Next Article in Journal
The Critical Role of 12-Methyl Group of Anthracycline Dutomycin to Its Antiproliferative Activity
Next Article in Special Issue
Asymmetrically Substituted Phospholes as Ligands for Coinage Metal Complexes
Previous Article in Journal
Synthesis of PP2A-Activating PF-543 Derivatives and Investigation of Their Inhibitory Effects on Pancreatic Cancer Cells
Previous Article in Special Issue
Discovery of Novel Non-Oxime Reactivators Showing In Vivo Antidotal Efficiency for Sarin Poisoned Mice
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

1,2σ3λ3-Oxaphosphetanes and Their P-Chalcogenides—A Combined Experimental and Theoretical Study

by
Florian Gleim
1,†,
Antonio García Alcaraz
2,†,
Gregor Schnakenburg
1,
Arturo Espinosa Ferao
2,* and
Rainer Streubel
1,*
1
Institut für Anorganische Chemie, Rheinische Friedrich-Wilhelms-Universität Bonn, Gerhard-Domagk-Straße 1, 53121 Bonn, Germany
2
Departamento de Química Orgánica, Facultad de Química, Campus de Espinardo, Universidad de Murcia, 30100 Murcia, Spain
*
Authors to whom correspondence should be addressed.
These authors contributed equally to this work.
Submission received: 30 April 2022 / Revised: 13 May 2022 / Accepted: 17 May 2022 / Published: 23 May 2022

Abstract

:
Although 1,2σ5λ5-oxaphosphetanes have been known for a long time, the “low-coordinate” 1,2σ3λ3-oxaphosphetanes have only been known since their first synthesis in 2018 via decomplexation. Apart from ligation of this P-heterocycle to gold(I)chloride and the oxidation using ortho-chloranil, nothing on their chemistry has been reported so far. Herein, we describe the synthesis of new 1,2σ3λ3-oxaphosphetane complexes (3ae) and free derivatives (4ae), as well as reactions of 4a with chalcogens and/or chalcogen transfer reagents, which yielded the P-chalcogenides (1416a; Ch = O, S, Se). We also report on the theoretical results of the reaction pathways of C-phenyl-substituted 1,2 σ3λ3-oxaphosphetanes and ring strain energies of 1,2σ4λ5-oxaphosphetane P-chalcogenides.

1. Introduction

Strained organic and inorganic ring systems [1] are of high interest, due to their special bonding situation and high reactivity; for example, oxetanes (I) (Figure 1) are important building blocks for the synthesis of more complicated molecules [2] and polymers [3]. The phosphorus-containing four-membered rings, the phosphetanes (II), drew attention because of their use as steering ligands in transition metal catalysis [4], and, more recently, their performance as organocatalyst [5,6,7]. The class of oxaphosphetanes can be regarded as an unusual combination of the features of oxetanes (I) and phosphetanes (II), which have been scarcely studied so far. Please note that isomeric 1,3-oxaphosphetanes (III) [8] and 1,2-oxaphosphetanes (IV) [9] do also exist.
In case of III and IV, the higher substituted compounds have been investigated more often. For example, 1,3σ4λ5-oxaphosphetanes are available through intramolecular Mitsunobu reactions [8], and bi- and tricyclic 1,3σ3λ3-oxaphosphetanes have recently been proposed in the decomposition of HPCO [10]. The high-coordinate 1,2σ5λ5-oxaphosphetanes (IVb) were known for a long time as intermediates in the Wittig reaction, although they do not occur in all cases [9,11,12], or in the deoxygenation of epoxides [12,13]. Recent calculations by Espinosa show that they also occur in the phosphite-initiated reductive dimerization of ketones [14]. Until now, very few crystal structures of 1,2σ5λ5-oxaphosphetanes (IVb) were reported [15,16,17,18].
In contrast, only the low-coordinate 1,2σ3λ3-oxaphosphetanes (IVa) were proposed [19] for a long time, with no stable derivative known. We synthesized the corresponding κP-pentacarbonylmetal(0) complexes (M = Cr, Mo, W) (V) either through ring expansion of epoxides using highly reactive phosphinidenoid complexes [20,21], or ring formation through intramolecular nucleophilic attack. Recently, the free ligand was obtained [22] using a decomplexation strategy [23] for octahedral complexes [M(CO)5L] by a combined thermal substitution with the chelating effect of bis(diphenylphosphino)ethane (DPPE). The first X-ray structure of a non-ligated 1,2σ3λ3-oxaphosphetane (IVa) was also reported together with the P-oxidation using ortho-chloranil and the P-complexation of gold(I)chloride [22].
Similar to 1,2σ3λ3-oxaphosphetanes (IVa), P-chalcogenides (VI) are rather elusive compounds. The 1,2-oxaphosphetane P-oxides (VI, E = O) were first reported by Regitz in 1973 [24], but it was later found by Inamoto that these products were in fact 3,4-dihydro-1H-2,3-benzoxaphosphorin 3-oxides [25]. In 1976, Inamoto proposed a 1,2-oxaphosphetane oxide as a reaction product of a phosphinidene oxide and trans-stilbene oxide [26]. However, the same author published a revised structure in 1991, showing that the product was in fact an acyclic secondary phosphine oxide [27]. In 1991, Hafez proposed an annulated 1,2-oxaphosphetane P-oxide-like structure for the photochemical reaction product of flavone with the Lawesson reagent. The product was only characterized by mass spectrometry, IR- and 1H-NMR spectroscopy, and elemental analysis, but as the publication lacks 13C- and 31P-NMR data and the four-membered ring bears no hydrogen atoms, this assignment might be incorrect [28]. In 1994, Okazaki reported the synthesis of a 1,2-oxaphosphetane P-oxide, kinetically stabilized through a bulky 2,4,6-triisopropylphenyl group at phosphorus [29]. Regarding the 1,2-oxphosphetane P-sulfides (VI, E = S), there is only one publication proposing 2-alkylthio-1,2-oxaphosphetane P-sulfides as the thermodynamically stable product of the reaction of 3-alkylamino-2-butenoic esters with phosphorus pentasulfide [30]. To the best of our knowledge, the 1,2-oxaphosphetane selenides and tellurides (VI, E = Se, Te) are unknown so far.
Herein, syntheses of new C4-substituted 1,2σ3λ3-oxaphosphetanes, the mechanistic evaluation of this reaction for model compounds using DFT calculations, as well as efforts to access their P-chalcogenides (VI, E = O, S, Se, Te) are described.

2. Results

2.1. Synthesis and Spectroscopic Characterization of 1,2σ3λ3-Oxaphosphetanes

Firstly, the protocol currently used for accessing 1,2-oxaphosphetanes [21,22,31] is significantly improved. In the absence of 12-crown-4 the P-triphenylmethyl (trityl) substituted Li/Cl phosphinidenoid complex 1 reacted cleanly with epoxides 2ad in THF yielding the oxaphosphetane complexes 3a,a’3d,d’ (Scheme 1). The new complexes 3b,b’d,d’ could be isolated as pairs of diastereomers (Table 1). Compounds 3a,a’d,d’ were then treated with 1,2-bis(diphenylphosphino)ethane (DPPE) at 80 °C for two days. The formation of the desired products 4a,a’d,d’ was shown by 31P{1H}-NMR spectroscopy (Scheme 1). For 31P{1H}-NMR parameters, as well as product ratios, see Table 2. For all 3b,b’e,e’, 3e*,e*’, and 4b,b’d,d’ diastereomeric pairs, the more highfield shifted 31P{1H}-NMR signal can be tentatively assigned to the cis-isomers and the downfield shifted signal to the trans-isomers, based on former calculations for 3a,a’ and 4a,a’ [22].
It should be noted that the change in isomer ratio from the complexes 3a,a’d,d’ to the free 1,2-oxaphosphetanes 4a,a’d,d’ can be attributed to the method of purification (extraction with n-pentane). At the end of the reaction, the isomer ratio closely resembles that of the starting material, the final difference arising from slightly different solubilities of the isomers in n-pentane.
Crystal structures of complexes 3b,b’d,d’ were obtained (see ESI), but the change of C4-substituent did not lead to significant changes of bond lengths or angles compared to similar known compounds reported in the literature [21]. In the case of unligated species 4b,b’ (Figure 2) and 4c,c’ (see ESI), their crystal structures were obtained after recrystallization from n-pentane. The bond lengths of 4b,b’c,c’ are very similar compared to their metal complexes 3b,b’c,c’. The bonds of phosphorus change by less than 2%. The change of the dihedral angle of the ring system is more prominent; for example, for 3b,b’ the dihedral angles are approximately 150° (cis) and 170° (trans), whereas the trans form of 4b,b’ is nearly planar and the cis-form bent more strongly (around 130°).
As in the case of the P-bis(trimethylsilyl)methyl substituted phosphinidenoid complex, the reaction of 1 with styrene oxide (2e) did not lead to the C4-, but preferentially to the C3-substituted 1,2-oxaphosphetane complexes [31]. The synthesis of a phenyl-substituted 1,2-oxaphosphetane was also attempted via reaction of 1 with the above mentioned oxirane derivative (2e), hoping to profit from the huge steric demand of the trityl group. However, this reaction led to a mixture of four isomers, the diastereomeric pairs of the C4- (3e,e’) and C3-substituted complexes (3e*,e*’) (Scheme 2), whose NMR data and ratios are collected in Table 3. The assignment of the 31P{1H}-NMR chemical shifts to the C3- and C4-substituted regioisomers is based on the P-bis(trimethylsilyl)methyl substituted case, where only the C3-substituted regioisomers are formed (proven by crystal structures), and where it is shown that they are downfield shifted in comparison to other C4-substituted derivatives [31]. Unfortunately, the mixture of 3e,e’ and 3e*,e*’ could not be separated using column chromatography, even at a lower temperature.

2.2. DFT-Based Mechanistic Proposal

Quantum chemical calculations were performed to provide further insights into mechanistic aspects of the formation of C-phenyl-substituted 1,2-oxaphosphetanes 3e,e’ and 3e*,e*’. For the sake of computational economy, a methyl group (instead of trityl) was used as P-substituent in the Li/Cl phosphinidenoid moiety. Additionally, diethylene glycol dimethyl ether (DEGDME) was used as a model to provide an almost saturated coordination sphere for the Li cation. The approach of complex 5 to styrene oxide 2e (taking the S enantiomer for the model study) gave rise to a van der Waals complex 6, where the Li(DEGDME) group is coordinated to the epoxide O atom in a barrierless, thermodynamically favorable process, furnishing complex 7 (Scheme 3). Given the high oxophilicity of phosphorus centers, nucleophilic attack of the negatively charged P atom to the electron-deficient O atom in the cationic part was first studied. By elimination of the solvated LiCl salt 8, terminal phosphinidene-epoxide adduct 9 was formed in a markedly endergonic transformation, for which a TS could not be located. The singularity of a terminal phosphinidene pentacarbonyltungsten(0) oxirane adduct was recently studied [32], showing the weakest O→P bond among the whole series of cyclic ethers adducts of phosphinidene complexes. In contrast, complex 9 displayed a strengthened P→O bond with similar bond strength descriptors values than those obtained when the O donor is dimethyl ether 9OMe2 (Table S1). Elongation of the less activated, non-benzylic C-O bond of 9 gives exergonically styrene 10 and phosphinidene oxide complex 11 through a moderate barrier (18.41 kcal mol−1). A similar result was found previously for a terminal phosphinidene molybdenum(0) thiirane complex, giving rise to ethylene and a side-on complexed phosphinidene sulfide [33]. On the contrary, P insertion into the benzylic C-O bond proceeds through a lower-energy TS (9.81 kcal mol−1) affording 1,2-oxaphosphetane 12* (C3-substituted) in a markedly exergonic process.
A more favorable pathway to obtain the desired products resulted from the direct nucleophilic attack of the P atom to the epoxide C atoms of 7 (Scheme 4). The attack at the more positively charged benzylic carbon (qN = 0.03 e) is slightly kinetically favored (∆∆EZPE = 0.69 kcal mol−1) (Figure 3), due to a higher C-O bond activation (WBI = 0.881, MBO = 0.778) and the low steric hindrance of the methyl group at the phosphanido moiety. Conversely, the attack to the non-benzylic carbon atom (qN = −0.10 e), with a comparatively strengthened C-O bond (WBI = 0.908, MBO = 0.945), leads to a more stable alkoxide 13 (Figure 3). However, when a tert-butyl group is attached to phosphorus, the attack to the non-benzylic carbon is (slightly) kinetically favored (see ESI). Therefore, in the real system with a trityl group, an even more favorable non-benzylic C-O insertion would be expected, due to the high steric hindrance. The cyclization to form the four-membered 1,2-oxaphosphetanes proceeds in both cases through similar energy TSs. The most stable isomer 12 is obtained (initially as the van der Waals complex 8·12) through the slightly higher energy barrier process. The pathways leading to minor diastereomers 12′ and 12*’ were also computed (see ESI).

2.3. Synthesis of 1,2-Oxaphosphetane P-Chalcogenides

As the main goal of the study was to synthesize various 1,2σ4λ5-oxaphosphetane chalcogenide derivatives, the 42:58 mixture of 4-methyl-1,2-oxaphosphetane 4a,4a’ was used as a good case in point.
In order to target P-oxide derivatives, reactions of the mixture 4a,a’ with various oxygen-transfer reagents were studied. Treating 4a,a’ with propylene oxide or trimethylamine N-oxide in toluene at r.t. was not effective to convert 4a,a’ into 14a,a’. The use of tert-butylhydroperoxide or meta-chloroperoxybenzoic acid (mCPBA) led to unselective reactions; however, the reaction using iodosylbenzene (Scheme 5) led to the selective formation of 1,2-oxaphosphetane P-oxides 14a,a’. The product was fully characterized by NMR spectroscopy, as well as ESI and APCI mass spectrometry. The 31P{1H}-NMR spectrum of the product solution showed two resonance signals of 62.1 ppm and 63.5 ppm, in a ratio of 66:34. This assignment fits well with the reported shift of the P-triisopropylphenyl substituted 1,2σ4λ5-oxaphosphetane P-oxide (δ (31P{1H}) = 48.7 ppm [29]).
To synthesize 1,2σ4λ5-oxaphosphetane P-sulfide 15a,a’, 4a,a’ was treated with elemental sulfur in toluene at ambient temperature (Scheme 6). The reaction occurred selectively; 15a,a’ was isolated via extraction from n-pentane and it was fully characterized by NMR spectroscopy and LIFDI mass spectrometry. 31P{1H}-NMR chemical shifts of the isomers of 15a,a’ were observed at 115.8 (40%) and 120.0 ppm (60%). These values are close to those reported for a 2,5-dihydro-1,2-benzoxaphosphole-2-sulfide (130.2 ppm [34]).
Under the same reaction conditions, but with a slightly longer reaction time (2 d instead of 1 d), 4a,a’ was treated with elemental (gray) selenium. The 1,2-oxaphosphetane-P-selenides 16a,a’ (Scheme 6) were formed in a selective manner and isolated as 29:71 mixture in good yields by filtration, and excess selenium was removed. The 16a,a’ mixture was fully characterized by NMR spectroscopy and LIFDI mass spectrometry. Its 31P{1H}-NMR spectrum showed two resonance signals with selenium satellites at 116.1 ppm (1J(Se,P) = 839.7 Hz) and 121.5 ppm (1J(Se,P) = 846.4 Hz), corresponding to the two diastereomers of 16a,a’. The 77Se{1H}-NMR spectrum showed two doublets at −10.7 and 79.4 ppm. A comparison to the (acyclic) tert-butyl-ethoxyphenylphosphane-P-selenide[33] (δ (31P{1H}) = 111.0 ppm, 1J(Se,P) = 786.3 Hz) showed very similar values for the phosphorus chemical shifts and coupling constants, whereas the selenium resonances of 16a,a’ are downfield-shifted (cf. δ (77Se{1H}) = −350.3 ppm [35]).
However, 4a,a’ did not react with elemental tellurium or tributylphosphane-P-telluride (as transfer reagent) to form 1,2-oxaphosphetane-P-tellurides under the same conditions, nor by heating to 80 °C.
A comparison of the 13C{1H}-NMR data of 4a,a’, 14a,a’, 15a,a’, and 16a,a’ (Table 4) reveals, for all chalcogenides, the 1J(P,CH2) and 2J(P,CH) coupling constants are increased as expected when oxidizing P(III) to P(V). In the case of 14a,a’, the 1J(P,CPh3) constant also increases, whereas it decreases when going from 4a,a’ to 15a,a’ and then to 16a,a’. The decrease in the coupling constant hints at a change of the hybridization of phosphorus and, concomitantly, the bond angles due to increased steric bulk near the ring system.

2.4. Ring Strain Energy of Model 1,2-Oxaphosphetane Derivatives

To obtain further insight into the chemistry of the 1,2-oxaphosphirane chalcogenides, their ring strain energies (RSEs) were computed for model 1,2-oxaphosphetane derivatives VIa-e (Scheme 7) using suitable homodesmotic reactions (see ESI), as previously completed for related three- and four-membered heterocycles [36,37,38,39,40,41,42,43,44]. RSEs values (Table 5) slightly increase in the order VIa < VIb < VIc < VId < VIe, therefore regularly increasing for heavier P-chalcogenides and also reproducing the reported variation on moving from the σ3λ3-1,2-oxaphosphetane VIa to its P-oxide derivative VIb [22].

3. Materials and Methods

3.1. Synthetic Details

The syntheses of all compounds were performed under an argon atmosphere, using common Schlenk techniques and dry solvents. All NMR spectra were recorded on a Bruker AVI-300 or a Bruker AV III HD Prodigy 500 spectrometer at 25 °C. The 1H and 13C NMR spectra were referenced to the residual proton resonances and the 13C NMR signals of the deuterated solvents and 31P to 85% H3PO4 as external standards, respectively. Please check the ESI for further experimental details.

3.2. Computational Details

DFT calculations were performed with the ORCA electronic structure program package (version 4.2.1, created by Frank Neese, Max Planck Institut für Kohlenforschung, Mülheim/Ruhr, Germany) [45]. All geometry optimizations were run in redundant internal coordinates with tight convergence criteria, in the gas phase, and using Grimme’s dispersion-corrected composite PBEh-3c level [46]. For the mechanistic study, solvent (THF) effects were taken into consideration with the CPCM solvation method [47] as implemented in ORCA. For Mo [48] and Te [49] atoms, the [def2-ECP(28)] effective core potential (ECP) was used. Harmonic frequency calculations verified the nature of ground states or transition states (TS), having all positive frequencies or only one imaginary frequency, respectively. TS structures were confirmed by following the intrinsic reaction path in both directions of the negative eigenvector. From these optimized geometries, all reported data were obtained by means of single-point (SP) calculations using the more polarized def2-TZVPP basis set [50]. Reported energies include the Zero-point energy (ZPE) correction term at the optimization level. In the case of mechanistic aspects, final energies were obtained by means of double-hybrid-meta-GGA functional PWPB95 [51,52], using the RI [53,54,55] approximation for the MP2 correlation part, together with the RI-JK approximation for Coulomb and exchange integrals in the DFT part. Additionally, the latest Grimme’s semiempirical atom-pair-wise London dispersion correction D4 was included [56]. For RSE calculations, final energies were computed with the near-linear scaling domain-based local pair natural orbital (DLPNO) [57] method, to achieve coupled cluster theory with single-double and perturbative triple excitations (CCSD(T)) [58] using the def2-TZVPP basis set.

4. Conclusions

The new unligated 1,2σ3λ3-oxaphosphetanes 4a,a’d,d’ were synthesized and characterized. As a good case in point, a mixture of 4a,a’ was used to investigate oxidation reactions, i.e., to access less (Ch = O) and/or unknown (Ch = S, Se) 1,2σ4λ5-oxaphosphetane P-chalcogenides. DFT calculations provided mechanistic insights into the formation of C-phenyl-substituted 1,2σ3λ3-oxaphosphetanes 3e,e’ and 3e*,e*’ using model derivatives. Nucleophilic attack to non-benzylic carbon of styrene oxide 2e followed by cyclization seems to be the preferred pathway which explains the preferred formation of 3e,e’. Ring strain energy calculations revealed the tendency to increase RSE values in going from 1,2σ3λ3- to 1,2σ4λ5-oxaphosphetane P-chalcogenides and among the latter from the lighter to the heavier chalcogens.

Supplementary Materials

The following supporting information can be downloaded at: https://0-www-mdpi-com.brum.beds.ac.uk/article/10.3390/molecules27103345/s1. General procedures, synthetic methods, and analytical data for 3a,a’, 3b,b’, 3c,c’, 3d,d’, 3e,e’, 4b,b’, 4c,c’, 4d,d’, 14a,a’, 15a,a’, and 16a,a’; Crystal structure data of 3b,b’, 3c,c’, 3d,d’, 4b,b’, and 4c,c’; Figures S1–S9: 1H-, 13C{1H}-, and 31P[1H}-NMR spectra of 3b,b’-3d,d’; Figure S10: 31P[1H}-NMR spectrum of 3e,e’ and 3e*,e*’; Figures S11–S19: 1H-, 13C{1H}-, and 31P[1H}-NMR spectra of 4b,b’-4d,d’; Figures S20–S22: 1H-, 13C{1H}-, and 31P[1H}-NMR spectra of 14a,a’; Figures S23–S25: 1H-, 13C{1H}-, and 31P[1H}-NMR spectra of 15a,a’; Figures S26–S29: 1H-, 13C{1H}-, 31P[1H}-, 77Se{1H}-NMR spectra of 16a,a’; Table S1: P-O bond properties of complexes 9 and 9OMe2; Table S2: C-O bond properties of 2e, 2e-Li(DEGDME) and 9; Scheme S1: Mechanistic proposal for the formation of 1,2-oxaphosphetanes 12′ and 12b*’ from the direct nucleophilic attack of the P atom to the epoxide C atoms of 7; Figure S30: Calculated (CPCMTHF/CCSD(T)/def2-TZVPP(ecp)) minimum energy profile for the conversion of 7 into 12′ and 12*’; Scheme S2: Nucleophilic attack of the P atom of 5tBu to C atoms of styrene oxide 2e giving rise to 13tBu and 13tBu*; Figure S31: Calculated (CPCMTHF/CCSD(T)/def2-TZVPP(ecp)) minimum energy profile for the conversion of 5tBu + 2e into 13tBu and 13tBu*; Scheme S3: Homodesmotic reactions for RSE evaluation on derivatives VIa-e; Calculated structures.

Author Contributions

Conceptualization, A.E.F. and R.S.; methodology, F.G. and A.G.A.; validation, A.E.F. and R.S.; formal analysis, investigation, F.G. and A.G.A.; resources, R.S.; data curation, A.E.F. and R.S.; writing—review and editing, F.G., G.S., A.E.F. and R.S.; supervision, A.E.F. and R.S. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data presented in this study are available in supplementary material.

Acknowledgments

We are grateful to the University of Bonn for financial support (R.S.), to the Servicio de Cálculo Científico at the University of Murcia for technical support and computational resources (A.E.F.) and the University of Murcia for a predoctoral fellowship (A.G.A). The authors would also like to acknowledge A.C. Filippou and D. Menche for the use of X-ray facilities and C. Rödde for the X-ray diffraction studies.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. He, G.; Shynkaruk, O.; Lui, M.W.; Rivard, E. Small Inorganic Rings in the 21st Century: From Fleeting Intermediates to Novel Isolable Entities. Chem. Rev. 2014, 114, 7815–7880. [Google Scholar] [CrossRef] [PubMed]
  2. Bull, J.A.; Croft, R.A.; Davis, O.A.; Doran, R.; Morgan, K.F. Oxetanes: Recent Advances in Synthesis, Reactivity, and Medicinal Chemistry. Chem. Rev. 2016, 116, 12150–12233. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  3. Klein, R.; Wurm, F.R. Aliphatic Polyethers: Classical Polymers for the 21st Century. Macromol. Rapid Commun. 2015, 36, 1147–1165. [Google Scholar] [CrossRef] [PubMed]
  4. Marinetti, A.; Carmichael, D. Synthesis and Properties of Phosphetanes. Chem. Rev. 2002, 102, 201–230. [Google Scholar] [CrossRef]
  5. Zhao, W.; Yan, P.K.; Radosevich, A.T. A Phosphetane Catalyzes Deoxygenative Condensation of α-Keto Esters and Carboxylic Acids via PIII/PV=O Redox Cycling. J. Am. Chem. Soc. 2015, 137, 616–619. [Google Scholar] [CrossRef]
  6. Reichl, K.D.; Dunn, N.L.; Fastuca, N.J.; Radosevich, A.T. Biphilic Organophosphorus Catalysis: Regioselective Reductive Transposition of Allylic Bromides via PIII/PV Redox Cycling. J. Am. Chem. Soc. 2015, 137, 5292–5295. [Google Scholar] [CrossRef] [Green Version]
  7. Nykaza, T.V.; Harrison, T.S.; Ghosh, A.; Putnik, R.A.; Radosevich, A.T. A Biphilic Phosphetane Catalyzes N–N Bond-Forming Cadogan Heterocyclization via PIII/PV=O Redox Cycling. J. Am. Chem. Soc. 2017, 139, 6839–6842. [Google Scholar] [CrossRef] [Green Version]
  8. Kaboudin, B.; Haghighat, H.; Yokomatsu, T. Synthesis of a New Class of Phosphinic Acids: Synthesis of Novel Four-Membered Cyclic Oxaphosphetanes by Intramolecular Mitsunobu Reaction of Bis(α-hydroxyalkyl)phosphinic Acids. Synthesis 2011, 2011, 3185–3189. [Google Scholar] [CrossRef]
  9. Wittig, G.; Geissler, G. Zur Reaktionsweise des Pentaphenyl-phosphors und einiger Derivate. Justus Liebigs Ann. Chem. 1953, 580, 44–57. [Google Scholar] [CrossRef]
  10. Hinz, A.; Labbow, R.; Rennick, C.; Schulz, A.; Goicoechea, J.M. HPCO—A Phosphorus-Containing Analogue of Isocyanic Acid. Angew. Chem. Int. Ed. 2017, 56, 3911–3915. [Google Scholar] [CrossRef] [Green Version]
  11. Wittig, G.; Schöllkopf, U. Über Triphenyl-phosphin-methylene als olefinbildende Reagenzien (I. Mitteil). Chem. Ber. 1954, 87, 1318–1330. [Google Scholar] [CrossRef]
  12. Vedejs, E.; Meier, G.P.; Snoble, K.A.J. Low-temperature characterization of the intermediates in the Wittig reaction. J. Am. Chem. Soc. 1981, 103, 2823–2831. [Google Scholar] [CrossRef]
  13. Wittig, G.; Haag, W. Über Triphenyl-phosphinmethylene als olefinbildende Reagenzien (II. Mitteil.1). Chem. Ber. 1955, 88, 1654–1666. [Google Scholar] [CrossRef]
  14. Espinosa Ferao, A. On the Mechanism of Trimethylphosphine-Mediated Reductive Dimerization of Ketones. Inorg. Chem. 2018, 57, 8058–8064. [Google Scholar] [CrossRef]
  15. Dieckbreder, U.; Lork, E.; Röschenthaler, G.-V.; Kolomeitsev, A.A. First P-trifluoromethylated ylides. Heteroat. Chem. 1996, 7, 281–284. [Google Scholar] [CrossRef]
  16. Hamaguchi, M.; Iyama, Y.; Mochizuki, E.; Oshima, T. First isolation and characterization of 1,2-oxaphosphetanes with three phenyl groups at the phosphorus atom in typical Wittig reaction using cyclopropylidenetriphenylphosphorane. Tetrahedron Lett. 2005, 46, 8949–8952. [Google Scholar] [CrossRef]
  17. Kawashima, T.; Kato, K.; Okazaki, R. Synthesis, Structure, and Thermolysis of a 3-Methoxycarbonyl-1, 2 λ5-oxaphosphetane. Angew. Chem. Int. Ed. Engl. 1993, 32, 869–870. [Google Scholar] [CrossRef]
  18. Caughlan, C.N.; Ramirez, F.; Pilot, J.F.; Smith, C.P. Crystal and molecular structure of a four-membered cyclic oxyphosphorane with pentavalent phosphorus, PO2(C6H5)2(CF3)4C3H2. J. Am. Chem. Soc. 1971, 93, 5229–5235. [Google Scholar] [CrossRef]
  19. Dianova, E.N.; Zabotina, E.Y.; Akhmethkhaova, I.Z.; Samuilov, Y.D. 5-Methyl-2-phenyl-1,2,3-diazaphosphole in reaction with 2,5-bis(methoxycarbonyl)-3,4-diphenylcyclopentadiene. Zh. Obs. Khim. 1991, 64, 1063–1066. [Google Scholar]
  20. Kyri, A.W.; Schnakenburg, G.; Streubel, R. A novel route to C-unsubstituted 1,2-oxaphosphetane and 1,2-oxaphospholane complexes. Chem. Commun. 2016, 52, 8593–8595. [Google Scholar] [CrossRef]
  21. Kyri, A.W. Investigations on 1,2-Oxaphosphetane Complexes. Ph.D. Thesis, University of Bonn, Bonn, Germany, 2017. [Google Scholar]
  22. Kyri, A.W.; Gleim, F.; García Alcaraz, A.; Schnakenburg, G.; Espinosa Ferao, A.; Streubel, R. “Low-coordinate” 1,2-oxaphosphetanes—A new opportunity in coordination and main group chemistry. Chem. Commun. 2018, 54, 7123–7126. [Google Scholar] [CrossRef] [PubMed]
  23. Espinosa Ferao, A.; Deschamps, B.; Mathey, F. Towards functional phospholide ions: Synthesis of a 3-ethoxycarbonyl derivative. Bull. Soc. Chim. Fr. 1993, 130, 695–699. [Google Scholar]
  24. Regitz, M.; Scherer, H.; Illger, W.; Eckes, H. Detection of Alkylideneoxophosphoranes by Cycloaddition of Carbonyl Compounds. Angew. Chem. Int. Ed. Engl. 1973, 12, 1010–1011. [Google Scholar] [CrossRef]
  25. Kawashima, T.; Inamoto, N. The Correct Structure of Cyclic Adducts of (Diphenylmethylene)oxophenylphosphorane with Aromatic Aldehydes. Bull. Chem. Soc. Jpn. 1991, 64, 713–715. [Google Scholar] [CrossRef] [Green Version]
  26. Nakayama, S.; Yoshifuji, M.; Okazaki, R.; Inamoto, N. Phosphinidenes and Related Intermediates. V. Reactions of Phosphinylidenes with cis- and trans-Stilbene Oxides. Bull. Chem. Soc. Jpn. 1976, 49, 1173–1174. [Google Scholar] [CrossRef] [Green Version]
  27. Kawashima, T.; Nakayama, S.; Yoshifuji, M.; Okazaki, R.; Inamoto, N. Revised Structure of 2-Cyclohexyl-3,4-diphenyl-1,2-oxaphosphetane 2-Oxide. Bull. Chem. Soc. Jpn. 1991, 64, 711–712. [Google Scholar] [CrossRef]
  28. Hafez, T.S.; El-Khoshnieh, Y.O.; Mahran, M.R.; Atta, S.M.S. Organophosphorus Chemistry, 20.1 the Behaviour of Certain γ-Pyrone Derivatives toward 2,4-BIS-(4-Methoxyphenyl)-1,3,2,4-Dithiaphosphetan-2,4-Disulphide (Lawesson Reagent). Phosphorus. Sulfur. Silicon Relat. Elem. 1991, 56, 165–171. [Google Scholar] [CrossRef]
  29. Kawashima, T.; Takami, H.; Okazaki, R. Synthesis and Thermolysis of Tetracoordinate 1,2-Oxaphosphetanes Stabilized by Steric Protection. Chem. Lett. 1994, 23, 1487–1490. [Google Scholar] [CrossRef]
  30. Dabrowska, U.; Dabrowski, J. IR-Spektren und Struktur von α-β-ungesättigten Carbonylverbindungen, XXI. Thione aus der Reaktion von Alkylaminocrotonestern mit Phosphorpentasulfid. Chem. Ber. 1976, 109, 1779–1786. [Google Scholar] [CrossRef]
  31. Kyri, A.W.; Nesterov, V.; Schnakenburg, G.; Streubel, R. Synthesis and Reaction of the First 1,2-Oxaphosphetane Complexes. Angew. Chem. Int. Ed. 2014, 53, 10809–10812. [Google Scholar] [CrossRef]
  32. Espinosa Ferao, A.; García Alcaraz, A.; Zaragoza Noguera, S.; Streubel, R. Terminal Phosphinidene Complex Adducts with Neutral and Anionic O-Donors and Halides and the Search for a Differentiating Bonding Descriptor. Inorg. Chem. 2020, 59, 12829–12841. [Google Scholar] [CrossRef] [PubMed]
  33. Espinosa Ferao, A.; Streubel, R. 1,2-Thiaphosphetanes: The Quest for Wittig-Type Ring Cleavage, Rearrangement, and Sulfur Atom Transfer. Inorg. Chem. 2020, 59, 3110–3117. [Google Scholar] [CrossRef] [PubMed]
  34. Kojima, S.; Sugino, M.; Matsukawa, S.; Nakamoto, M.; Akiba, K. First Isolation and Characterization of an Anti-Apicophilic Spirophosphorane Bearing an Oxaphosphetane Ring:  A Model for the Possible Reactive Intermediate in the Wittig Reaction. J. Am. Chem. Soc. 2002, 124, 7674–7675. [Google Scholar] [CrossRef] [PubMed]
  35. Kimura, T.; Murai, T. P-Chiral Phosphinoselenoic Chlorides and Phosphinochalcogenoselenoic Acid Esters:  Synthesis, Characterization, and Conformational Studies. J. Org. Chem. 2005, 70, 952–959. [Google Scholar] [CrossRef] [PubMed]
  36. Rey, A.; Espinosa Ferao, A.; Streubel, R. Quantum Chemical Calculations on CHOP Derivatives—Spanning the Chemical Space of Phosphinidenes, Phosphaketenes, Oxaphosphirenes, and COP–Isomers. Molecules 2018, 23, 3341. [Google Scholar] [CrossRef] [Green Version]
  37. Espinosa Ferao, A. Kinetic energy density per electron as quick insight into ring strain energies. Tetrahedron Lett. 2016, 57, 5616–5619. [Google Scholar] [CrossRef]
  38. Streubel, R.; Faßbender, J.; Schnakenburg, G.; Espinosa Ferao, A. Formation of Transient and Stable 1,3-Dipole Complexes with P,S,C and S,P,C Ligand Skeletons. Organometallics 2015, 34, 3103–3106. [Google Scholar] [CrossRef]
  39. Villalba Franco, J.M.; Schnakenburg, G.; Sasamori, T.; Espinosa Ferao, A.; Streubel, R. Stimuli-Responsive Frustrated Lewis-Pair-Type Reactivity of a Tungsten Iminoazaphosphiridine Complex. Chem.—A Eur. J. 2015, 21, 9650–9655. [Google Scholar] [CrossRef]
  40. Albrecht, C.; Schneider, E.; Engeser, M.; Schnakenburg, G.; Espinosa, A.; Streubel, R. Synthesis and DFT calculations of spirooxaphosphirane complexes. Dalt. Trans. 2013, 42, 8897–8906. [Google Scholar] [CrossRef]
  41. Espinosa, A.; Gómez, C.; Streubel, R. Single Electron Transfer-Mediated Selective endo- and exocyclic Bond Cleavage Processes in Azaphosphiridine Chromium(0) Complexes: A Computational Study. Inorg. Chem. 2012, 51, 7250–7256. [Google Scholar] [CrossRef]
  42. Schulten, C.; von Frantzius, G.; Schnakenburg, G.; Streubel, R. Deoxygenation of isocyanates via transient electrophilic terminal phosphinidene complexes: Are strained P-heterocycles involved? Heteroat. Chem. 2011, 22, 275–286. [Google Scholar] [CrossRef]
  43. Espinosa, A.; Streubel, R. Computational Studies on Azaphosphiridines, or How to Effect Ring-Opening Processes through Selective Bond Activation. Chem.—A Eur. J. 2011, 17, 3166–3178. [Google Scholar] [CrossRef] [PubMed]
  44. Krahe, O.; Neese, F.; Streubel, R. The Quest for Ring Opening of Oxaphosphirane Complexes: A Coupled-Cluster and Density Functional Study of CH3PO Isomers and Their Cr(CO)5 Complexes. Chem. Eur. J. 2009, 15, 2594–2601. [Google Scholar] [CrossRef] [PubMed]
  45. Neese, F. The ORCA program system. Wiley Interdiscip. Rev. Comput. Mol. Sci. 2012, 2, 73–78. [Google Scholar] [CrossRef]
  46. Grimme, S.; Brandenburg, J.G.; Bannwarth, C.; Hansen, A. Consistent structures and interactions by density functional theory with small atomic orbital basis sets. J. Chem. Phys. 2015, 143, 54107. [Google Scholar] [CrossRef]
  47. Barone, V.; Cossi, M. Quantum Calculation of Molecular Energies and Energy Gradients in Solution by a Conductor Solvent Model. J. Phys. Chem. A 1998, 102, 1995–2001. [Google Scholar] [CrossRef]
  48. Andrae, D.; Häußermann, U.; Dolg, M.; Stoll, H.; Preuß, H. Energy-adjustedab initio pseudopotentials for the second and third row transition elements. Theor. Chim. Acta 1990, 77, 123–141. [Google Scholar] [CrossRef]
  49. Peterson, K.A.; Figgen, D.; Goll, E.; Stoll, H.; Dolg, M. Systematically convergent basis sets with relativistic pseudopotentials. II. Small-core pseudopotentials and correlation consistent basis sets for the post-d group 16–18 elements. J. Chem. Phys. 2003, 119, 11113–11123. [Google Scholar] [CrossRef] [Green Version]
  50. Schäfer, A.; Huber, C.; Ahlrichs, R. Fully optimized contracted Gaussian basis sets of triple zeta valence quality for atoms Li to Kr. J. Chem. Phys. 1994, 100, 5829–5835. [Google Scholar] [CrossRef]
  51. Goerigk, L.; Grimme, S. A thorough benchmark of density functional methods for general main group thermochemistry, kinetics, and noncovalent interactions. Phys. Chem. Chem. Phys. 2011, 13, 6670–6688. [Google Scholar] [CrossRef]
  52. Goerigk, L.; Grimme, S. Efficient and Accurate Double-Hybrid-Meta-GGA Density Functionals—Evaluation with the Extended GMTKN30 Database for General Main Group Thermochemistry, Kinetics, and Noncovalent Interactions. J. Chem. Theory Comput. 2011, 7, 291–309. [Google Scholar] [CrossRef] [PubMed]
  53. Weigend, F.; Häser, M.; Patzelt, H.; Ahlrichs, R. RI-MP2: Optimized auxiliary basis sets and demonstration of efficiency. Chem. Phys. Lett. 1998, 294, 143–152. [Google Scholar] [CrossRef]
  54. Weigend, F.; Häser, M. RI-MP2: First derivatives and global consistency. Theor. Chem. Acc. 1997, 97, 331–340. [Google Scholar] [CrossRef]
  55. Bernholdt, D.E.; Harrison, R.J. Large-scale correlated electronic structure calculations: The RI-MP2 method on parallel computers. Chem. Phys. Lett. 1996, 250, 477–484. [Google Scholar] [CrossRef] [Green Version]
  56. Caldeweyher, E.; Bannwarth, C.; Grimme, S. Extension of the D3 dispersion coefficient model. J. Chem. Phys. 2017, 147, 34112. [Google Scholar] [CrossRef] [PubMed]
  57. Riplinger, C.; Sandhoefer, B.; Hansen, A.; Neese, F. Natural triple excitations in local coupled cluster calculations with pair natural orbitals. J. Chem. Phys. 2013, 139, 134101. [Google Scholar] [CrossRef]
  58. Pople, J.A.; Head-Gordon, M.; Raghavachari, K. Quadratic configuration interaction. A general technique for determining electron correlation energies. J. Chem. Phys. 1987, 87, 5968–5975. [Google Scholar] [CrossRef]
Figure 1. Oxetane (I), phosphetanes (σ3λ3 IIa, σ5λ5 IIb), 1,3-oxaphosphetanes (σ3λ3 IIIa, σ5λ5 IIIb), 1,2-oxaphosphetanes (σ3λ3 IVa, σ5λ5 IVb), 1,2σ3λ3-oxaphosphetane metal complexes (V), and 1,2σ4λ5-oxaphosphetanes (VI).
Figure 1. Oxetane (I), phosphetanes (σ3λ3 IIa, σ5λ5 IIb), 1,3-oxaphosphetanes (σ3λ3 IIIa, σ5λ5 IIIb), 1,2-oxaphosphetanes (σ3λ3 IVa, σ5λ5 IVb), 1,2σ3λ3-oxaphosphetane metal complexes (V), and 1,2σ4λ5-oxaphosphetanes (VI).
Molecules 27 03345 g001
Scheme 1. Synthesis of complexed (3a-d) and unligated 1,2-oxaphosphetanes (4a-d).
Scheme 1. Synthesis of complexed (3a-d) and unligated 1,2-oxaphosphetanes (4a-d).
Molecules 27 03345 sch001
Figure 2. Molecular structures of 1,2-oxaphosphetane 4b,b’ in the solid state. Hydrogen atoms are omitted and the thermal ellipsoids are set at the 50% probability level. Split layers C2A:C2 equals 33:67, C2′A:C2′ equals 45:55. Selected bond lengths in Å, angles in degree, second entry corresponds to the crystallographic positions denoted with a dash: P-O 1.6792(14)/1.6803(13), P-C1 1.854(2)/1.849(2), P-C6 1.9229(19)/1.9148(18), O-C2 1.538(3)/1.542(3), O-C2A 1.451(6)/1.477(4), C1-C2 1.562(3)/1.599(4), C1-C2A 1.609(7)/1.552(5), O-P-C1 79.93(8)/80.07(8), C2-O-P 96.67(13)/97.70(13), C2A-O-P 88.0(3), 88.04(19), O-C2-C1 94.35(18)/92.7(2), O-C2A-C1 95.8(4)/97.3(3), C2-C1-P 89.05(14)/89.23(14), and C2A-C1-P 77.6(2)/80.03(19).
Figure 2. Molecular structures of 1,2-oxaphosphetane 4b,b’ in the solid state. Hydrogen atoms are omitted and the thermal ellipsoids are set at the 50% probability level. Split layers C2A:C2 equals 33:67, C2′A:C2′ equals 45:55. Selected bond lengths in Å, angles in degree, second entry corresponds to the crystallographic positions denoted with a dash: P-O 1.6792(14)/1.6803(13), P-C1 1.854(2)/1.849(2), P-C6 1.9229(19)/1.9148(18), O-C2 1.538(3)/1.542(3), O-C2A 1.451(6)/1.477(4), C1-C2 1.562(3)/1.599(4), C1-C2A 1.609(7)/1.552(5), O-P-C1 79.93(8)/80.07(8), C2-O-P 96.67(13)/97.70(13), C2A-O-P 88.0(3), 88.04(19), O-C2-C1 94.35(18)/92.7(2), O-C2A-C1 95.8(4)/97.3(3), C2-C1-P 89.05(14)/89.23(14), and C2A-C1-P 77.6(2)/80.03(19).
Molecules 27 03345 g002
Scheme 2. Synthesis of C-phenyl substituted 1,2-oxaphosphetane complexes 3e,e’ and 3e*,e*’.
Scheme 2. Synthesis of C-phenyl substituted 1,2-oxaphosphetane complexes 3e,e’ and 3e*,e*’.
Molecules 27 03345 sch002
Scheme 3. Reaction of model Li/Cl phosphinidenoid complex 5 with (S)-styrene oxide 2e giving rise to 9 and its evolution through C-O bond cleavages. Computed ZPE-corrected energies (kcal mol−1) for both minima and TS (marked with a ‡ superscript) at the CPCMTHF/CCSD(T)/def2-TZVPP(ecp) level, in brackets.
Scheme 3. Reaction of model Li/Cl phosphinidenoid complex 5 with (S)-styrene oxide 2e giving rise to 9 and its evolution through C-O bond cleavages. Computed ZPE-corrected energies (kcal mol−1) for both minima and TS (marked with a ‡ superscript) at the CPCMTHF/CCSD(T)/def2-TZVPP(ecp) level, in brackets.
Molecules 27 03345 sch003
Scheme 4. Mechanistic proposal for the formation of 1,2-oxaphosphetanes 12 and 12b* from the direct nucleophilic attack of the P atom to the epoxide C atoms of 7.
Scheme 4. Mechanistic proposal for the formation of 1,2-oxaphosphetanes 12 and 12b* from the direct nucleophilic attack of the P atom to the epoxide C atoms of 7.
Molecules 27 03345 sch004
Figure 3. Calculated (CPCMTHF/CCSD(T)/def2-TZVPP(ecp)) minimum energy profile for the conversion of 7 into 12 and 12*.
Figure 3. Calculated (CPCMTHF/CCSD(T)/def2-TZVPP(ecp)) minimum energy profile for the conversion of 7 into 12 and 12*.
Molecules 27 03345 g003
Scheme 5. Synthesis of 1,2-oxaphosphetane P—oxides 14a,a’.
Scheme 5. Synthesis of 1,2-oxaphosphetane P—oxides 14a,a’.
Molecules 27 03345 sch005
Scheme 6. Synthesis of heavier 1,2-oxaphosphetane P—chalcogenides 15a,a’16a,a’.
Scheme 6. Synthesis of heavier 1,2-oxaphosphetane P—chalcogenides 15a,a’16a,a’.
Molecules 27 03345 sch006
Scheme 7. Model 1,2-oxaphosphetane-P-chalcogenides VIa-e studied computationally.
Scheme 7. Model 1,2-oxaphosphetane-P-chalcogenides VIa-e studied computationally.
Molecules 27 03345 sch007
Table 1. Selected NMR spectroscopic data of 3a,a’d,d’ measured in CDCl3, chemical shifts in ppm, and coupling constants in Hz. See Supplementary Materials for more data and spectra.
Table 1. Selected NMR spectroscopic data of 3a,a’d,d’ measured in CDCl3, chemical shifts in ppm, and coupling constants in Hz. See Supplementary Materials for more data and spectra.
3a,a’3b,b’3c,c’3d,d’
Ratio51:4950:5050:5048:52
δ(P)185.6/207.2183.5/206.1187.6/208.8187.3/208.7
δ(CH2)2.92/2.962.90/2.352.91/2.952.90/2.98
δ(CH2*)3.01/3.182.90/2.993.00/3.122.98/3.13
δ(CH)5.35/4.694.54/4.974.48/5.135.12/4.48
δ(CH2)41.7/39.937.5/39.940.6/38.538.5/40.5
1J(P-C)18.3/21.721.5/18.518.3/21.421.5/8.3
δ(CPh3)68.0/67.267.1/67.867.2/67.967.1/67.9
1J(P-C)10.6/9.09.3/10.79.3/10.79.4/10.7
δ(CH)77.7/82.085.9/90.180.9/85.081.1/85.2
2J(P-C)11.7/11.711.3/11.211.6/11.411.6/11.5
* Denotes second set of the magnetically non-equivalent CH2-protons arising from C4-substituted regioisomers.
Table 2. 31P{1H}-NMR spectroscopic data of 4a,a’d,d’, chemical shifts in ppm, coupling constants in Hz.
Table 2. 31P{1H}-NMR spectroscopic data of 4a,a’d,d’, chemical shifts in ppm, coupling constants in Hz.
4a,a’4b,b’4c,c’4d,d’
Ratio42:5834:6639:6118:82
δ(P)163.7/199.0 1161.2/196.7 1166.2/199.0 2166.3/199.4 3
1 Measured in CDCl3. 2 Measured in C6D6. 3 Measured in n-pentane.
Table 3. Selected NMR spectroscopic data of 3e,e’ and 3e*,e*’, measured in the reaction solution, chemical shifts in ppm, and coupling constants in Hz.
Table 3. Selected NMR spectroscopic data of 3e,e’ and 3e*,e*’, measured in the reaction solution, chemical shifts in ppm, and coupling constants in Hz.
3e3e’3e*3e*’
Ratio4045105
δ(P)191.5210.9237.3244.9
δ(CH2)39.740.475.976.2
nJ(P-C)18.821.713.113.6
Table 4. Selected NMR spectroscopical data of 4a,a’, 14a,a’, 15a,a’, and 16a,a’, chemical shifts in ppm, and coupling constants in Hz. See Supplementary Materials for more data and spectra.
Table 4. Selected NMR spectroscopical data of 4a,a’, 14a,a’, 15a,a’, and 16a,a’, chemical shifts in ppm, and coupling constants in Hz. See Supplementary Materials for more data and spectra.
4a,a’ 114a,a’ 115a,a’ 216a,a’ 2
δ(P)163.7/199.062.1/63.5115.8/120.0116.1/121.5
δ(CH2)2.22/2.452.53/2.562.49/2.222.62/2.43
δ(CH2*)2.69/2.482.75/2.962.49/2.722.75/3.00
δ(CH)5.12/4.624.25/5.044.78/4.054.88/4.26
δ(CH2)33.7/31.239.5/38.944.8/44.445.6/44.6
1J(P-C)13.6/7.960.8/64.051.5/49.844.6/43.5
δ(CPh3)63.4/62.965,6/66.270.4/69.870.4/69.8
1J(P-C)52.3/50.773.3/71.946.1/47.833.4/35.4
δ(CH)77.6/83.072.1/74.375.3/74.576.1/76.1
2J(P-C)4.6/2.220.2/20.219.4/19.919.4/19.4
1 Measured in CDCl3. 2 Measured in C6D6. * Denotes second set of the magnetically non-equivalent CH2-protons arising from C3-substituted regioisomers.
Table 5. Computed CCSD(T)/def2-TZVPP(ecp)//PBEh-3c RSEs (in kcal mol−1) for model compounds VIa-e.
Table 5. Computed CCSD(T)/def2-TZVPP(ecp)//PBEh-3c RSEs (in kcal mol−1) for model compounds VIa-e.
VIaVIbVicVidVie
18.9519.5719.9320.3520.59
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Gleim, F.; Alcaraz, A.G.; Schnakenburg, G.; Ferao, A.E.; Streubel, R. 1,2σ3λ3-Oxaphosphetanes and Their P-Chalcogenides—A Combined Experimental and Theoretical Study. Molecules 2022, 27, 3345. https://0-doi-org.brum.beds.ac.uk/10.3390/molecules27103345

AMA Style

Gleim F, Alcaraz AG, Schnakenburg G, Ferao AE, Streubel R. 1,2σ3λ3-Oxaphosphetanes and Their P-Chalcogenides—A Combined Experimental and Theoretical Study. Molecules. 2022; 27(10):3345. https://0-doi-org.brum.beds.ac.uk/10.3390/molecules27103345

Chicago/Turabian Style

Gleim, Florian, Antonio García Alcaraz, Gregor Schnakenburg, Arturo Espinosa Ferao, and Rainer Streubel. 2022. "1,2σ3λ3-Oxaphosphetanes and Their P-Chalcogenides—A Combined Experimental and Theoretical Study" Molecules 27, no. 10: 3345. https://0-doi-org.brum.beds.ac.uk/10.3390/molecules27103345

Article Metrics

Back to TopTop