteroat. Chem. 2016, 27, 279–289. [Google Scholar] [CrossRef]
  35. Katahara, S.; Takahashi, T.; Nomura, K.; Uchiyama, M.; Sato, T.; Chida, N. Copper-Catalyzed Electrophilic Etherification of Arylboronic Esters with Isoxazolidines. Chem. Asian J. 2020, 15, 1869–1872. [Google Scholar] [CrossRef] [PubMed]
Figure 1. N-substituted isoxazolidines.
Figure 1. N-substituted isoxazolidines.
Molecules 27 05390 g001
Scheme 1. Reaction mechanism of oxidation with RuO4 of compounds 1ac in α position to the nitrogen atom.
Scheme 1. Reaction mechanism of oxidation with RuO4 of compounds 1ac in α position to the nitrogen atom.
Molecules 27 05390 sch001
Figure 2. Three-dimensional plots of the transition states TS1 for the endo and exo pathways of compounds 1ac. Displacement vectors for TS imaginary frequencies are shown as dotted lines and distances are reported in angstroms.
Figure 2. Three-dimensional plots of the transition states TS1 for the endo and exo pathways of compounds 1ac. Displacement vectors for TS imaginary frequencies are shown as dotted lines and distances are reported in angstroms.
Molecules 27 05390 g002
Figure 3. Three-dimensional plots of representative conformations of ion pairs (IP) and the corresponding transition states (TSIP) leading to P1 for compounds 1ac. For TSs imaginary frequencies are shown as dotted lines and distances are reported in angstroms.
Figure 3. Three-dimensional plots of representative conformations of ion pairs (IP) and the corresponding transition states (TSIP) leading to P1 for compounds 1ac. For TSs imaginary frequencies are shown as dotted lines and distances are reported in angstroms.
Molecules 27 05390 g003
Figure 4. Three-dimensional plots of TS2 leading to P2 for compounds 1ac. For TSs imaginary frequencies are shown as dotted lines and distances are reported in angstroms.
Figure 4. Three-dimensional plots of TS2 leading to P2 for compounds 1ac. For TSs imaginary frequencies are shown as dotted lines and distances are reported in angstroms.
Molecules 27 05390 g004
Scheme 2. Alternative route for a second H-transfer from the endo position of isoxazolidines 1ac.
Scheme 2. Alternative route for a second H-transfer from the endo position of isoxazolidines 1ac.
Molecules 27 05390 sch002
Figure 5. Three-dimensional plots of TSs_sec for compounds 1a and 1c, TSs2 for compound 1b, and the corresponding products P_sec. For TSs imaginary frequencies are shown as dotted lines and distances are reported in angstroms.
Figure 5. Three-dimensional plots of TSs_sec for compounds 1a and 1c, TSs2 for compound 1b, and the corresponding products P_sec. For TSs imaginary frequencies are shown as dotted lines and distances are reported in angstroms.
Molecules 27 05390 g005
Scheme 3. Alternative route for the second H-transfer from β-position of cyclohexyl ring for the exo pathway of 1b.
Scheme 3. Alternative route for the second H-transfer from β-position of cyclohexyl ring for the exo pathway of 1b.
Molecules 27 05390 sch003
Scheme 4. First step of the further oxidation reaction on P2-endo (2a–c), leading to 7 and corresponding carbonyl compounds.
Scheme 4. First step of the further oxidation reaction on P2-endo (2a–c), leading to 7 and corresponding carbonyl compounds.
Molecules 27 05390 sch004
Figure 6. Three-dimensional plots of TSs1–2b and P1 b products. TSs imaginary frequencies are shown as dotted lines and distances are reported in angstroms.
Figure 6. Three-dimensional plots of TSs1–2b and P1 b products. TSs imaginary frequencies are shown as dotted lines and distances are reported in angstroms.
Molecules 27 05390 g006
Scheme 5. Oxidation reaction on P2-endo (2a,c) and P2-exo (3a,c), leading to P2 b (8).
Scheme 5. Oxidation reaction on P2-endo (2a,c) and P2-exo (3a,c), leading to P2 b (8).
Molecules 27 05390 sch005
Scheme 6. Reagents and conditions: (i) RuO2, NaIO4, ethyl acetate, H2 O, rt., 90 min.
Scheme 6. Reagents and conditions: (i) RuO2, NaIO4, ethyl acetate, H2 O, rt., 90 min.
Molecules 27 05390 sch006
Scheme 7. Reagents and conditions: (i) RuO2, NaIO4, ethyl acetate, H2 O, rt, 90 min.
Scheme 7. Reagents and conditions: (i) RuO2, NaIO4, ethyl acetate, H2 O, rt, 90 min.
Molecules 27 05390 sch007
Scheme 8. Reagents and conditions: (i) RuO2, NaIO4, ethyl acetate, H2 O, rt, 90 min.
Scheme 8. Reagents and conditions: (i) RuO2, NaIO4, ethyl acetate, H2 O, rt, 90 min.
Molecules 27 05390 sch008
Table 1. Relative free energies (kcal/mol) of TSs and percentages of the corresponding products at 298 K of the oxidation reaction with RuO4 of compounds 1ac.
Table 1. Relative free energies (kcal/mol) of TSs and percentages of the corresponding products at 298 K of the oxidation reaction with RuO4 of compounds 1ac.
Compounds ΔG
(kcal/mol)
%Σ
(endo and exo %)
1aTS1_Bn_endo
TS1_Bn_exo_proR
TS1_Bn_exo_proS
0.00
0.12
0.10
3737
3163
32
1bTS1_cy_endo_4C1
TS1_cy_endo_1C4
TS1_cy_exo_4C1
TS1_cy_exo_1C4
1.13
1.42
0.00
0.24
8
5
13
52
35
87
1cTS1_pOMe_endo
TS1_pOMe _exo_proR
TS1_pOMe _exo_proS
2.32
0.00
0.21
11
58
41
99
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Legnani, L.; Giofré, S.V.; Iannazzo, D.; Celesti, C.; Veltri, L.; Chiacchio, M.A. Chemoselective Oxidation of Isoxazolidines with Ruthenium Tetroxide: A Successful Intertwining of Combined Theoretical and Experimental Data. Molecules 2022, 27, 5390. https://doi.org/10.3390/molecules27175390

AMA Style

Legnani L, Giofré SV, Iannazzo D, Celesti C, Veltri L, Chiacchio MA. Chemoselective Oxidation of Isoxazolidines with Ruthenium Tetroxide: A Successful Intertwining of Combined Theoretical and Experimental Data. Molecules. 2022; 27(17):5390. https://doi.org/10.3390/molecules27175390

Chicago/Turabian Style

Legnani, Laura, Salvatore V. Giofré, Daniela Iannazzo, Consuelo Celesti, Lucia Veltri, and Maria Assunta Chiacchio. 2022. "Chemoselective Oxidation of Isoxazolidines with Ruthenium Tetroxide: A Successful Intertwining of Combined Theoretical and Experimental Data" Molecules 27, no. 17: 5390. https://doi.org/10.3390/molecules27175390

Article Metrics

Back to TopTop