Next Article in Journal
Clay Nanotubes Loaded with Diazepam or Xylazine Permeate the Brain through Intranasal Administration in Mice
Next Article in Special Issue
On the Efficiency of the Density Functional Theory (DFT)-Based Computational Protocol for 1H and 13C Nuclear Magnetic Resonance (NMR) Chemical Shifts of Natural Products: Studying the Accuracy of the pecS-n (n = 1, 2) Basis Sets
Previous Article in Journal
Pro- and Anti-Inflammatory Prostaglandins and Cytokines in Humans: A Mini Review
Previous Article in Special Issue
New pecJ-n (n = 1, 2) Basis Sets for Selenium Atom Purposed for the Calculations of NMR Spin–Spin Coupling Constants Involving Selenium
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

C- and N-Phosphorylated Enamines—An Avenue to Heterocycles: NMR Spectroscopy

A. E. Favorsky Irkutsk Institute of Chemistry, Siberian Branch of the Russian Academy of Sciences, 1 Favorsky St., 664033 Irkutsk, Russia
Int. J. Mol. Sci. 2023, 24(11), 9646; https://doi.org/10.3390/ijms24119646
Submission received: 13 April 2023 / Revised: 25 May 2023 / Accepted: 28 May 2023 / Published: 1 June 2023
(This article belongs to the Special Issue NMR Spectroscopy in Materials Chemistry)

Abstract

:
The review presents extensive data (from the works of the author and literature) on the structure of C- and N-chlorophosphorylated enamines and the related heterocycles obtained by multipulse multinuclear 1H, 13C, and 31P NMR spectroscopy. The use of phosphorus pentachloride as a phosphorylating agent for functional enamines enables the synthesis of various C- and N-phosphorylated products that are heterocyclized to form various promising nitrogen- and phosphorus-containing heterocyclic systems. 31P NMR spectroscopy is the most convenient, reliable and unambiguous method for the study and identification of organophosphorus compounds with different coordination numbers of the phosphorus atom, as well as for the determination of their Z- and E-isomeric forms. An alteration of the coordination number of the phosphorus atom in the phosphorylated compounds from 3 to 6 leads to a drastic screening of the 31P nucleus from about +200 to −300 ppm. The unique structural features of nitrogen–phosphorus-containing heterocyclic compounds are discussed.

Graphical Abstract

1. Introduction

The chemistry of organophosphorus compounds (OPCs) is one of the rapidly developing areas of organic chemistry. Phosphorus compounds have attracted much interest in the past decades due to their wide applications ranging from synthetic chemistry to materials science and life sciences. Interest in organophosphorus compounds is due to their importance for the chemistry of macromolecular materials, complexones, chelating agents, extragents, antipyrenes, drugs, phosphorus-containing pesticides and other materials [1,2,3,4,5,6,7,8,9]. Due to a wide range of physiological activity and easy degradation into the simplest non-toxic products, organophosphorus compounds rank among the best plant protection chemicals. Although organophosphorus chemistry is known as a mature and specialized field, researchers would like to develop new methods for the synthesis of OPCs to improve the safety and sustainability of chemical processes [10,11,12,13].
A well-known method for obtaining unsaturated OPCs having phosphorus chloride groups at the double bond is the reaction of phosphorus pentachloride with various organic nucleophiles—alkenes, alkadienes, alkynes, ethers and esters, tertiary amines and other compounds [14,15,16,17,18,19,20]. The availability of the starting reagents, mild reaction conditions and the ability to vary the structure of the final OPCs are advantages of this method. This reaction has not exhausted its synthetic capabilities in terms of the covering of new nucleophiles. Therefore, research in this area still remains a challenge.
Organic derivatives of phosphorus pentachloride in various coordination states, including the products of phosphorylation of alkenes and heterocycles based on them, can be conveniently studied by 31P NMR spectroscopy. As is known, phosphorus pentachloride, a common electrophilic phosphorylating agent, dissociates into ions both in the crystalline state and in polar solvents and exists in a pentacoordinated state in nonpolar solvents:
2   P C l 5 P C l 4 + + P C l 6
A variation in the coordination number of phosphorus from 3 to 6 significantly changes the shielding of the 31P nucleus from about +200 to −300 ppm:
Phosphorous chloride PCl 3 PCI 4 + PCl 5 PCl 6
Coordination number3456
31P NMR chemical shift, pm+200+80−80−300
The tetra-, penta- and hexacoordinated states of phosphorus pentachloride can be very easily characterized by multinuclear multipulse NMR spectroscopy and, in particular, by 31P NMR spectroscopy [9].
The phosphorylation reaction of a number of nitrogen-containing organic nucleophiles involves the electrophilic addition of phosphorus pentachloride to the double bond of enamines introduced into the reaction or formed during the phosphorylation of enamides, amides, ureides, tertiary amines and other nitrogen-containing organic nucleophiles. Phosphorus-containing enamines and enamides successfully combine a double bond and phosphorus and nitrogen organic fragments in a molecule. Due to the presence of several reaction centers, they are key compounds in the synthesis of promising functionalized heterocycles. In addition, organic compounds containing phosphorus and nitrogen at the same time are interesting objects for studying conjugation, stereochemistry and geometric isomerism.
The stereochemistry of functionalized enamines incorporating phosphorus atoms is a challenging topic in heteroatom chemistry because the correct interpretation of their chemical behavior and biological activity depends on understanding the factors that determine the stereochemical features and relative stability of their isomers.
The structural aspects of phosphorylated unsaturated compounds, such as N-vinylazoles and their derivatives, are considered in detail in the review [9]. It has been shown that many N-vinylazoles react with phosphorus pentachloride to form organophosphorus compounds containing tetra-, penta- and hexacoordinated phosphorus atoms.

2. The Phosphorylated Tertiary Amines

Tertiary amines are known to react intensely with phosphorus chlorides. This often leads to the resinification of the reaction mixture and does not allow the reaction products to be isolated as individual compounds. The reactions of tertiary amines with phosphorus chlorides deliver donor–acceptor complexes followed by the chlorination of alkyl groups, which in most cases leads to iminium salts with the N–C bond cleavage [21,22,23]. In the presence of excess tertiary amine, iminium salts are transformed into enamines, which can undergo C-phosphorylation with phosphorus pentachloride.
The phosphorylation of tertiary N-ethylamines with phosphorus pentachloride has been studied in detail by two-dimensional (2D) and multinuclear 1H, 13C and 31P NMR spectroscopy [14,24,25]. The tertiary amines, containing at least two ethyl groups (triethylamine, diethylaniline, diethylbenzylamine), react with phosphorus pentachloride under mild conditions (15–20 °C) to form hexachlorophosphorates of 2-aminoethenyltrichlorophosphonium (13) and hexachlorophosphorates of 2-amino-1-chloroethenyltrichlorophosphorates (46). The following sequence of transformations was suggested to explain the formation of compounds 16 (Scheme 1).
The reaction is initiated by the formation of donor–acceptor complexes of tertiary amines with phosphorus pentachlorides, followed by chlorination of the complex compounds at the methylene group in the α-position to the nitrogen atom. Due to the instability of tertiary chloramines, they lose chlorine in the form of an anion, which leads to the formation of iminium salts. In the presence of excess tertiary amine, iminium salts are dehydrochlorinated into enamines, the latter being phosphorylated with phosphorus pentachloride. Phosphorylated enamines (13), obtained from tertiary amines due to the high basicity of dialkyl- and alkyl(aryl)amino groups, have an increased nucleophilicity of the double bond and are easily chlorinated by phosphorus pentachloride at the double bond, in contrast to phosphorylated derivatives of N-vinylheterocycles, for example, N-vinylbenzotriazole and N-vinylcarbazole [9,26,27]. The chlorination products are easily dehydrochlorinated due to the high C-H acidity of protons (P–CH=) with the formation of hexachlorophosphorates 46 (Scheme 1, Table 1).
This was confirmed by studies of the reaction mixture by 31P NMR spectroscopy [24]. For example, two signals are observed in 31P NMR spectrum of the reaction products of phosphorus pentachloride with triethylamine: a doublet of doublets at 89.9 ppm 2JPH 35.0 Hz and 3JPH 25.9 Hz and a doublet at 80.7 ppm 3JPH 14.1 Hz assigned to organyltrichlorophosphonium cations 1 and 4 (Table 1).
The amount of hexachlorophosphorate 1 in the reaction mixture in the 31P NMR data decreases in time, while that of chlorenamine 4 grows. This testifies that enamine 4 is formed from hexachlorophosphate 1. The phosphorylation process is completed after the formation of enamine (4).
Phosphorylation of triethylamine in the presence of excess amine led to the arising in the reaction mixture of dichloroanhydride of 2-(diethylamino)-ethenylphosphonius acid (7), which was indicated by the presence of the signal at 160 ppm in the 31P NMR spectrum. Apparently, compound 7 is obtained through the reduction of 2-aminoethenyl-1-trichlorophosphonium hexachlorophosphate (1) with excess amine (Scheme 2).
It is not possible to separate crystalline hexachlorophosphates (16) from triorganylammonium chlorides presented in the reaction mixture, since the solubility of compounds 16 in organic solvents is close to that of triorganylammonium salts. This poses a serious preparative problem in the studies of the phosphorylation of tertiary amines with phosphorus pentachloride. Phosphorylation products of tertiary amines can be isolated in pure form by converting enaminotrichlorophosphonium hexachlorophosphates into phosphonic dichlorides 813, which is achieved by treating hexachlorophosphates with sulfur dioxide or dry acetone (Scheme 2). As hydrochlorides of tertiary amines are insoluble in diethyl ether, and chloroanhydrides of phosphonous acids are soluble, compounds 813 have been isolated in pure form by extraction with diethyl ether. The structure of phosphorus-containing enamines (16, 813) has been characterized by 1H, 13C and 31P NMR spectroscopy (Table 1).
The values of coupling constants 3JHH 13–14 Hz for the phosphorylated enamines having both vinyl protons indicate the trans-orientation of protons in the molecule. The 31P NMR spectra of phosphorylated chlorenamines are characterized by the coupling constant 3JPH values being decreased by 7–10 Hz as compared to phosphorylated enamines, containing no chlorine atom at the double bond. This fact is widely used to identify chlorophosphorylated compounds in organic synthesis.
The presence of a chlorine atom exactly at the carbon atom, nearest to the phosphorus-containing substituent in compounds 46 and 1113, is unambiguously confirmed by 13C NMR spectral data (Table 1). The high value of coupling constants 1JCH 164–168 Hz (N–CH=) in the 13C NMR spectra of these compounds can testify that the carbon atom is bound with the electronegative nitrogen atom of the amino group. The values of coupling constant 2JCP 35–40 Hz indicate the trans-(PN)-configuration (E-isomer) of hexachlorophosphorates of N,N-diorganylamino-1-chlorethenyl-phosphonium (46) and dichloroanhydride of N,N-diorganylamino-1-chloroethenylphosphonous acids (1113). A similar pattern (regularity) was observed in the phosphorylation products of N-vinylazoles and a number of other enamines [28]. A low value of 3JPH (11–26 Hz) (Table 1) indicates the cis-orientation of the phosphorus atom relative to the vinyl proton. In the case of the trans-orientation, 3JPH values reach 54–68 Hz [28].
The phosphorylation of N-methyl-N-ethylamine leads to the substitution of two chlorine atoms by the alkenyl groups in the PCl5 molecule to form a derivative of enaminophosphonium acids. Along with the reaction route leading to phosphorylated chloramine (14), the second direction is realized to afford chloroanhydride of bis-[2(N-methyl-N-phenylamino)ethyl] phosphonous acids (15) (Scheme 3).
In the 31P NMR spectrum, besides the doublet signal at δ 34.2 ppm with the value 3JPH = 11.8 Hz characteristic of chloranhydrides of 2-diorganylamino-1-chlorethenyl-phosphonous acid and related to compound 14, a signal at 35.6 ppm is present (Table 2). The splitting of this signal in a triplet of triplets unequivocally indicates the binding of the phosphorus atom with two vinyl groups in compound 15. NMR spectral data of oxaphosphinine (16) are given for comparison (Table 2). The chlorophosphoryl group (POCl) in 16 also is bound to two vinyl groups, but because of the heterocyclic nature of compound 16, vinyl protons are oriented in cis-position towards each other [29]. Coupling constant values 3JPH for compound 16 exceed twice those of compound 15, which correlates well with various geometry of structural fragments of these molecules containing a chlorophosphoryl group and vinyl proton. Similar conclusions follow from the comparison of 3JHH values for compounds 15 and 16 (12 Hz and 6.4 Hz, respectively).
The phosphorylation of N,N-diethylaniline in tetrachloromethane is accompanied (apart from phosphorylation) by the chlorination of organophosphorus compounds into the ring. In the 31P NMR spectrum of organyltrichlorophosphonium cations, four signals are observed. A signal at 79.8 ppm (dd) with 2JPH 34.3 and 3JPH 21.4 Hz is assigned to compound 3. Apparently, a high-frequency signal of smaller intensity (82.7 ppm), also split into a doublet of doublets with very close constants 2JPH 33.6 and 3JPH 22.9 Hz, belongs to compound 17 (Scheme 4). Chlorination of the ethyl group in hexachlorophosphorate (3) is improbable, as the nitrogen atom is bound to the acceptor ethyltrichlorophosphonium group. Electrophilic hydrogen substitution into the para-position of the phenyl ring is more likely because it is sterically less shielded.
Two doublets of different intensities are observed in the high field of the spectrum. The major signal at δ 72.0 ppm with 3JPH = 13.7 Hz is attributed to compound 6 and completely coincides with the characteristics of hexachlorophosphorate 6, which has been synthesized in benzene (Table 1). The minor doublet at δ 74.0 ppm with 3JPH = 14.5 Hz is assigned to compound 18. These parameters of the 31P NMR spectrum indicate close similarity of compounds 6 and 18.

3. The Structure of the Reaction Products of Triethylamine with Organyltrichlorophosphonium Hexachlorophosphorates

The most abundant phosphorylating agents are alkyl-, aryl- and tetrachloro(styryl)phosphoranes [14,30]. The interaction of triethylamine with organic derivatives of phosphorus pentachloride, organyltrichlorophosphonium hexachlorophosphorates, in particular, 2-ethoxyethenyltrichlorophosphonium hexachlorophosphorate, has been studied by NMR spectroscopy [31]. The doublet signal at 40.4 ppm (3JPH = 18.3 Hz) in the 31P NMR spectrum refers to the phosphorus atom in 1-chloro-2-ethoxyethenylphosphonyl dichloride (19). The appearance of a doublet at 36–40 ppm in the 31P NMR spectrum instead of a doublet of doublets is unexpected. Evidently, one vinyl hydrogen atom is replaced by the chlorine atom. The chlorination of the double bond followed by dehydrochlorination in the presence of triethylamine indicates the high electrophilicity of the chlorine atom in the donor–acceptor complexes formed between phosphorus pentachloride and triethylamine. It is most likely that these complexes are able to dissociate partially to ions (Scheme 5). The chlorine atom in the formed cation is electrophilic enough for the chlorination of a multiple bond. 1-Chloro-2-ethoxyethenylphosphonyl dichloride (19) is formed, most likely, according to Scheme 5.
The appearance of the dark orange color of the reaction mixture can be due to the partial dehydrochlorination of compound 19 with triethylamine to 2-ethoxyethynylphosphonyl chloride (20), which is unstable and undergoes easy resinification (Scheme 5) (Table 3). The intense singlet (dichlorophosphoryl substituent bound to the C≡C group) is observed at 6 ppm in the 31P NMR spectrum. This is consistent with the data of [32]. The 31P NMR spectrum also contains four doublets with pairwise coinciding coupling constants JPP observed in the 31P NMR spectrum of a mixture of the products of the reaction of trichloro(2-ethoxyethenyl)phosphonium hexachlorophosphorate with triethylamine. These data indicate that the reaction affords two compounds, each of which contains two magnetically nonequivalent phosphorus atoms in the P=O groups. The doublets at 28.3 and 21.4 ppm (2JPP = 56.5 Hz) belong to compound 21, while the other compound, 22, is characterized by the doublets at 25.6 and 21.9 ppm (2JPP = 48.8 Hz) (Table 3).
The spectral pattern of the obtained mixture of the reaction products changes over time. The concentration of 1-chloro-2-ethoxyethenylphosphonyl dichloride (19) decreases in the mixture, and the amount of two compounds (each containing two phosphorus atoms) noticeably increases. The signals at 6 ppm disappear from the spectrum. These spectral data show that compound 19 and phosphorus-containing acetylene 20 participate in the formation of the compounds containing two phosphorus atoms in the molecules. Since compound 19 is stable in the absence of triethylamine, we assume that triethylamine is also involved in the formation of diphosphorylated compounds. Based on these data and 1H and 13C NMR results, we propose the scheme for the synthesis of diphosphorylated heterocyclic compounds 21 and 22, i.e., 1,4-oxaphosphinines containing the dichlorophosphoryl substituent in position 2 (Scheme 6).
The formation of 1,4-oxaphosphinine 21 starts from the conversion of 1-chloro-2-ethoxyethenylphosphonyl dichloride (19) with triethylamine to phosphorus-containing acetylene, 2-ethoxyethynylphosphonyl dichloride (20). Then, the attack of the trichlorophosphonium group of trichloro(2-ethoxyethenyl)phosphonium hexachlorophosphorate on the triple bond of compound 20 leads to diphosphorylated ketene A. The dichlorophosphonium group of diphosphorylated ketene A is transformed into the —POCl— group (ketene B) due to the phosphorylated derivatives of ethyl vinyl ether that are present in the reaction mixture. Then, heterocyclization occurs, preceded by the intramolecular E,Z-isomerization of the ethoxyvinyl fragment of the diphosphorylated ketene molecule (B,C). The formed oxaphosphinine 21 is acetal; hence, a portion of this product is transformed into 2-chlorosubstituted oxaphosphinine 22 in the presence of phosphorus pentachloride. Heterocyclization permits the rationalization of the rather high coupling constants 3JP1H = 57.2 Hz for compound 21 and 62.0 Hz for compound 22. The high values of coupling constant 4JP2H = 40.4 Hz for compound 21 and 36.6 Hz for compound 22 can be explained by the W-shaped configuration of the fragment of the oxaphosphinine molecule binding the P2 phosphorus atom to the vinyl proton (Table 3).
The 31P NMR chemical shift values of the tetracoordinated phosphorus atom of POCl2 in compounds 19 and 20 differ several times: 80.1 and 6.0 ppm, respectively (Table 3). The presence of a triple bond (20) has a great shielding influence on the position of the resonance signal of phosphorus in the 31P NMR spectra.
The 13C NMR spectra (with proton decoupling) of a mixture of compounds 21 and 22 exhibit two groups of signals characterizing the C1 and C2 atoms directly bound to the phosphorus atoms. Two doublets of doublets at 106.4–108.7 ppm shifted relatively to each other by 8.5 Hz relate to the C1 atoms in compounds 21 and 22. The high values of coupling constants 1JP1C1 = 157.7 and 150.0 Hz and 1JP2C1 = 67.5 and 68.0 Hz indicate that two different phosphorus atoms are directly bonded to the C-1 carbon atom. The significant value of 1JP1C2 = 101.2 and 100.0 Hz in compounds 21 and 22 indicates that one phosphorus atom is bonded to the C-2 carbon atom. Thus, the 13C NMR spectral data are consistent with the structures of oxaphosphinines 21 and 22.
Trichloro(styryl)phosphonium hexachlorophosphorate was used as such a model compound (Scheme 7).
Its reaction with triethylamine is accompanied by a weak coloration of the reaction mixture. The 31P NMR spectrum exhibits a singlet at 219.6 ppm (PCl3). A specific feature of the spectrum is the presence of two intense signals in the range characteristic of organic chloride of trivalent phosphorus. It is most likely that styrylphosphonyl (23) and 1-chloro-2-phenylethynylphosphonyl (24) dichlorides are formed (Scheme 7). Compound 23 is characterized by a doublet of doublets at 162.8 ppm (2JPH = 9.2 Hz, 3JPH = 19.1 Hz), and a doublet at 157.8 ppm (3JPH = 22.1 Hz) belongs to compound 24. A reason for the formation of compound 24 is the chlorination at the multiple bond. It should be mentioned that compound 24 was obtained [30] by heating styryltetrachlorophosphorane under much harsher conditions. The replacement of the oxygen-containing ethoxy group in hexachlorophosphorate by the phenyl group containing no oxygen atoms results in the absence of styrylphosphonyl chloride in the reaction mixture. Thus, the 31P NMR chemical shifts for PCl2 in compounds 23 and 24 in the region of 160 ppm are characteristic of a tricoordinated (trivalent) phosphorus atom.
Thus, the NMR studies of chlorophosphorylated enamines on the basis of accessible tertiary amines and a wide range of other nitrogen-containing organic nucleophiles contributes essentially to the chemistry of unsaturated organophosphorus compounds. 31P NMR spectroscopy provides the most reliable and unambiguous method for the investigation of EZ-isomeric structures of phosphorylated enamines.

4. The Structural Aspects of the Phosphorylation Products of Enamides

Phosphorus-containing enamines and enamides attract the attention of researchers due to their synthetic potential and biological properties. Organophosphorus complexes are widely used in biological, pharmaceutical and material fields. Great demands on their diversity and availability encourage chemists to intensify research efforts to find effective and general routes to the synthesis of their derivatives and to study the structure of the reaction products [33,34,35,36,37]. Moreover, various enamides are suitable reactants and show high reactivity.
The low stability of alkyl- and aryl-substituted tertiary enamines in acidic media hinders the application of phosphorus pentachloride as a phosphorylating agent. So, N, N-diphenylethenamine was reacted with PCl5 in benzene with instant resinification, and in the reaction mixture after its treatment with sulfurous anhydride, only 31P NMR was used to identify 2-diphenylaminoethene-1chloro-1-phosphonic acid dichloride—E-(PN) isomer (Scheme 8) (see footnote in Table 1).
Enamines, containing strong electron-withdrawing substituents at the nitrogen atom and having a less nucleophilic character, behave in the phosphorylation reaction like vinyl ethers; therefore, it is possible to obtain organophosphorus compounds not contaminated with by-products in high yield.
The phosphorylation of N-vinyl-substituted tertiary amides, in which nitrogen nucleophilicity is significantly weakened, has been studied [14,38,39,40] (Scheme 9, Table 4).
Tertiary enamides can be attacked by phosphorus pentachloride at two reaction centers, namely at the double bond of the vinyl fragment and at the carbonyl group, and the reactivity of the carbonyl group should increase upon passing from N-aryl-substituted to N-alkyl-substituted enamides.
It was found [39,40] that the studied enamides react with phosphorus pentachloride in the same way as with alkenes, i.e., at the vinyl group. Hexachlorophosphates 2533 are isolated as finely crystalline precipitates that are easily hydrolyzed in air. The N-benzyl derivative of hexachlorophosphorate PhCH2N(COMe)CH=CH- PCl 3 +   PCl 6 (33a) is formed as a slowly crystallizing oil.
Phosphorylation products of the most reactive N-vinyl-N-alkylacetamides undergo such a rapid transformation with the participation of the carbonyl group that compounds of the 2533 type cannot be detected. The replacement of the alkyl substituent at the nitrogen atom by the more electron-withdrawing benzyl group somewhat slows down the attack of phosphorus on carbonyl oxygen, and compound 33a is stable at room temperature for 1 h. Aryl substituents at nitrogen, competing for its lone electron pair with a phosphorus-containing double bond and a carbonyl group, further reduce the nucleophilicity of the latter. This leads to a significant increase in the stability of compounds 2529. Phosphorylation products of N-vinyl lactams and N-vinylimides are stable for several days, with compound 33 being the most stable, apparently due to efficient conjugation in the condensed phthalimide system.
The 3JHH value (15–16 Hz) of vinyl protons for the studied compounds 2542 indicates E-(PN)-isomers; i.e., when going from hexachlorophosphates (2533) to the corresponding enaminophosphoric acid dichlorides (3442), the configuration of the molecule does not change. The 31P values of chemical shifts for compounds 3442 are typical for acid chlorides of alkenylphosphonic acids and vary in the region of 33–37 ppm.
The values of the constants 3JPH = 25–27 and 3JPH = 21–25 Hz are typical for E-isomers, for hexachlorophosphates (3442) and for acid dichlorides (3442), respectively. Compounds 3442 are more stable than complex compounds 2533. The most stable are 2-(N-phenyl-N-benzoylamido)ethene-1-phosphonic acid dichloride (35) and 2-(N-phthalimido)ethene-1-phosphonic acid dichloride (42). 2-(N-phthalimido)ethene-1-trichlorophosphonium hexachlorophosphate (33) is stable at room temperature and under weak heating; at 100°C, it turns into 2-(N-3,3-dichloro-1-hydroxyisoindolinyl)-ethenephosphonic acid dichloride (43) (Scheme 10) [14,40].
The intramolecular mechanism of the formation of the phosphoryl group is evidenced by the Z-configuration of dichloride 43, which inevitably arises during the intramolecular electrophilic attack of the dichlorophosphonium cation at one of the carbonyl oxygen atoms of compound 33. The trans-cis-isomerization of the enaminotrichlorophosphonium group, which is necessary for the formation of the transition state (A), occurs either under the catalytic effect of hydrogen chloride present in the reaction medium or due to high-temperature rotation around the multiple bond without opening it. At a temperature above 100 °C, the partial resinification of the reaction mixture occurs, and the amount of chlorination products and various organophosphorus compounds of unknown structure, resulting from thermal degradation, increases.
The 1H, 13C and 31P NMR chemical shifts and coupling constants of phosphonic dichloride 43 are presented in Table 5. The 31P NMR spectrum of compound 43 contains a single signal at 23.3 ppm in the form of a doublet of doublets. The Z-geometry of 43 follows from the values of the constants 3JHH = 10.5 and 3JPH = 57.1 Hz. A small value of the constant 2JPC = 8.5 Hz also corresponds to the Z-(PN)-configuration. The 35Cl NQR spectrum of acid dichloride 43 consists of four signals (ν77 26.030, 26.766, 38.072, 38.228 MHz). Two low-frequency signals refer to chlorine atoms in the POCl2 group since they are in the frequency range characteristic of compounds of the RPOCl2 series. The other two signals belong to the chlorine atoms of the CCl2 group (Table 5) [40].
The quantitative transition of the E-(PN)-isomer of hexachlorophosphate (33) to the Z-(PN)-isomer 43, as well as the stability of the latter even at 100 °C, is unusual for the chemistry of phosphorus-containing enamines [39,41].
2-(N-succinimido)ethenyltrichlorophosphonium hexachlorophosphorate (32), in contrast to hexachlorophosphorate 33, is already unstable at room temperature and gradually undergoes deep transformations, especially in the presence of phosphorus pentachloride, culminating in the formation of a bicyclic heterocycle: 2-oxo-2,6-dichloropyrrolo[1,2-e]1,5,2-oxazaphosphorin-7-trichlorophosphonium (44) (Scheme 11) [14,41].
The E,Z-isomerization appears to be catalyzed by the hydrogen chloride present in the reaction medium. Intermediates A and B cannot be fixed in the reaction medium due to their rapid conversion to the Z-isomer of 2-(N-5,5-dichloropyrrolidone)ethenephosphonic acid dichloride C. The fixation of the geometry of intermediate C with the maximum distance of the dichlorophosphoryl group from the dichloromethylene group is due to steric factors; i.e., general regularity of intramolecular transformations of phosphorylated enimides containing phthalimide and succinimide groups is observed. Intermediate C was not isolated individually, but the signal at 21.6 ppm (dd) with 3JPH = 57.1 and 2JPH = 26.2 Hz in the 31P NMR spectrum corresponds to this compound. The values of the chemical shift and the coupling constants are close to those of dichloride 43 formed upon heating 33 (Scheme 10). The intermediate C undergoes rapid heterocyclization to bicyclic heterocycle D, followed by the elimination of hydrogen chloride and the formation of heterocyclic dienamine E. For intermediates D and E, the 31P NMR spectra show close values of chemical shifts of phosphorus atoms and coupling constants: 13.6 ppm (dd) with 3JPH = 20.0, 2JPH = 42.0 Hz, and 13.0 ppm (dd) with 3JPH = 22.0, 2JPH = 40.0 Hz, respectively. The compound E is phosphorylated at the double bond of the Δ2–pyrroline ring with excess phosphorus pentachloride to form 2-oxo-2,6-dichloropyrrolo[1,2-e]1,5,2-oxazaphosphorine-7-trichlorophosphonium hexachlorophosphorate (44). The 31P NMR spectrum of hexachlorophosphorate (44) shows three signals: 83.5 ppm, dd ( PCl 3 + ); 11.1 ppm, ddd (POCl); and a singlet of −297.7 ppm ( PCl 6 ) (Table 5). In the 1H NMR spectrum, doublets of doublets at 6.44 and 7.78 ppm with proton–proton constant 3JHH = 10.0 Hz correspond to vinyl protons, which indicates the cis-orientation of vinyl protons.
The action of SO2 on compound 44 gave 2-oxo-2,6-dichloro-7-dichlorophosphorylpyrrolo[1,2-e]1,5,2-oxazaphosphorine (45). Heterocycles 44 and 45, despite the presence of a chloropyrrole fragment in the molecule, are stable in the absence of moisture. Two signals are observed in the 31P NMR spectrum of oxazaphosphorine 45. The signal at 18.6 ppm corresponds to the exocyclic phosphorus atom, and the signal at 10.4 ppm belongs to the P(O)Cl group. The cis-configuration of the N-vinylchlorophosphoryl group is confirmed by the constants 3JHH = 10.5 and 3JPH = 40.5 Hz. It follows from the 13C NMR spectrum data that each phosphorus atom is bonded directly to one carbon atom. The spin–spin interaction of C-9 and C-8 carbon nuclei with the nuclei of both phosphorus atoms is consistent with the bicyclic structure of compound 45 (Table 5).
It was established [14,42] that 1-(N-carbazolyl)-1-butene is easily phosphorylated by phosphorus pentachloride to form 1-ethyl-2-(N-carbazolyl)ethene-1-trichlorophosphonium hexachlorophosphorate (46) and 1-ethyldichloride-2-(N-carbazolyl)ethene-1-phosphonic acid (47) (Scheme 12). In the 31P NMR spectrum of hexachlorophosphorate 46, two signals are observed: 96.6 ppm, td ( PCl 3 + ), 3JPH = 33.0 Hz (CH2, t), 3JPH = 26.0 Hz (d) and 11.1 ppm singlet −297.7 ( PCl 6 ). The value 3JPH = 26.0 Hz clearly indicates the E-isomer (P,N) of the alkenyltrichlorophosphonium cation 46. The 31P NMR spectrum of compound 47 shows 39.1 ppm, td (POCl2), 3JPH = 27.0 Hz (CH2, t), 3JPH = 18.5 Hz (=CH, d).
N-Acetylcarbazole reacts with PCl5 (Scheme 12) with the formation of the corresponding complex salt, which, after the elimination of SO2, turns into 2-carbazolyl-2-chlorovinyl-1-phosphonic acid dichloride (49) [14,43]. The 31P NMR spectrum of compound 49 contains a single doublet at 22.4 ppm with 2JPH = 21.6 Hz.
The attack of PCl5 in N-phenyl-substituted acetamides occurs at the oxygen of the C=O group; the intermediate unstable compound cleaves POCl3, turning into α-chloro-substituted tertiary enamine, which is further phosphorylated by excess phosphorus pentachloride at the carbon–carbon double bond to afford unsaturated organophosphorus compounds 50 and 51 (Scheme 13) [43,44]. Hexachlorophosphorate 51 is easily converted by the action of Alk 4 N +   I into dichlorophosphine 51a, and upon treatment with SO2 into phosphonic acid dichloride 52.
The 31P NMR spectrum of hexachlorophosphorate 50 has an intense narrow signal of −298.1 ppm ( PCl 6 ) and a doublet at 71.4 ppm, with 2JPH = 32.8 Hz ( PCl 3 + ). The 31P NMR spectrum of complex salt 51 contained two signals: a doublet at 75.7 ppm with 2JPH = 36.5 Hz, related to the organyltrichlorophosphonium cation ( PCl 3 + ), and −298.0 ppm, a broadened singlet characteristic of the PCl 6 anion. The 31P NMR spectrum of compound 51a contains two doublets, 161.9 ppm, 2JPH = 2.5 Hz, and 152.9 ppm, 2JPH = 4.4 Hz, in a 6:1 ratio, related to the E- and Z-isomers. In the 1H NMR spectrum, the doublets at 5.58 ppm with 2JPH = 2.5 Hz and 6.14 ppm with 2JPH = 4.4 Hz correspond to the vinyl protons of these isomers. The proton spectrum of 52 contains a multiplet in the region of 7.2 ppm (Ph) and a doublet at 5.14 ppm, 2JPH = 19.5 Hz. The 31P NMR spectrum has a single signal at 24.6 ppm with 2JPH = 19.5 Hz.
The acetyl group of enamides of enamidotrichlorophosphonium hexachlorophosphorates, initially formed during phosphorylation of enamides, participates in further transformations of organophosphorus compounds, which involve phosphorus chloride substituents at the double bond. It is established that enamidotrichlorophosphonium hexachlorophosphorates containing an acetamide group in the molecule in the presence of PCl5 are rapidly converted upon heating and more slowly at room temperature into 2-N-alkyl(aryl)-N-2-(dichlorophosphoryl)-alkenylamino-2-chloroethenylphosphonium hexachlorophosphorates (5357) (Scheme 14) [14,40,41,42,45,46,47].
The formation of compounds 5357 is the result of an intramolecular attack of the carbonyl group by the organyl trichlorophosphonium oxygen cation, leading to the appearance of a vinyl chloride group, which then reacts with PCl5 similarly to the phosphorylation of enamines. Hexachlorophosphorates 5357 cannot be isolated individually without impurities, since the reaction mixture, along with them, always contains products of further transformations. However, after treatment of compounds 5357 with sulfur dioxide, stable diphosphonic acid chlorides are formed (5862) (Table 6).
The value of the constant 3JHH = 14–16 Hz, as well as 3JPH = 20–26 Hz, indicates E-isomerization of (P,N)-organophosphorus compounds 5362. For compounds 53, 54, 56, 58 and 60, the constants 2JPH and 3JPH in the 31P NMR spectra are almost equal in magnitude.
The 35Cl NQR spectrum of compound 59 consists of four signals: ν77 26.608, 26.734, 26.933 and 36.775 MHz. The high-frequency signal (36.775 MHz) relates to the chlorine atom bonded to the carbon atom, and the three low-frequency signals belong to the chlorine atoms bonded to the phosphorus group POCl2.

5. The Structure of Enamide-Derived Azaphosphorines

It has been established that heating or long-term room-temperature storage of bisphosphorylated divinylamines (5356, 56a) delivers cyclic hexachlorophosphorates of 1,4-azaphosphoniarines (6367). The formation of compounds 6367 is the result of an intramolecular attack by the ethenyltrichlorophosphonium cation at the π-bond of the ethenyldichlorophosphoryl group. The transformation of disphosphorylated dialkenyls into heterocycles (6367) is noteworthy in that the authors of [14,47] using the trichlorophosphonium cation managed to phosphorylate a carbon–carbon double bond having a strong electron-withdrawing substituent. Compound 57 is unable to form the 1,4-azaphosphorine ring, apparently due to steric hindrance created by three substituents at the double bond for intramolecular cation attack of 1,4-azaphosphorine derivatives (6872) (Scheme 15) (Table 7) [46,47].
The transformation of compounds 6367 under the action of SO2 into 1,4-azaphosphorines (6872) occurs much more slowly than similar transformations of acyclic alkenyltrichlorophosphonium hexachlorophosphonates.
The coupling constant 2JPH of the proton of the P–CH= group ( PCl 3 + ) is almost not observed in the 31P NMR spectra, and only for some compounds is it possible to fix the constant in the range of 0.6 Hz. The insignificant value of the geminal constant is characteristic of 1,4-heterophosphorine cycles. The unusually large coupling constant across four bonds (4JPH) is due to the W-shaped planar structure of the molecular fragment of these compounds, which includes both phosphorus atoms and a proton. The value of the coupling constant 2JPP (15–18 Hz) is typical for geminal bonded phosphorus atoms. The existence of the C-Cl bond in the molecule of compounds 6372 is confirmed by the presence of signals in the 35Cl NQR spectra in the frequency range of 38.7–39.4 MHz.
Similarly, the reaction of N-methyl-N,N-diacetamide with phosphorus pentachloride leads to 1,4-azaphosphorine containing an exocyclic phosphoryl group [41,48], while the phosphorylation of N,N-diacetamide proceeds differently [49] (Scheme 16).
The reaction, carried out under moderate vacuum conditions, furnishes 1H-2,4,4,6-tetrachloro-1,4-dihydro-1,4-azaphosphorinonium hexachlorophosphorate (75). The formation of compound 75 is probably preceded by imine–enamine rearrangement, which converts imidoyl chloride A to dialkenylamine B. Intramolecular phosphorylation of the latter leads to the formation of heterocycle 75, which is transformed by formic acid to 1H-4-oxo-2,4,6-trichloro-1,4-dihydro-1,4-azaphosphorine (76) (Table 8).
The phosphorus signal of PCl 6   in the 31P NMR spectra of compounds 73 and 75 appears in the region of −297.2 and −298.9 ppm, respectively. The large value of the direct coupling constant 1JCH=180.0–187.0 Hz may indicate the influence of the lone electron pair (LEP) of the oxygen atom on the C-H bond, and in addition, the proximity of the carbon atom to the electronegative phosphorus (or nitrogen) atom in the cycle [50,51,52,53]. NQR spectroscopy data also indicate the formation of heterocyclic compounds. It should be noted that 35Cl NQR spectroscopy is successfully used to prove the structure of chlorine-containing compounds [54].
So, diacetamide reacts with phosphorus pentachloride to give the cyclic phosphonium salt (75) which is converted to azaphosphorinone (76) upon treatment with formic acid. The 31P NMR chemical shift values of POCl in 1,4-oxaphosphinines 21 and 22, azaphosphorines 6872 and azaphosphorinone 76 appear in the region 24–28 ppm and practically do not depend on the nature (structure) of the heterocycle.
The phosphorylation of a representative of cyclic N-acylated hydrazones, 1-acetyl-3-methyl-6-oxo-1,4,5,6-tetrahydropyridazine, was carried out (Scheme 17) [55].
The process proceeds similarly to the phosphorylation of N,N-disubstituted acetamides [43]. The carbonyl group of the heterocycle lowers the nucleophilicity of the N-acetyl nitrogen atom, directing the reaction to the formation of α-chlorenamine A. The intermediate A is further phosphorylated by phosphorus pentachloride to afford complex salt 77, from which dichloride 2-(3-methyl-6- oxo-1,4,5,6-tetrahydropyridazin-1-yl)-2-chloroethenylphosphonic acid (78) is easily obtained. The 31P NMR spectrum of dichloride 78 contains a doublet at 33.1 ppm with 2JPH = 24.8 Hz.
Enamino ketones as well as ethyl ethers of phenylaminocrotonic acids are convenient starting compounds for the synthesis of nitrogen–phosphorus-containing heterocycles via phosphorylation [56,57,58]. 4-Arylamino-3-penten-2-ones react with PCl5 to give 4-(N-aryl-N′-dichlorophosphorylamino)-2-chloro-1,3-dienyltrichlorophosphonium hexachlorophosphorates 7981. Compounds 7981 and diphosphonic acid chloride 82 readily formed from 79 undergo gradual heterocyclization to 1,2-dihydro-1,2-azaphosphorines 8587 and compound 88 with the elimination of phosphorus oxychloride (Scheme 18) (Table 9) [56,57].
The 1H, 13C and 31P NMR chemical shifts and coupling constants of phosphonic chloride 88 (CDCl3) are presented in Table 9. The 31P NMR chemical shifts of 85, 86, and 87 have the following values (ppm): 56.2 dd, 3JPH = 34.4 and 2JPH = 16.5 Hz; 57.3 d, 2JPH = 17.8 Hz; and 60.4 d, 2JPH = 20.1 Hz, respectively.
It has been found that a longer (for several hours) treatment of ethyl phenylaminocrotonoate with phosphorus pentachloride leads to 1,4-quinolone 89 containing a trichlorophosphonium group in position 3 (Scheme 19) (Table 9) [58]. The heterocyclization is preceded by the formation of phosphorus-containing phenylaminocrotonic acid chloride. Its carbonyl group exhibits increased electrophilicity.
We found that 2-methyl-1,4-quinolone is not phosphorylated by phosphorus pentachloride, which confirms the proposed scheme for the reaction of crotonate with PCl5, according to which the phosphorylation at the double bond is preceded by acylation of the benzene ring. The phosphorylation at the NH nitrogen atom of the quinolone is also possible, but the NH bond in the resulting phosphonic acid is regenerated in the course of hydrolysis.
The nitrogen chemical shift in the 15N NMR spectrum of compound 89 is at −220.1 ppm, which is characteristic of a pyrrole-type nitrogen atom [52,53].
It should be noted that unsaturated ketones behave similarly to enamino ketones when phosphorylated with phosphorus pentachloride [59,60]. C-phosphorylation products dienyltrichlorophosphonium hexachlorophosphorates have been studied by means of 2D and multinuclear 1H, 13C and 31P NMR spectroscopy. The reaction most probably starts with the attack of phosphorus pentachloride at the carbonyl group to form O-phosphorylated product A (Scheme 20) [59].
In the presence of excess phosphorus pentachloride, tetrachlorophosphorane A undergoes ionization to give intermediate B, which decomposes to afford diene C. The latter is phosphorylated by excess phosphorus pentachloride similarly to the known alkenes with the formation of hexachlorophosphorates 9094. The 31P NMR spectra of compounds 9094 have signals at 84–89 ppm belonging to the PCl 3 + cation; a broadened signal at −297 to −298 ppm corresponds to the PCl 6 anion (Table 10).
The 13C NMR spectra of hexachlorophosphorates 9094 contain doublet signals in the region of 116–119 ppm, which are characteristic of the carbon atom directly bonded to the phosphorus atom (Table 10). The coupling constant 1JPC lies in the range of 115–117 Hz, but the spin–spin coupling constant 2JPH between the phosphorus atom and vinyl protons turns out to be quite sensitive to the effect of substituents (41–56 Hz).
Chlorophosphorylated 1,3-dienes are also formed by the action of PCl5 on diacetone alcohol [60]. It has been established that two pentadienyltrichlorophosphonium hexachlorophosphates (95, 96) are produced in approximately equal amounts via phosphorylation. Under the action of dimethylacetamide, hexachlorophosphorates (95, 96) are converted into a mixture of two isomers: 2-chloro-2,4-pentadiene-4-methyl-5-phosphonic (97) and 2-methyl-2,4-pentadiene-4-chloro- 5-phosphonic (98) acids (Scheme 21).
The process starts from the attack by the hydroxy oxygen atom on the phosphorus atom of phosphorus pentachloride, which is evidenced by a vigorous evolution of hydrogen chloride. The formed intermediate A readily undergoes cyclization to trichlorophosphorane B as a result of the intramolecular attack by the carbonyl oxygen atom of intermediate A on the phosphorus atom of tetrachlorophosphorane group. The attack by the carbonyl group of compound A at phosphorus pentachloride present in the reaction mixture in excess is unlikely because the intramolecular process is preferable. The intermediate B readily dissociates to ions, and the formed quasiphosphonium salt C, according to the second step of the Arbuzov reaction, is transformed to compound D. The subsequent conversion of D to dienes E and F proceeds in the presence of phosphorus pentachloride with the elimination of trichlorophosphorus oxide and hydrogen chloride, similarly to the process of the formation of alkenes in the phosphorylation of tertiary alcohols containing at least two methyl groups next to the hydroxy group with phosphorus pentachloride. The process is finalized by the phosphorylation of pentadienes with phosphorus pentachloride (Scheme 21).
The structure of the obtained products has been established by two-dimensional and multinuclear 1H, 13C and 31P NMR spectroscopy. In the 31P NMR spectra of dienyltrichlorophosphonium hexachlorophosphorates (95, 96) recorded both in nitrobenzene and in nitromethane, two doublets of approximately the same intensity and a broadened singlet in the region of −296 ppm, characteristic of the PCl 6 anion, are observed in the region 90.5 ppm (d, 2JPH 52.8 Hz) and 85.7 ppm (d, 2JPH 56.0 Hz). The 13C NMR spectrum of a mixture of compounds 95 and 96 shows two doublets at 116.5 ppm (C-1, 1JPC 114.9 Hz) and at 114.8 ppm (C-1, 1JPC 111.2 Hz) characteristic of carbon atoms directly bonded to the phosphorus atom. In the 31P NMR spectrum of a mixture of isomers 97 and 98, signals detected in the region 26.2 ppm (d, POCl2, 2JPH1 35.3 Hz) and 25.6 ppm (d, POCl2, 2JPH1 36.1 Hz) relate to compounds 97 and 98, respectively. The 1H NMR spectrum of compound 97 contains signals at 2.36 ppm (d, C(2)Me, 4JPH 3.5 Hz) and 2.29 ppm (s, C(4)Me). The 1H NMR spectroscopy data of diene 97 indicate the cis-orientation of the methyl group towards the POCl2 moiety in the propenyl fragment. This is evidenced by the value of the constant 4JPH = 3.5 Hz.
It was established [61] that acetaldoxime reacts with PCl5 to form the stable phosphorus-containing enamine 99. The trichlorophosphase group formed in compound 99 reduces the nucleophilicity of the nitrogen atom, which ensures the stability of this compound (Scheme 22).
2-(Trichlorophosphazo)-2-chloroethenyltrichlorophosphonium hexachlorophosphorate (99) is isolated as a white powder. The 31P NMR spectrum of compound 99 shows δ 80.5 ppm, d ( PCl 3 + ), 2JPH = 44.2 Hz; δ 15.9 ppm, d (P=N), 4JPH = 15.4 Hz; δ –297.8 ppm, br s ( PCl 6 ). In the 1H NMR spectrum in the region of vinyl protons, there is a single signal at 5.67 ppm, dd, which indicates that both PCl3 groups belong to the same molecule.
Oxime derivatives containing a vinyl group at the oxygen atom react with PCl5 in the same way as vinyl esters [62]. O-vinyloximes are phosphorylated at the O-vinyl group at room temperature, forming alkenyltrichlorophosphonium hexachlorophosphorates (Scheme 23).
Under the action of sulfur dioxide, hexachlorophosphorates are converted into E-ethenylphosphonic acid dichlorides (100, 101) (3JHH = 13.0 Hz). The structure of the compounds has been established by 2D and multinuclear 1H, 13C and 31P NMR spectroscopy [62].

6. The Structural Features of Acetylurea-Derived Diazaphosphorines

Organophosphorus compounds obtained by phosphorylation of urea derivatives and related compounds are widely used in the life sciences, medicine chemistry, pharmaceuticals, organic catalysis and materials science [63,64,65,66,67,68,69,70,71,72]. Efficient P-C or P-N bond formation using the various available phosphorus sources plays a critical role in the construction of organophosphorus compounds. Among them, PCl5 acts as an essential source of phosphorus; therefore, its use to form P-C or P-N covalent bonds is being actively developed.
The problem of obtaining phosphorylated nitrogen-containing heterocyclic compounds, which play an important role in medicine, agrochemistry and modern metal complex catalysis, deserves special attention. The need to select conditions for each particular type of heterocycles, harsh reaction conditions (high temperature, the presence of strong bases and acids) and the high cost of the catalysts used significantly limit the application of this approach in practice. An alternative devoid of these disadvantages can be the construction of a heterocyclic ring using universal phosphorus pentachloride. With a rational approach, the “multiple” reactivity of such a reagent will provide an arsenal of substrate-controlled transformations leading to structurally diverse products.
Due to the wide practical application, the chemistry of organophosphorus compounds is developing rapidly. Among numerous series, phosphorus-containing heterocycles have been synthesized, which can be represented as structural analogues of azabenzenes: pyridine, pyrimidine, s-triazine. The interaction of urea derivatives with phosphorus pentachloride and the phosphorylation products have been successfully studied [73,74,75,76,77,78,79,80,81,82,83,84]. It is known that the results of phosphorylation of urea derivatives depend on the conditions of the process.
The direction of phosphorylation of N-acetylurea under the action of phosphorus pentachloride significantly depends on the ratio of reagents (Scheme 24) [73,76,77,78,79]. Further transformations of intermediate A are determined by the presence or absence of PCl5 in the reaction medium. At a molar ratio of urea and PCl5 of 1:5, hexachlorophosphorate 102 is obtained. However, in the absence of an excess of PCl5 in the reaction medium, intramolecular phosphorylation of the C=C double bond of azabutadiene A by the trichlorophosphazol group occurs to deliver 2,2,4,6-tetrachloro-2,2-dihydro-1,5,2-diazaphosphorine 103 (Scheme 24) (Table 11). The action of formic acid on complex compound 102 gives N1-dichlorophosphoryl-N2-(1-chloro-2-dichlorophosphorylethenyl)-C-chloroformamidine 102a [73,79].
The 31P NMR spectrum of compound 102 is represented by three signals: δ, ppm, 82.8 d ( PCl 3 + ), 2JPH = 43.2 Hz; 23.7 s ( PCl 3 ); −297.2 s ( PCl 6 ). A single signal at 6.55 ppm (=CH) is observed in the proton spectrum. The 13C NMR spectrum of complex compound 102 shows the following: δ, ppm, 156.00 dd (ClC=CH), 2JPC = 10.2 and 4JPC = 2.0 Hz; 144.12 dd (ClC=N), 2JPC = 6.8 and 4JPC = 2.0 Hz; 95.62 (=CH). The presence of a spin–spin interaction of each of the carbon atoms simultaneously with two phosphorus atoms proves the presence of two phosphorus atoms in one molecule.
The 31P NMR spectrum of chloroformamidine 102a has two signals, δ, ppm: 25.3 d (C-POCl2), 2JPH = 19.8 Hz, and 3.3 s (N-POCl2). In the 1H spectrum, the signal =CH of the proton is in the region of 6.2 ppm. Two high-frequency signals in the 35Cl NQR spectrum of compound 102a at 36.210 and 37.243 MHz refer to the chlorine atoms of the C–Cl bond, the latter being related to the chlorine atom in the C(Cl)=CH group. Two low-frequency signals at 27.638 and 27.088 MHz refer to chlorine atoms in the POCl2 group.
Heating of the reaction mixture obtained by the interaction of PCI5 and chloroacetylurea is accompanied by the release of hydrogen chloride and leads to the formation of 2,2,3,4,6-pentachloro-2,2-dihydro-1,5,2-diazaphosphorine (104) (Scheme 25) [76].
The 31P NMR spectrum of diazaphosphorine (104) contains a singlet at 44.7 ppm, located in the region characteristic of azaphosphorines containing a PCl2 group in the heterocycle (Table 11). The structure of heterocycle 104 is also confirmed by the data of 35Cl NQR spectroscopy [76].
Upon transition of diazaphosphorine (DAP) (103) to the crystalline state, the conjugation of the P=N, C=C and C=N bonds is disturbed or significantly weakened, as indicated by a significant change in the characteristic vibrations in the IR spectra of these structures [78]. The violation of conjugation occurs, most likely, as a result of the dimerization of diazaphosphorine (103) at the P=N bond, and the resulting dimer (103a) is stable only in the crystalline state, and upon melting or dissolution, it dissociates extremely easily to give the monomeric form (103) (Scheme 26).
In the 31P NMR spectra of solutions of compound 103 at room and low (−80 °C) temperatures, no signals are observed in the region characteristic of the pentacoordinated phosphorus atom, which confirms the conclusion that only the monomeric form is present in the solution. The dimeric structure of DAP (103a) in the crystalline state is confirmed by NQR and photoelectron spectroscopy data, as well as by quantum chemical calculations [78,85]. A weighty argument in favor of the dimeric structure (103a) is the significant splitting of the resonant doublet signal (ν77 27.788 and 29.308 Hz), which characterizes the PCl2 fragment in the NQR 35Cl spectrum of compound 103a. This indicates that the electronic distribution (state) of chlorine atoms differs significantly. Upon dimerization of 103 to 103a, the phosphorus atom becomes pentacoordinated, thereby acquiring a trigonal bipyramidal configuration. As a result, the nitrogen atoms of the four-membered cycle can occupy axial (one) and equatorial (second) positions. Accordingly, one chlorine atom becomes axial, and the other becomes equatorial, which causes the observed splitting of the ν77 (P-Cl) signal. In the NQR 35Cl spectra of chlorophosphoranes, the signals of chlorine nuclei occupying the equatorial position in the trigonal bipyramid are observed at higher frequencies than the signals of axial chlorine atoms.
The interaction of PCl5 and N-methyl-N′-acetylurea unexpectedly affords 2-(trichlorophosphazo)-2-chloroethenyltrichlorophosphonium hexachlorophosphorate (99) (see Scheme 22), and a minor amount of 1-methyl-2,2,4,6-tetrachloro-1,2-dihydro-1,5,2-diazaphosphorinonium hexachlorophosphorate (105) is formed (Scheme 27). Under the action of SO2, compound 105 is readily converted to 1-methyl-2,4,6-trichloro-1,2-dihydro-1,5,2-diazaphosphorine (106) (Table 11). Both heterocycles are easily hydrolyzable crystals, stable without moisture access [14,75].
The structures of compound 99 and heterocycles 105 and 106 were proved by 1H, 13C and 31P NMR methods, as well as 35Cl NQR spectroscopy data. The 31P NMR spectrum of compound 105 has a doublet at 64.4 ppm corresponding to the phosphorus atom of the PCl2 group and a broadened singlet in the region of −297.5 ppm (Table 11).
Under the action of PCl5 on N-methyl-N′-chloroacetylurea, the chloroacetyl group is not phosphorylated; the reaction furnishes a compound with a four-membered diazaphosphetidine ring—1-methyl-2,2,2-trichloro-3-(1,1,2,2-tetrachloroethyl)-1,3-diaza-2-phosphetidinone (107) (Scheme 28) [84].
The structure of compound 107 has been established by NMR spectroscopy data. In the 31P NMR spectrum of phosphetidinone 107, a single signal refers to the pentacoordinated phosphorus atom: −69.1 ppm, qd, 3JPH = 24.4 and 4JPH = 1.0 Hz (Table 11).
The phosphorylation of N-acyl-N,N′-dimethylurea derivatives with phosphorus pentachloride gives rise to six-membered heterocycles, 1,5-dimethyl-2,2,4-trichloro-6-oxo-3-alkyl-1,5,2-diazaphosphoniarin hexachlorophosphorates (108110) (Scheme 29) [74].
The phosphorylation begins with the attack of phosphorus pentachloride at the oxygen atom of the acetyl group to produce α-chloramine. The enurea is further phosphorylated at the double bond with the formation of acyclic hexachlorophosphorate 2-(N,N′-dimethylurea)-2-chloro-1-alkylethenyltrichlorophosphonium A. The subsequent rapid intramolecular attack of the chlorophosphonium cation at the secondary nitrogen atom of urea ends with cyclization to deliver diazaphosphorines (108110). The action of SO2 on compounds 108110 results in 1,5-dimethyl-2,4-dichloro-2,6-dioxo-3-alkyl-1,2,5,6-tetrahydro-1,5,2-diazaphosphoniarins (111113) (Table 12).
The signals of the tetracoordinated phosphorus atom of the PCl 2 + group in compounds 6367 (Table 8), 73 and 75 (Table 8), 8587 and 105 (Table 11), and 108110 (Table 12) in the 31P NMR spectra are observed in a higher frequency field (45–64 ppm) than those of the POCl moiety in products 111113 (17–19 ppm) (Table 12), which is apparently more associated with the existence of a molecule in a charged form.
Reduction of complex compounds 108110 with tetraethylammonium iodide gives 1,5-dimethyl-2,4-dichloro-6-oxo-3-alkyl-1,2,5,6-tetrahydro-1,5,2-diazaphosphoniarines (114, 115) (Scheme 29) (Table 12) [74]. In the 31P NMR spectrum of compound 115, a multiplet signal is observed in the region of 99.3 ppm, which corresponds to the phosphorus atom of the PCl group. It should be noted that the 31P NMR chemical shift of the tricoordinated (trivalent) phosphorus atom of the PCl group (114, 115) is quite large, as expected, and lies in the region of approximately 100 ppm (Table 12).
N-acetyl-N,N′-ethyleneurea (1-acetylimidazolidin-2-one) reacts with PCl5 to form 2-(2-oxo-1-imidazolidinyl)-2-chloroethenyltrichlorophosphonium hexachlorophosphorate (116) (Scheme 30) [81].
Phosphorus pentachloride attacks the oxygen atom of the acetyl group of acetylethyleneurea, but steric and conformational hindrances in cyclic ureide do not allow intramolecular attack of the nitrogen atom N-H by the organyltrichlorophosphonium cation (116) with the formation of a heterocycle, as observed in the phosphorylation of acyclic trisubstituted ureides [74]. Compound 116, upon treatment with SO2, is converted into 2-(2-oxo-1-imidazolidinyl)-2-chloroethenylphosphonic dichloride (117); its hydrolysis yields 2-oxo-2-(2-oxo-1-imidazolidinyl) ethylphosphonic acid (118) (Table 13).
The 31P NMR chemical shifts of 116 have the following values (ppm): 80.2 d, 2JPH = 38.9 Hz ( PCl 3 + ) and −295.5 br s ( PCl 6 ).
A generalization of the analysis of 31P NMR chemical shifts of nitrogen–phosphorus-containing enamines and the corresponding heterocycles shows that a change in the coordination number of the phosphorus atom from 3 to 6 significantly changes δ31P from about + 200 to −300 ppm. As regards the spin–spin coupling constants, the Karplus equation is applicable to the saturated systems of the H–C(sp3)–C(sp3)–H type. The vicinal interaction in such systems strongly (more precisely, mainly) depends on the value of the dihedral angle φ. If one of the carbon atoms is in the sp2 hybridization state, the Karplus equation becomes inapplicable. In our case, we deal with the unsaturated compounds (enamines), in which carbon atoms are in the state of sp2 hybridization.

7. Conclusions

The reactions of phosphorus pentachloride with enamines, tertiary amines, enamides, diacetamides, N-vinyl-substituted cyclic imides and N-acetylureas can be employed as a platform for the simple and convenient synthesis of hard-to-reach nitrogen-containing organophosphorus compounds and new types of heterocycles based on them. The development of efficient and reliable methods for the formation of carbon–phosphorus bonds is of great importance in connection with the widespread use of organophosphorus compounds in chemistry, materials science and biology. In recent years, interest in functionalized organophosphorus compounds has been steadily increasing, as evidenced by a large number of publications in the field of their chemical and structural studies [86,87,88,89,90,91,92]. It should be noted that the skillful use of phosphorus pentachloride as a phosphorylating agent for derivatives of enamines, enamides, tertiary amines, acetyl ureas and other nucleophiles leads to the production of various C- and N-chlorophosphorylated compounds, which, under certain conditions, are transformed to form promising nitrogen- and phosphorus-containing heterocyclic systems of diverse structures.
NMR spectroscopy studies of chlorophosphorylated enamines based on available tertiary amines and a wide range of other nitrogen-containing organic nucleophiles make a significant contribution to the development of the chemistry of unsaturated organophosphorus compounds. The use (involvement) of multipulse and multinuclear NMR spectroscopy, in particular 31P NMR, simplifies the tasks associated with establishing the structure of phosphorylation products and stereochemical behavior (e.g., for E,Z-isomeric compounds). Analysis of the 31P NMR chemical shifts of phosphorus–nitrogen-containing enamines and the corresponding heterocycles indicates that a variation in the coordination number of the phosphorus atom from 3 to 6 leads to a dramatic change in the screening of the 31P nucleus from about +200 to −300 ppm.
31P NMR spectroscopy is the most convenient, powerful and simple express method for recognizing the coordination number of the phosphorus atom in organophosphorus compounds, as well as for determining (identifying) the structure of isomeric forms of organophosphorus products. A certain contribution to the determination of the structure of chlorophosphorylated compounds is made by 35Cl NQR spectroscopy.
This review article can be considered as the first step towards a further deeper analysis of NMR spectroscopy and quantum-chemical data for a wide range of organic compounds, in particular, saturated and unsaturated compounds and heterocycles based on them. In this regard, it will be possible to apply the Karplus equation, which works mainly for saturated systems of the H–C(sp3)–C(sp3)–H type.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

Dedicated to the anniversary of the famous specialist in the field of chemistry of organophosphorus compounds, Vladimir Rozinov. I am sincerely grateful to V. Rozinov, an outstanding chemist, for consultations and helpful comments in writing this review.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Karl, D.M. Phosphorus, the staff of life. Nature 2000, 406, 31–33. [Google Scholar] [CrossRef] [PubMed]
  2. Demkowicz, S.; Rachon, J.; Dasko, M.; Kozak, W. Selected organophosphorus compounds with biological activity. Applications in medicine. RSC Adv. 2016, 6, 7101–7112. [Google Scholar] [CrossRef]
  3. Kochian, L.V. Rooting for more phosphorus. Nature 2012, 488, 466–467. [Google Scholar] [CrossRef] [PubMed]
  4. Montchamp, J.-L. Phosphinate Chemistry in the 21st Century: A Viable Alternative to the Use of Phosphorus Trichloride in Organophosphorus Synthesis. Acc. Chem. Res. 2014, 47, 77–87. [Google Scholar] [CrossRef]
  5. Budnikova, Y.H. Opportunities and challenges for combining electro- and organometallic catalysis in C(sp2)-H phosphonation. Pure Appl. Chem. 2018, 91, 16. [Google Scholar] [CrossRef]
  6. Müller, C. Copper(I) complexes of low-coordinate phosphorus(III) compounds. In Copper(I) Chemistry of Phosphines, Functionalized Phosphines and Phosphorus; Elsevier: Amsterdam, The Netherlands, 2019; pp. 1–19. [Google Scholar] [CrossRef]
  7. Clausing, S.T.; Morales, D.; Orthaber, S.A. Preparation, photo- and electrochemical studies of a homoleptic imine-phosphaalkene Cu(I) complex. Inorg. Chim. Acta 2020, 513, 119958. [Google Scholar] [CrossRef]
  8. Zagidullin, A.A.; Sakhapov, I.F.; Miluykov, V.A.; Yakhvarov, D.G. Nickel Complexes in C-P Bond Formation. Molecules 2021, 26, 5283. [Google Scholar] [CrossRef]
  9. Larina, L.I. Organophosphorus Azoles Incorporating a Tetra-, penta-, and hexacoordinated Phosphorus atom: NMR spectroscopy and Quantum Chemistry. Molecules 2023, 28, 669. [Google Scholar] [CrossRef]
  10. Chen, W.-Q.; Ma, J.-A. Stabilized Nucleophiles with Electron Deficient Alkenes, Alkynes, Allenes. In Comprehensive Organic Synthesis, 2nd ed.; Knochel, P., Molander, G.A., Eds.; Elsevier: Amsterdam, The Netherlands, 2014; Volume 4, pp. 1–85. [Google Scholar] [CrossRef]
  11. Zarate, C.; Van Gemmeren, M.; Somerville, R.; Martin, R. Phenol Derivatives: Modern Electrophiles in Cross-Coupling Reactions. Adv. Organometall. Chem. 2016, 66, 143–222. [Google Scholar] [CrossRef]
  12. Gusarova, N.K.; Trofimov, B.A. Organophosphorus chemistry based on elemental phosphorus: Advances and horizons. Russ. Chem. Rev. 2020, 89, 225–249. [Google Scholar] [CrossRef]
  13. Gafurov, Z.N.; Kagilev, A.A.; Kantyukov, A.O.; Sinyashin, O.G.; Yakhvarov, D.G. The role of organonickel reagents in organophosphorus chemistry. Coord. Chem. Rev. 2021, 438, 213889. [Google Scholar] [CrossRef]
  14. Rozinov, V.G. Chlorophosphorylated Enamines. Synthesis and Heterocyclization. Ph.D. Thesis, Irkutsk State University, Irkutsk, Russa, 1989; 420p. (In Russian). [Google Scholar]
  15. Mitrasov, Y.N.; Smolina, I.N.; Yegorova, G.Y. Reaction of allylalkanoates with phosphorus pentachloride. Russ. J. Gen. Chem. 2008, 78, 2154–2155. [Google Scholar] [CrossRef]
  16. Klapotke, T.M.; Krumm, B.; Moll, R. Synthesis and Structure of 2,2,2- Nitrilotriacetyl Chloride with a Flat NC3 Pyramide. Z. Naturforsch. 2013, 68, 735–738. [Google Scholar] [CrossRef]
  17. Mitrasov, Y.N.; Sosnov, D.A.; Kondrateva, O.V. Reaction of spiro[2.2]pentane with phosphorus pentachloride. Russ. J. Gen. Chem. 2013, 83, 1167. [Google Scholar] [CrossRef] [Green Version]
  18. Romero, A.H. Modified procedure for the synthesis of 2-chloroquinoline-3-carbaldehydes using phosphorus pentachloride. Synth. Commun. 2016, 46, 287–291. [Google Scholar] [CrossRef]
  19. Da Paixao Soares, F.; Groaz, E.; Herdewijn, P. Phosphorus Pentachloride Promoted gem-Dichlorination of 2′- and 3′-Deoxynucleosides. Molecules 2018, 23, 1457. [Google Scholar] [CrossRef] [Green Version]
  20. Salkeeva, L.K.; Muratbekova, A.A.; Minayeva, E.V.; Voitichek, P. Phosphorylation of glycoluryl derivatives with phosphorus pentachloride. Bull. Karag. Univer. Chem. Ser. 2021, 101, 12–18. [Google Scholar] [CrossRef]
  21. Bohme, H.; Viene, H.L. (Eds.) Iminium Salts in Organic Chemistry; Part 1, Methods and Results; Wiley: New York, NY, USA, 1976; p. 619. [Google Scholar]
  22. Kukhar, V.P.; Shevchenko, M.V. Amidation of trivalent phosphorus acid chlorides with bis(dialkylamino)methanes. Russ. J. Gen. Chem. 1982, 52, 562–566. [Google Scholar]
  23. Shevchenko, M.V.; Kukhar, V.P. Amidation of Halogenated Phosphorus Derivatives with Formic Acid Orthomorpholide. Russ. J. Gen. Chem. 1984, 54, 1499–1503. [Google Scholar]
  24. Larina, L.I.; Rozinov, V.G.; Komarova, T.N. Studies of phosphorylation reaction of tertiary amines by NMR spectroscopy. Russ. Bull. Chem. 2013, 62, 39–46. [Google Scholar] [CrossRef]
  25. Pensionerova, G.A.; Rozinov, V.G.; Glukhikh, V.I. The phosphorylation of tertiary amines with phosphorus pentachloride. Russ. J. Gen. Chem. 1978, 48, 2796. [Google Scholar]
  26. Rozinov, V.G.; Pensionerova, G.A.; Donskikh, V.I.; Kalabina, A.V.; Domnina, E.S.; Skvortsova, G.G. Unsaturated organophosphorus compounds based on 1-vinylbenzotriazole. Russ. J. Gen. Chem. 1983, 53, 697–698. (In Russian) [Google Scholar]
  27. Rozinov, V.G.; Pensionerova, G.A.; Donskih, V.I.; Sergienko, L.M.; Korostova, S.E.; Mikhaleva, A.I.; Dolgushin, G.V. Reaction of alkyl- and phenylsubstituted N-vinylpyrroles with phosphorus pentachlorides. Russ. J. Gen. Chem. 1986, 56, 790–804. [Google Scholar]
  28. Larina, L.I.; Rozinov, V.G.; Dmitrichenko, M.Y.; Eskova, L.A. NMR investigation of chlorophosphorylation products of N-vinylazoles. Magn. Reson. Chem. 2009, 47, 149–157. [Google Scholar] [CrossRef]
  29. Rozinov, V.G.; Kolbina, V.E.; Donskikh, V.I. Reaction of divinyl ether with phosphorus pentachloride. Russ. J. Gen. Chem. 1987, 57, 847–849. [Google Scholar]
  30. Fridland, S.V.; Malkov, Y.K. Reactions and Investigation Methods of Organic Compounds; Knunyants, I.L., Melnikova, N.N., Grapova, A.F., Simonova, V.D., Eds.; Chemistry: Moscow, Russia, 1986; pp. 106–149. [Google Scholar]
  31. Larina, L.I.; Komarova, T.N.; Rozinov, V.G. NMR study of reactions of organyltrichlorophosphonium hexachlorophosphorates with triethylamine. Russ. Chem. Bull. 2015, 64, 732–737. [Google Scholar] [CrossRef]
  32. Moskva, V.V.; Novruzov, S.A.; Ismailov, V.M.; Razumov, A.I.; Zykova, T.N.; Akhmedov, S.T. Derivatives of substituted vinylphosphonic acids. 21. Transformations of substituted vinylphosphonates under the action of nucleophilic reagents. Russ. J. Gen. Chem. 1976, 46, 534–542. [Google Scholar]
  33. Sultana, T.; Khan, M.W. A facile synthesis of 5-Iodo-6-substituted Pyrimidines from Uracil-6-Carboxylic acid (Orotic acid). Dhaka Univ. J. Pharm. Sci. 2014, 12, 97–102. [Google Scholar] [CrossRef]
  34. Khan, W.; Haque, E.; Ahmed, F. An expeditious synthesis of 6-Amido-(1H,3H)-Pyrimidine- 2,4-Diones from Uracil-6-Carboxylic Acid. Dhaka Univ. J. Pharm. Sci. 2015, 13, 23–29. [Google Scholar] [CrossRef]
  35. Liu, Y.T.; Song, S.M.; Yin, D.W. Solvent-free green synthesis of 4-amino-5-substituted-1,2,4-triazole-3-thione by Grinding. Adv. Mater. Res. 2015, 1088, 367–370. [Google Scholar] [CrossRef]
  36. Zhang, D.-L.; Li, C.-K.; Zeng, R.-S.; Shoberu, A.; Zou, J.-P. Manganese(III)-mediated selective phosphorylation of enamides: Direct synthesis of β-phosphoryl enamides. Org. Chem. Front. 2019, 6, 236–240. [Google Scholar] [CrossRef]
  37. Lei, T.; Liang, G.; Cheng, Y.Y.; Chen, B.; Tung, C.H.; Wu, L.Z. Cobaloxime Catalysis for Enamine Phosphorylation with Hydrogen Evolution. Org. Lett. 2020, 22, 5385–5389. [Google Scholar] [CrossRef] [PubMed]
  38. Rozinov, V.G.; Izhboldina, L.P.; Malysheva, S.F.; Donskikh, V.I.; Kalabina, A.V. Phosphorylation of N-vinylacetanilide with phosphorus pentachloride. Russ. J. Gen. Chem. 1983, 53, 934–935. (In Russian) [Google Scholar]
  39. Rozinov, V.G.; Izhboldina, L.P.; Pensionerova, G.A.; Donskih, V.I.; Malysheva, S.F.; Glukhikh, V.I.; Sergienko, L.M.; Ratovskiy, G.V. Phosphor-containing enamines. II. Phosphorylation of N-vinyl-substituted tertiary amides, lactams and cyclic imides with phosphorus pentachloride. Russ. J. Gen. Chem. 1984, 54, 1051–1060. (In Russian) [Google Scholar]
  40. Rozinov, V.G.; Izhboldina, L.P.; Donskih, V.I.; Dolgushin, G.V.; Ratovskiy, G.V. Thermal transformations of phosphorylated N-vinylphthalimide. Russ. J. Gen. Chem. 1988, 58, 942–943. [Google Scholar] [CrossRef]
  41. Rozinov, V.G.; Dmitrichenko, M.Y.; Kolbina, V.E.; Dolgushin, G.V. Phosphorus-containing enamines from acyclic and cyclic amides. II. Phosphorylation of acetamides, their homologs, and lactams. Russ. J. Gen. Chem. 1999, 69, 1377–1385. [Google Scholar]
  42. Rozinov, V.G.; Pensionerova, G.A.; Izhboldina, L.P.; Donskih, V.I. Reaction of β-substituted enamines with phosphorus pentachloride. Russ. J. Gen. Chem. 1985, 55, 2626–2627. [Google Scholar]
  43. Pensionerova, G.A.; Rozinov, V.G.; Kalabina, A.V.; Donskih, V.I.; Glukhikh, V.I.; Ratovskiy, G.V.; Dmitrichenko, M.Y. Phosphorylation of N,N-disubstituted amides with phosphorus pentachloride. Russ. J. Gen. Chem. 1982, 52, 2491–2499. [Google Scholar]
  44. Pensionerova, G.A.; Rozinov, V.G.; Glukhikh, V.I. Phosphorylation of tertiary amides with phosphorus pentachloride. Russ. J. Gen. Chem. 1981, 51, 1201–1202. [Google Scholar]
  45. Rozinov, V.G.; Izhboldina, L.P.; Donskih, V.I.; Dolgushin, G.V. Bisphosphorylated N-vinyl-N-chlorovinylamine on the base of N-vinyl-N-butylacetamide. Russ. J. Gen. Chem. 1984, 54, 222–223. [Google Scholar]
  46. Rozinov, V.G.; Izhboldina, L.P.; Donskih, V.I.; Kalabina, A.V. Reaction of phosphorus pentachloride with enamides—A new source of 1,4-azaphosphorins. Russ. J. Gen. Chem. 1987, 57, 478–479. [Google Scholar]
  47. Rozinov, V.G.; Izhboldina, L.P.; Donskih, V.I.; Ratovskiy, G.V.; Sergienko, L.M.; Dolgushin, G.V.; Dmitrichenko, M.Y. Phosphor-containing enamines. V. Features of the formation of 1,4-azaphosphorines during phosphorylation of enamides. Russ. J. Gen. Chem. 1989, 59, 997–1018. [Google Scholar]
  48. Dmitrichenko, M.Y.; Donskih, V.I.; Rozinov, V.G.; Dolgushin, G.V.; Feshin, V.P.; Kalabina, A.V. Phosphorylation of N-methyl-N,N-diacetamide. Russ. J. Gen. Chem. 1987, 57, 1912–1913. [Google Scholar]
  49. Dmitrichenko, M.Y.; Rozinov, V.G.; Zinchenko, S.V.; Zherebtsov, A.A. 1,4-Azaphosphorine from N,N-diacetamide. Russ. J. Gen. Chem. 1990, 60, 1919–1920. [Google Scholar] [CrossRef]
  50. Larina, L.I.; Elokhina, V.N.; Yaroshenko, T.I.; Nakhmanovich, A.S.; Dolgushin, G.V. 1H, 13C, 15N and 19F NMR study of acetylation products of heterocyclic thiosemicarbazones. Magn. Reson. Chem. 2007, 45, 667–673. [Google Scholar] [CrossRef] [PubMed]
  51. Larina, L.I.; Lopyrev, V.A. Nitroazoles: Synthesis, Structure and Applications; Springer: New York, NY, USA, 2009; 446p. [Google Scholar]
  52. Larina, L.I. Tautomerism and Structure of azoles: Nuclear Magnetic Resonance Spectroscopy. Adv. Heterocycl. Chem. 2018, 124, 233–321. [Google Scholar] [CrossRef]
  53. Larina, L.I. The structure of biologically active functionalized azoles: NMR spectroscopy and quantum chemistry. Magnetochemistry 2022, 8, 52. [Google Scholar] [CrossRef]
  54. Larina, L.I. Nuclear Quadrupole Resonance Spectroscopy: Tautomerism and structure of functional azoles. Crystals 2019, 9, 366. [Google Scholar] [CrossRef] [Green Version]
  55. Dmitrichenko, M.Y.; Rozinov, V.G.; Eliseeva, G.D.; Ratovskiy, G.V.; Kushnarev, D.F.; Zhilyakov, A.V. Phosphorylation of N-acetylhydropyridazine. Russ. J. Gen. Chem. 1997, 67, 162. [Google Scholar]
  56. Kolbina, V.E.; Dolgushin, G.V.; Rozinov, V.G. Phosphorylated amino-1,3-pentendienes. Russ. J. Gen. Chem. 1992, 62, 471–472. [Google Scholar]
  57. Borisov, A.A.; Rozinov, V.G.; Bohzenkov, G.V.; Larina, L.I. Enamino ketones in the synthesis of phosphorus-containing heterocycles. Russ. J. Gen. Chem. 2004, 74, 1822–1823. [Google Scholar] [CrossRef]
  58. Borisov, A.A.; Kolbina, V.E.; Dmitrichenko, M.Y.; Larina, L.I.; Rozinov, V.G. Phosphorus-containing quinolone derived from phenylaminocrotonate. Russ. J. Gen. Chem. 2005, 75, 1676–1677. [Google Scholar] [CrossRef]
  59. Larina, L.I.; Komarova, T.N.; Rozinov, V.G. C-phosphorylation of unsaturated ketones with phosphorus pentachloride. Russ. J. Gen. Chem. 2016, 86, 512–514. [Google Scholar] [CrossRef]
  60. Larina, L.I.; Komarova, T.N.; Rozinov, V.G. Phosphorus-containing dienes derived from diacetone alcohol: Synthesis and study by NMR spectroscopy. Russ. Chem. Bull. 2015, 64, 2744–2746. [Google Scholar] [CrossRef]
  61. Rozinov, V.G.; Kolbina, V.E.; Dmitrichenko, M.Y. Features of phosphorylation of oximes and amides under mild conditions. Russ. J. Gen. Chem. 1997, 67, 519–520. [Google Scholar]
  62. Rozinov, V.G.; Dmitrichenko, M.Y.; Larina, L.I.; Schmidt, E.Y.; Michaleva, A.I. Unsaturated organophosphorus compounds based on O-vinyl oximes. Russ. J. Gen. Chem. 2001, 71, 1041–1042. [Google Scholar] [CrossRef]
  63. Sokolov, F.; Brusko, V.; Safin, D.; Cherkasov, R.; Zabirov, N. Coordination diversity of N-phosphorylated amides and ureas towards viiib group cations. Transit. Met. Chem. New Res. 2008, 101–150. Available online: https://dspace.kpfu.ru/xmlui/handle/net/141420 (accessed on 12 April 2023).
  64. Kantlehner, W.; Kreß, R.; Mezger, J.; Ziegler, G. Orthoamide und Iminiumsalze, LXXXVIII. Synthese N,N,N′,N′,N″,N″-persubstituierter Guanidiniumsalze aus N,N′-persubstituierten Harnstoff/Säurechlorid-Addukten. Z. Naturforsch. B 2015, 70, 9–27. [Google Scholar] [CrossRef]
  65. Bhagat, V.; O’Brien, E.; Zhou, J.; Becker, M.L. Caddisfly inspired phosphorylated poly(ester-urea)-based degradable bone adhesives. Biomacromolecules 2016, 17, 3016–3024. [Google Scholar] [CrossRef]
  66. Bhagat, V.; Matthew, L.; Becker, M.L. Degradable adhesives for surgery and tissue engineering. Biomacromolecules 2017, 18, 3009–3039. [Google Scholar] [CrossRef] [Green Version]
  67. Hawkins, L.J.; Wang, M.; Zhang, B.; Xiao, Q.; Wang, H.; Storey, K.B. Glucose and urea metabolic enzymes are differentially phosphorylated during freezing, anoxia, and dehydration exposures in a freeze tolerant frog. Comparat. Biochem. Physiol. Part D 2019, 30, 1–13. [Google Scholar] [CrossRef] [PubMed]
  68. Bakibaev, A.A.; Zhumanov, K.B.; Panshina, S.Y.; Gorbin, S.I.; Malkov, V.S.; Tsoy, I.G.; Massalimova, B.K.; Matniyazova, G.K.; Baybazarova, E.A. Synthesis methods of phosphorylated carbamide containing acyclic and heterocyclic compounds. Bull. Karag. Univer. Chem. Ser. 2019, 95, 115–117. [Google Scholar] [CrossRef]
  69. Kesseli, F.P.; Lauer, C.S.; Baker, I.; Mirica, K.A.; Van Citters, W. Identification of a calcium phosphoserine coordination network in an adhesive organo-apatitic bone cement system. Acta Biomater. 2020, 105, 280–289. [Google Scholar] [CrossRef] [PubMed]
  70. Shokri, M.; Dalili, F.; Kharaziha, M.; Eslaminejad, B.M.; Tafti, H.A. Strong and bioactive bioinspired biomaterials, next generation of bone adhesives. Adv. Coll. Interface Sci. 2022, 305, 102706. [Google Scholar] [CrossRef] [PubMed]
  71. Bingol, H.B.; Bender, J.C.M.E.; Opsteen, J.A.; Leeuwenburgh, S.C.G. Bone adhesive materials: From bench to bedside. Mater. Today Bio 2023, 19, 100599. [Google Scholar] [CrossRef]
  72. Ma, Y.; Zhang, X.; Ma, C.; Xia, W.; Hu, L.; Dong, X.; Xiong, Y. Electrochemically oxidative phosphating of aldehydes and ketones. J. Org. Chem. 2023, 88, 4264–4272. [Google Scholar] [CrossRef]
  73. Rozinov, V.G.; Dmitrichenko, M.Y.; Donskih, V.I.; Dolgushin, G.V.; Feshin, V.P. 1,3,4-Diazaphosphorine from N-acetylurea. Russ. J. Gen. Chem. 1987, 57, 228–229. [Google Scholar]
  74. Dmitrichenko, M.Y.; Rozinov, V.G.; Donskih, V.I.; Ratovskiy, G.V.; Sergienko, L.M.; Dolgushin, G.V.; Valeev, R.B. Phosphorus-containing enureas. I. Phosphorylation of N-acetyl-N,N’-dimethylureas with phosphorus pentachloride. Russ. J. Gen. Chem. 1988, 58, 2252–2261. [Google Scholar]
  75. Dmitrichenko, M.Y.; Dolgushin, G.V.; Rozinov, V.G.; Donskih, V.I. Phosphorylation of N-methyl-N’-acetylurea. Russ. J. Gen. Chem. 1990, 60, 462–463. [Google Scholar]
  76. Dmitrichenko, M.Y.; Rozinov, V.G.; Donskih, V.I.; Ratovskiy, G.V.; Dolgushin, G.V.; Vitkovsky, V.Y.; Shilin, S.V. 1,5,2-Diazaphosphorines. II. Synthesis and peculiarities of cyclization of phosphorus-containing azabutadienes on the base of acetyl- and chloroacetylureas. Russ. J. Gen. Chem. 1991, 61, 643–656. [Google Scholar]
  77. Dmitrichenko, M.Y.; Rozinov, V.G.; Donskih, V.I.; Ratovskiy, G.V.; Efremov, V.G.; Dolgushin, G.V. 1,5,2-Diazaphosphorines. III. Phosphorus-containing azabutadienes and heterocycles from N-methyl-N’-acetylurea. Russ. J. Gen. Chem. 1992, 62, 306–318. [Google Scholar]
  78. Dmitrichenko, M.Y.; Rozinov, V.G.; Ratovskiy, G.V.; Dolgushin, G.V.; Donskih, V.I.; Sergienko, L.M. 1,5,2-Diazaphosphorines. IV. Structure and chemical transformation of 2,2,4,6-tetrachloro-2,2-dihydro-1,5,2-diazaphosphorine. Russ. J. Gen. Chem. 1993, 63, 1976–1989. [Google Scholar]
  79. Dmitrichenko, M.Y.; Donskih, V.I.; Rozinov, V.G.; Dolgushin, G.V.; Ratovskiy, G.V.; Efremov, V.G.; Rubkina, V.V. Phosphorus-containing chloroethenylchloroformamidines from N-acetylureas. Russ. J. Gen. Chem. 1989, 59, 227–228. [Google Scholar]
  80. Dmitrichenko, M.Y.; Rozinov, V.G.; Ratovskiy, G.V.; Dolgushin, G.V.; Donskih, V.I.; Sergienko, L.M. 1,5,2-Diazaphosphorines. V. Features of amination of heterocycles of varying degrees of unsaturation. Russ. J. Gen. Chem. 1995, 65, 426–430. [Google Scholar]
  81. Larina, L.I.; Dmitrichenko, M.Y.; Rozinov, V.G.; Benedyuk, A.V. Phosphorylation of N-acetyl-N,N’-ethyleneurea with phosphorus pentachloride. Russ. J. Gen. Chem. 2005, 75, 484–485. [Google Scholar] [CrossRef]
  82. Larina, L.I.; Rozinov, V.G.; Chernyshev, K.A. The Products of Phosphorylation of N,N-Dialkylureas and Dialkylcyanamides with Phosphorus Pentachloride. NMR Spectroscopy Study. Russ. J. Gen. Chem. 2012, 82, 72–76. [Google Scholar] [CrossRef]
  83. Chernyshev, K.A.; Larina, L.I.; Rozinov, V.G.; Krivdin, L.B. Quantum-chemical calculations of NMR chemical shifts of organic molecules: IX. Electronic structure of [dimethylamino(chloro)methylideneamino]trichlorophosphonium hexachlorophosphate: Azomethine or azophosphine? Russ. J. Org. Chem. 2013, 49, 195–197. [Google Scholar] [CrossRef]
  84. Dmitrichenko, M.Y.; Rozinov, V.G.; Donskih, V.I. Synthesis of 1-methyl-2,2,2-trichloro-3-(1,1,2,2-tetrachloroethyl)-1,3-diaza-2-phosphetidinone. Russ. J. Gen. Chem. 1989, 59, 715–716. [Google Scholar]
  85. Turchaninov, V.K.; Dolgushin, G.V.; Dmitrichenko, M.Y.; Larina, L.I. AM1 study of photoelectron spectra. 9. 2,2,4,6-Tetrachloro-2,2-dihydro-1,5,2-diazaphosphorinine. Russ. Chem. Bull. 1996, 45, 781–785. [Google Scholar] [CrossRef]
  86. Ghanadpour, M.; Carosio, F.; Larsson, P.T.; Wagberg, L. Phosphorylated 619 Cellulose Nanofibrils: A Renewable Nanomaterial for the Preparation of Intrinsically 620 Flame-Retardant Materials. Biomacromolecules 2015, 16, 3399–3410. [Google Scholar] [CrossRef]
  87. Long, H.; Huang, C.; Zheng, Y.-T.; Li, Z.-Y.; Jie, L.-H.; Song, J.; Zhu, S.; Xu, H.-C. Electrochemical C–H phosphorylation of arenes in continuous flow suitable for late-stage functionalization. Nat. Commun. 2021, 12, 6629–6636. [Google Scholar] [CrossRef]
  88. Rol, F.; Sillard, C.; Bardet, M.; Yarava, J.R.; Emsley, L.; Gablin, C.; Leonard, D.; Belgacem, N.; Bras, J. Cellulose phosphorylation comparison and analysis of phosphorate position on cellulose fibers. Carbohydr. Polym. 2020, 229, 115294–115332. [Google Scholar] [CrossRef] [PubMed]
  89. Liu, B.; Li, J.; Hu, Y.; Chen, Q.; Jiu, Y.; Li, S.; Maruoka, K.; Huo, Y.; Zhang, H.-L. Visible-Light-Induced α-C(sp 3)–H Phosphinylation of Unactivated Ethers under Photocatalyst- and Additive-Free Conditions. J. Org. Chem. 2022, 87, 11281–11291. [Google Scholar] [CrossRef] [PubMed]
  90. Mei, Y.; Yan, Z.; Liu, L.L. Facile synthesis of the dicyanophosphide anion via electrochemical activation of white phosphorus: An avenue to organophosphorus compounds. J. Amer. Chem. Soc. 2022, 144, 1517–1522. [Google Scholar] [CrossRef] [PubMed]
  91. Bonfant, G.; Balestri, D.; Perego, J.; Comotti, A.; Bracco, S.; Koepf, M.; Gennari, M.; Marchio, L. Phosphine oxide porous organic polymers incorporating cobalt(II) ions: Synthesis, characterization, and investigation of H 2 production. ASC Omega 2022, 7, 6104–6112. [Google Scholar] [CrossRef]
  92. Mondal, A.; Thiel, N.O.; Dore, R.; Feringa, B.L. P-chirogenic phosphorus compounds by stereoselective Pd-catalysed arylation of phosphoramidites. Nat. Catal. 2022, 5, 10–19. [Google Scholar] [CrossRef]
Scheme 1. The formation of hexachlorophosphorates of 2-aminoethenyltrichloro-phosphonium (13) and hexachlorophosphorates of 2-amino-1-chloroethenyltrichlorophosphorates (46).
Scheme 1. The formation of hexachlorophosphorates of 2-aminoethenyltrichloro-phosphonium (13) and hexachlorophosphorates of 2-amino-1-chloroethenyltrichlorophosphorates (46).
Ijms 24 09646 sch001
Scheme 2. The formation of 2-(diethylamino)ethenylphosphonius acid (7) and transformation of enaminotrichlorophosphonium hexachlorophosphates (16) into dichloroanhydrides of phosphonous acids (813).
Scheme 2. The formation of 2-(diethylamino)ethenylphosphonius acid (7) and transformation of enaminotrichlorophosphonium hexachlorophosphates (16) into dichloroanhydrides of phosphonous acids (813).
Ijms 24 09646 sch002
Scheme 3. Formation of phosphorylated chloramine (14) and chloranhydride of bis-[2(N-methyl,N-phenylamino)ethyl] phosphonous acid (15).
Scheme 3. Formation of phosphorylated chloramine (14) and chloranhydride of bis-[2(N-methyl,N-phenylamino)ethyl] phosphonous acid (15).
Ijms 24 09646 sch003
Scheme 4. The phosphorylation of N,N-diethylaniline with phosphorus pentachloride.
Scheme 4. The phosphorylation of N,N-diethylaniline with phosphorus pentachloride.
Ijms 24 09646 sch004
Scheme 5. The formation of 1-chloro-2-ethoxyethenylphosphonyl dichloride (19) and 2-ethoxyethynylphosphonyl chloride (20).
Scheme 5. The formation of 1-chloro-2-ethoxyethenylphosphonyl dichloride (19) and 2-ethoxyethynylphosphonyl chloride (20).
Ijms 24 09646 sch005
Scheme 6. The formation of 1,4-oxaphosphinines 21 and 22.
Scheme 6. The formation of 1,4-oxaphosphinines 21 and 22.
Ijms 24 09646 sch006
Scheme 7. The formation of dichlorophosphines 23 and 24.
Scheme 7. The formation of dichlorophosphines 23 and 24.
Ijms 24 09646 sch007
Scheme 8. The formation of 2-diphenylaminoethene-1chloro-1-phosphonic acid dihydroanhydride (13a).
Scheme 8. The formation of 2-diphenylaminoethene-1chloro-1-phosphonic acid dihydroanhydride (13a).
Ijms 24 09646 sch008
Scheme 9. The phosphorylated enamides 2542.
Scheme 9. The phosphorylated enamides 2542.
Ijms 24 09646 sch009
Scheme 10. The formation of 2-(N-3,3-dichloro-1-hydroxyisoindolinyl)-ethenephosphonic acid dichloride (43).
Scheme 10. The formation of 2-(N-3,3-dichloro-1-hydroxyisoindolinyl)-ethenephosphonic acid dichloride (43).
Ijms 24 09646 sch010
Scheme 11. The formation of hexachlorophosphorate 44 and oxazaphosphorine 45.
Scheme 11. The formation of hexachlorophosphorate 44 and oxazaphosphorine 45.
Ijms 24 09646 sch011
Scheme 12. The formation of hexachlorophosphorates 46 and 48 and phosphonic dichloride 47 and 49 of N-substituted carbazole.
Scheme 12. The formation of hexachlorophosphorates 46 and 48 and phosphonic dichloride 47 and 49 of N-substituted carbazole.
Ijms 24 09646 sch012
Scheme 13. The phosphorylation of N-methyl-N-phenyl- and N,N-diphenylacetamides.
Scheme 13. The phosphorylation of N-methyl-N-phenyl- and N,N-diphenylacetamides.
Ijms 24 09646 sch013
Scheme 14. The phosphorylation of N-alkenylenamides.
Scheme 14. The phosphorylation of N-alkenylenamides.
Ijms 24 09646 sch014
Scheme 15. The formation of hexachlorophosphorates of 1,4-azaphosphoniarines (6367) and 1,4-azaphosphorines (6872).
Scheme 15. The formation of hexachlorophosphorates of 1,4-azaphosphoniarines (6367) and 1,4-azaphosphorines (6872).
Ijms 24 09646 sch015
Scheme 16. The formation of 1,4-azaphosphorinonium hexachlorophosphorates 73 and 75 and 1,4-azaphosphorines 74 and 76.
Scheme 16. The formation of 1,4-azaphosphorinonium hexachlorophosphorates 73 and 75 and 1,4-azaphosphorines 74 and 76.
Ijms 24 09646 sch016
Scheme 17. The phosphorylation of N-acetylhydropyridazine.
Scheme 17. The phosphorylation of N-acetylhydropyridazine.
Ijms 24 09646 sch017
Scheme 18. The phosphorylation of enamino ketones.
Scheme 18. The phosphorylation of enamino ketones.
Ijms 24 09646 sch018
Scheme 19. The phosphorylation of ethyl ether of phenylaminocrotonic acid.
Scheme 19. The phosphorylation of ethyl ether of phenylaminocrotonic acid.
Ijms 24 09646 sch019
Scheme 20. Chlorophosphorylation of unsaturated ketones.
Scheme 20. Chlorophosphorylation of unsaturated ketones.
Ijms 24 09646 sch020
Scheme 21. Chlorophosphorylation of diacetone alcohol.
Scheme 21. Chlorophosphorylation of diacetone alcohol.
Ijms 24 09646 sch021
Scheme 22. The phosphorylation of acetaldoxime.
Scheme 22. The phosphorylation of acetaldoxime.
Ijms 24 09646 sch022
Scheme 23. The phosphorylation of O-vinyl oximes.
Scheme 23. The phosphorylation of O-vinyl oximes.
Ijms 24 09646 sch023
Scheme 24. The chlorophosphorylation of N-acetylurea.
Scheme 24. The chlorophosphorylation of N-acetylurea.
Ijms 24 09646 sch024
Scheme 25. The formation of 2,2,3,4,6-pentachloro-2,2-dihydro-1,5,2-diazaphosphorine (104).
Scheme 25. The formation of 2,2,3,4,6-pentachloro-2,2-dihydro-1,5,2-diazaphosphorine (104).
Ijms 24 09646 sch025
Scheme 26. The formation of dimeric diazaphosphorine (103a).
Scheme 26. The formation of dimeric diazaphosphorine (103a).
Ijms 24 09646 sch026
Scheme 27. The formation of diazaphosphorines 105 and 106.
Scheme 27. The formation of diazaphosphorines 105 and 106.
Ijms 24 09646 sch027
Scheme 28. The formation of phosphetidinone 107.
Scheme 28. The formation of phosphetidinone 107.
Ijms 24 09646 sch028
Scheme 29. The formation of six-membered heterocycles 108115.
Scheme 29. The formation of six-membered heterocycles 108115.
Ijms 24 09646 sch029
Scheme 30. The chlorophosphorylation of 1-acetylimidazolidin-2-one.
Scheme 30. The chlorophosphorylation of 1-acetylimidazolidin-2-one.
Ijms 24 09646 sch030
Table 1. The 1H, 13C and 31P NMR chemical shifts (δ, ppm) and coupling constants (J, Hz) of phosphorylated enamines 113.
Table 1. The 1H, 13C and 31P NMR chemical shifts (δ, ppm) and coupling constants (J, Hz) of phosphorylated enamines 113.
No.Structure1H a13C31P b
1Ijms 24 09646 i0018.00 dd, 1H, N–CH=
3JPH = 25.9, 3JHH = 13.7
5.65 dd, 1H, P–CH=
2JPH = 33.7, 3JHH = 13.7
153.25 d, N–CH=, 2JPC = 47.3
81.47 d, P–CH=, 1JPC = 195.4
89.9 dd PCl 3 +
2JPH = 33.7
3JPH = 25.9
2Ijms 24 09646 i0027.81 dd, 1H, N–CH=
3JPH = 23.8, 3JHH = 13.9
5.50 dd, 1H, P–CH=
2JPH = 34.0, 3JHH = 13.9
155.12 d, N–CH=, 2JPC = 48.1
80.92 d, P–CH=, 1JPC = 200.2
80.1 dd PCl 3 +
2JPH = 34.0
3JPH = 23.8
3Ijms 24 09646 i0037.60 dd, 1H, N–CH=
3JPH = 21.4, 3JHH = 13.9
4.55 dd, 1H, P–CH=
2JPH = 34.3, 3JHH = 13.9
150.40 d, N–CH=, 2JPC = 52.7
80.29 d, P–CH=, 1JPC = 200.8
79.8 dd PCl 3 +
2JPH = 34.3
3JPH = 21.4
4Ijms 24 09646 i0047.50 d, 1H, N–CH=
3JPH = 14.1
158.11 d, N–CH=, 2JPC = 37.5
77.26 d, P–CH=, 1JPC = 187.1
80.7 d PCl 3 +
3JPH = 14.1
5Ijms 24 09646 i0057.65 d, 1H, N–CH=
3JPH = 13.9
157.20 d, N–CH=, 2JPC = 35.4
77.82 d, P–CH=, 1JPC = 179.5
82.0 d PCl 3 +
3JPH = 13.9
6Ijms 24 09646 i0067.50 d, 1H, N–CH=
3JPH = 13.7
159.21 d, N–CH=, 2JPC = 30.9
75.64 d, P–CH=, 1JPC = 159.6
72.0 d PCl 3 +
3JPH = 13.7
7Ijms 24 09646 i0077.25 dd, 1H, N–CH=
3JPH = 18.3, 3JHH = 13.1
5.36 dd, 1H, P–CH=
2JPH = 9.0, 3JHH = 13.1
3.27 m 4H, CH2 ,
1.29 m 3HCH3
156.80 d, N–CH=, 2JPC = 31.0
85.11 d, P–CH=, 1JPC = 201.2
50.67 s, CH2
14.86 s, CH3
160.0 dd PCl2
2JPH = 9.0
3JPH = 18.3
8Ijms 24 09646 i0087.06 dd, 1H, N–CH=
3JPH = 21.5, 3JHH = 13.7
5.12 dd, 1H, P–CH=
2JPH = 25.8, 3JHH = 13.7
3.51 m 4H, CH2,
1.27 m 6H, CH3
150.88 d, N–CH=, 2JPC = 38.9
85.45 d, P–CH=, 1JPC = 208.9
43.51 s, CH2
13.84 s, CH3
35.0 dd POCl2
2JPH = 25.8
3JPH = 21.5
9Ijms 24 09646 i0097.25 dd, 1H, N–CH=
3JPH = 21.2, 3JHH = 14.2
7.1–7.3 m, 5H, Ph
5.20 dd, 1H, P–CH=
2JPH = 25.5, 3JHH = 14.2
4.62 s, 2H, CH2-Ph
3.54 m, 2H, CH2
1.28 m, 3H, CH3
149.27 d, N–CH=, 2JPC = 37.1
86.23 d, P–CH=, 1JPC = 208.9
125.20 m, Ph
51.60 s, CH2-Ph
49.70 s, CH2
15.23 s, CH3
36.2 dd POCl2
2JPH = 25.5
3JPH = 21.2
10Ijms 24 09646 i0107.20 dd, 1H, N–CH=
3JPH = 21.8, 3JHH = 14.1
7.1–7.3 m, 5H, Ph
4.95 dd, 1H, P–CH=
2JPH = 26.4, 3JHH = 14.1
3.73 m, 2H, CH2
1.25 m, 3H, CH3
150.10 d, N–CH=, 2JPC = 32.8
85.32 d, P–CH=, 1JPC = 201.4
120.40 m, Ph
48.34 s, CH2
15.00 s, CH3
37.0 dd POCl2
2JPH = 26.4
3JPH = 21.8
11Ijms 24 09646 i0117.15 d, 1H, N–CH=
3JPH = 11.6
3.45 m, 4H, CH2
1.18 m, 6H, CH3
145.19 d, N–CH=, 2JPC = 39.4
1JCH=165.7
88.64 d, P–CH=, 1JPC = 212.5
46.91 s, CH2
14.84 s, CH3
36.1 d POCl2
3JPH = 11.6
12Ijms 24 09646 i0127.34 d, 1H, N–CH=
3JPH = 12.1
7.2–7.4 m, 5H, Ph
5.37 s, 2H, CH2-Ph
3.53 m, 2H, CH2
1.23 m, 3H, CH3
148.95 d, N–CH=, 2JPC = 39.8
86.02 d, P–CH=, 1JPC = 210.3
124.24 m, Ph
54.26 s, CH2 -Ph
49.15 s, CH2
14.65 s, CH3
35.9 d POCl2
3JPH = 12.1
13Ijms 24 09646 i0137.51 d, 1H, N–CH=
3JPH = 11.1
7.1–7.3 m, 5H, Ph
3.55 m, 2H, CH2
1.20 m, 3H, CH3
151.44 d, N–CH=, 2JPC = 33.5
84.57 d, P–CH=, 1JPC = 200.4
127.40 m, Ph
51.34 s, CH2
15.07 s, CH3
34.8 d POCl2
3JPH = 11.1
a It is impossible to identify the signals of protons of both ethyl and phenyl groups in compounds 16 (hexachlorophosphorates), since they and the tertiary amine hydrochlorides present in the crystalline reaction product have similar ethyl and phenyl substituents (δ1H: 1.20–1.28, 3.2–3.5 and 7.2–7.5 ppm, respectively, CH3, CH2, Ph). b The phosphorus signal of PCl 6 anion in the 31P NMR spectra of organyltrichlorophosphonium hexachlorophosphorates 16 appears in the region −(297–298) ppm. For (13a) Ph2N—CH=C(Cl)-POCl2, the value of δ 31P is 34.6 ppm with constant 3JPH = 11.6 Hz.
Table 2. The 1H, 13C and 31P NMR chemical shifts (δ, ppm) and coupling constants (J, Hz) of phosphorylated enamines 1418.
Table 2. The 1H, 13C and 31P NMR chemical shifts (δ, ppm) and coupling constants (J, Hz) of phosphorylated enamines 1418.
No.Structure1H13C31P
14Ijms 24 09646 i0147.47 d, 1H, N–CH=
3JPH = 11.8
7.3 m, 5H, Ph
3.02 s, 3H, CH3
148.40 d, N–CH=, 2JPC = 30.5
123.4 m Ph
81.37 d, P–CH=, 1JPC = 201.0
18.07 CH3
34.2 d POCl2
3JPH = 11.8
15Ijms 24 09646 i0157.63 dd, 1H, N–CH=
3JPH = 16.4, 3JHH = 12.0
7.18 m, 5H, Ph
6.68 dd, 1H, P–CH=
3JHH = 12.0, 2JPH = 5.1
3.11 s, 3H, CH3
150.20 d, N–CH=, 2JPC = 36.9
123.40 m Ph
88.11 d, P–CH=, 1JPC = 140.9
15.12 CH3
35.6 tt POCl
3JPH = 16.4
2JPH = 5.1
16Ijms 24 09646 i0167.59 dd, 1H, O–CH=
3JPH = 31.4, 3JHH = 6.4
5.68 dd, 1H, P–CH=
3JHH = 6.4, 2JPH = 1.5
149.27 d, O–CH =
120.23 d, P–CH= 1JPC = 120.9
10.7 tt POCl
3JPH = 31.4
2JPH = 1.5
17Ijms 24 09646 i0177.28 dd, 1H, N–CH=
3JPH = 22.9, 3JHH = 13.8
7.30 m, 4H, Ph
4.90 dd, 1H, P–CH=
2JPH = 33.6, 3JHH = 13.8
3.53 m, 2H, CH2
1.21 m, 3H, CH3
156.10 d, N–CH=, 2JPC = 32.8
129.24 m Ph
85.32 d, P–CH=, 1JPC = 201.4
48.34 CH2
15.00 CH3
82.7 dd PCl 3 +
2JPH = 33.6
3JPH = 22.9
−297.9 br s
PCl 6
18Ijms 24 09646 i0187.45 d, 1H, N–CH=
3JPH = 14.5
7.24 m, 4H, Ph
3.41 m, 2H, CH2
1.16 m, 3H, CH3
145.19 d, N–CH=, 2JPC = 39.6
86.64 d, P–CH=, 1JPC = 200.5
46.91 CH2
14.84 CH3
74.0 d PCl3
3JPH = 14.5
−298.0 br s
PCl 6
Table 3. The 1H, 13C and 31P NMR chemical shifts (δ, ppm) and coupling constants (J, Hz) of phosphorylated compounds 1924.
Table 3. The 1H, 13C and 31P NMR chemical shifts (δ, ppm) and coupling constants (J, Hz) of phosphorylated compounds 1924.
No.Structure1H13C31P
19Ijms 24 09646 i0197.34 d, 1H, CH=
3JPH = 18.3
4.50 m, 2H, CH2
1.52 m, 3H, CH3
135.30 s, CH=
125.76 d, CCl
1JPC = 92.2
42.54 s, OCH2
15.32 s, CH3
80.1 d
3JPH = 18.3
20Ijms 24 09646 i0204.42 m, 2H, CH2
1.50 m, 3H, CH3
102.30 CO
94.30 CP
44.50 OCH2
13.40 CH3
6.0 s
21Ijms 24 09646 i0218.15 ddd, 1H, O–CH=
3JP1H = 57.2, 5JP2H = 19.1
3JHH = 11.0
7.59 ddd, 1H, P–CH=
4JP2H = 40.4, 2JP1H = 27.0
3JHH = 11.0
4.30 m, 2H, CH2
1.55 m, 3H, CH3
138.5 d, C-2
1JP1C2 = 101.2
106.4 dd, C-1
1JP1C1 = 157.7
1JP2C1 = 67.5
28.3 ddd, P1
3JP1H = 57.2,
2JPP = 56.5, 2JP1H = 27.0
21.4 ddd P2
2JPP = 56.5, 4JP2H = 40.4
5JP2H = 19.1
22Ijms 24 09646 i0228.20 ddd, 1H, O–CH=
3JP1H = 62.0, 5JP2H = 16.8
3JHH = 11.0
7.50 d, 1H, P–CH=
4JP2H = 36.6, 2JP1H = 27.5
3JHH = 11.0
139.1 d, C-2
1JP1C2 = 100.0
108.7 dd, C-1
1JP1C1 = 150.0
1JP2C1 = 68.0
25.6 ddd P1
3JP1H = 62.0
2JPP = 48.8, 2JP1H = 27.5
21.9 ddd P2
2JPP = 48.8, 4JP2H = 36.6
5JP2H = 16.8
23Ijms 24 09646 i0237.58 dd, 1H, P–CH=
3JHH = 15.3, 2JPH = 9.2
7.50 m, 5H, Ph
7.30 dd, 1H, Ph–CH=
3JPH = 19.1, 3JHH = 15.3
142.2 d, C-P
1JPC = 98.0
112.4 d, C-Ph
2JPC = 35.0
162.8 dd PCl2
3JPH = 19.1
2JPH = 9.2
24Ijms 24 09646 i0247.50 m, 5H, Ph
7.40 d, 1H, Ph–CH=
3JPH = 22.1
142.2 d, C-P
1JPC = 99.0
118.3 d, C-Ph
2JPC = 33.0
157.8 d PCl2
3JPH = 22.1
Table 4. The 1H and 31P NMR chemical shifts (δ, ppm) and coupling constants (J, Hz) of phosphorylated enamides 2533 (CH3NO2) and 3442 (CDCl3).
Table 4. The 1H and 31P NMR chemical shifts (δ, ppm) and coupling constants (J, Hz) of phosphorylated enamides 2533 (CH3NO2) and 3442 (CDCl3).
Ijms 24 09646 i025
RNo.31P a PCl 3 +   No.1H31P, POCl2
Ijms 24 09646 i0262594.1 dd
2JPH = 40.2
3JPH = 25.9
348.50 dd, 1H, N–CH=
3JPH = 21.8, 3JHH = 15.0
7.32 m, 5H, Ph
4.75 dd, 1H, P–CH=
2JPH = 27.8, 3JHH = 15.0
1.91 3H, CH3
34.7 dd
2JPH = 27.8
3JPH = 21.8
Ijms 24 09646 i0272693.7 dd
2JPH = 40.0
3JPH = 26.0
358.54 dd, 1H, N–CH=
3JPH = 21.9, 3JHH = 15.0
7.35 m, 10H, Ph
5.12 dd, 1H, P–CH=
2JPH = 27.8, 3JHH = 15.0
33.7 dd
2JPH = 27.8
3JPH = 21.9
Ijms 24 09646 i0282794.0 dd
2JPH = 39.6
3JPH = 26.6
367.81 dd, 1H, N–CH=
3JPH = 21.9, 3JHH = 15.2
7.20 m, 4H, Ph
5.00 dd, 1H, P–CH=
2JPH = 27.8, 3JHH = 15.2
2.15, 3H, CH3
1.98 3H, CH3
34.7 dd
2JPH = 27.8
3JPH = 21.9
Ijms 24 09646 i0292894.0 dd
2JPH = 40.0
3JPH = 26.1
378.32 dd, 1H, N–CH=
3JPH = 21.5, 3JHH = 15.1
7.33 m, 4H, Ph
4.88 dd, 1H, P–CH=
2JPH = 28.5, 3JHH = 15.1
4.15 3H, OMe;
1.96 3H, CH3
34.1 dd
2JPH = 28.5
3JPH = 21.5
Ijms 24 09646 i0302993.5 dd
2JPH = 40.0
3JPH = 26.0
388.77 dd, 1H, N–CH=
3JPH = 21.6, 3JHH = 14.6
8.1–7.4 m, 7H, Napht
4.72 dd, 1H, P–CH=
2JPH = 27.9, 3JHH = 14.6
1.88 3H, CH3
33.7 dd
2JPH = 27.9
3JPH = 21.6
Ijms 24 09646 i0313092.1 dd
2JPH = 39.7
3JPH = 26.8
397.84 dd, 1H, N–CH=
3JPH = 21.7, 3JHH = 15.1
5.30 dd, 1H, P–CH=
2JPH = 26.9, 3JHH = 15.1
3.53 m, 2H, NCH2
2.49 m, 2H, CH2CO
2.1 m, 2H, CH2
33.9 dd
2JPH = 26.9
3JPH = 21.7
Ijms 24 09646 i0323192.5 dd
2JPH = 40.4
3JPH = 26.3
408.12 dd, 1H, N–CH=
3JPH = 24.0, 3JHH = 15.2
5.41 dd, 1H, P–CH=
2JPH = 26.5, 3JHH = 15.2
3.55 m, 2H, NCH2
2.58 m, 2H, CH2CO
1.6 m, 6H, CH2
37.3 dd
2JPH = 26.5
3JPH = 24.0
Ijms 24 09646 i0333298.9 dd
2JPH = 40.0
3JPH = 29.4
417.53 dd, 1H, N–CH=
3JPH = 26.2, 3JHH = 15.6
7.31 dd, 1H, P–CH=
2JPH = 30.5, 3JHH = 15.6
2.84 m, 4H, CH2CO
33.1 dd
2JPH = 30.5
3JPH = 26.2
Ijms 24 09646 i0343394.0 dd
2JPH = 39.6
3JPH = 26.6
427.64 dd, 1H, N–CH=
3JPH = 25.4, 3JHH = 15.6
7.85 m, 4H, Ph
7.12 dd, 1H, P–CH=
2JPH = 30.1, 3JHH = 15.6
34.4 dd
2JPH = 30.1
3JPH = 25.4
a The phosphorus signal of PCl 6 anion in the 31P NMR spectra of hexachlorophosphorates 2533 appears in the region −(296–298) ppm; the 3JHH value varies within 15–16 Hz.
Table 5. The 1H, 13C and 31P NMR chemical shifts (δ, ppm) and coupling constants (J, Hz) of phosphonic dichlorides 43 and 45 (CDCl3) and hexachlorophosphorate (44) (CH3NO2).
Table 5. The 1H, 13C and 31P NMR chemical shifts (δ, ppm) and coupling constants (J, Hz) of phosphonic dichlorides 43 and 45 (CDCl3) and hexachlorophosphorate (44) (CH3NO2).
No.R1H13C31P
43Ijms 24 09646 i0357.80 m, 4H, H-(4–7)
7.10 dd, 1H, N–CH=
3JPH = 57.1, 3JHH = 10.5
6.28 dd, 1H, P–CH=
2JPH = 26.2, 3JHH = 10.5
157.70 C-1
140.98 C-9
131.14 C-4
128.20 d, N–CH=
2JPC = 8.5
127.53 C-7
121.05 C-8
119.64 C-6
118.57 C-5
114.55 d, P–CH=
1JPC = 151.4
86.77 C-3
23.3 dd
2JPH = 26.2
3JPH = 57.1
44Ijms 24 09646 i0367.78 dd, 1H, N–CH=
3JPH = 40.8, 3JHH = 10.0
6.44 dd, 1H, P–CH=
2JPH = 20.0, 3JHH = 10.0
6.22 d, 1H, O–CH=
3JHH = 6.0
142.60 dd, C-9
2JPC = 23.3 d, 3JPC = 10.0
133.40 d, C-4, 2JPC = 6.0
117.84 d, C-6, 2JPC = 16.2
113.75 d, C-7, 1JPC = 185.3
101.50 dd, C-8
2JPC = 16.6 d, 3JPC = 5.5
83.5 dd PCl 3 +
3JPH = 6.0
5JPP = 2.8
11.1 ddd POCl
3JPH = 40.8
2JPH = 20.0
4JPH = 2.8
−297.7 br s
PCl 6
45Ijms 24 09646 i0377.67 dd, 1H, N–CH=
3JPH = 40.5, 3JHH = 10.5
6.17 dd, 1H, P–CH=
2JPH = 18.6, 3JHH = 10.5
6.23 d, 1H, O–CH=
3JHH = 6.1
139.82 dd, C-9
2JPC = 25.6 d, 3JPC = 11.0
135.29 d, C-4 2JPC = 6.1
115.95 d, C-6, 2JPC = 18.3
114.80 d, C-7, 1JPC = 194.7
96.50 dd, C-8
2JPC = 14.6 d, 3JPC = 4.5
18.6 dd POCl2
3JPH = 6.1
5JPP = 2.8
10.4 ddd POCl
3JPH = 40.5
2JPH = 18.6
4JPH = 2.8
Table 6. The 1H and 31P NMR chemical shifts (δ, ppm) and coupling constants (J, Hz) of phosphorylated acetamides 5357 (CH3NO2) and 5862 (CDCl3).
Table 6. The 1H and 31P NMR chemical shifts (δ, ppm) and coupling constants (J, Hz) of phosphorylated acetamides 5357 (CH3NO2) and 5862 (CDCl3).
Ijms 24 09646 i038
No.R1R31P, POCl231P a, PCl 3 +   No.1H31P, P1OCl231P, POCl2
53HPr31.2 dd
2JPH = 23.4
3JPH = 23.4
82.7 d
2JPH = 32.2
587.50 dd, 1H, N–CH=
3JPH = 23.0, 3JHH = 15.0
5.75 dd, 1H, P–CH=
2JPH = 17.0
5.48 dd, 1H, P1–CH=
2JPH = 23.0, 3JHH = 15.0
3.64 t, NCH2, 3JHH = 6.7
1.49 m, 2H, CH2
0.98 t, 3H, CH3, 3JHH = 6.9
33.7 dd
2JPH = 23.0
3JPH = 23.0
25.4 d
2JPH = 17.0
54HBu30.7 dd
2JPH = 22.4
3JPH = 22.4
81.5 d
2JPH = 30.3
597.73 dd, 1H, N–CH=
3JPH = 21.8, 3JHH = 14.8
5.78 d, 1H, P–CH=
2JPH = 17.1
5.40 dd, 1H, P1–CH=
2JPH = 23.9, 3JHH = 14.8
3.62 t, NCH2, 3JHH = 6.8
1.44 m, 4H, (CH2)2
0.92 t, 3H, CH3, 3JHH = 6.7
32.5 dd
2JPH = 23.9
3JPH = 21.8
24.1 d
2JPH = 17.1
55HPhCH230.2 dd
2JPH = 24.2
3JPH = 23.4
81.0 d
2JPH = 30.6
607.54 dd, 1H, N–CH=
3JPH = 26.0, 3JHH = 15.2
7.28 m, 5H, Ph
5.69 d, 1H, P–CH=
2JPH = 16.2
5.48 dd, 1H, P1–CH=
2JPH = 26.0, 3JHH = 15.2
4.68 s, 2H, NCH2
33.0 dd
2JPH = 26.0
3JPH = 26.0
24.7 d
2JPH = 16.2
56HPh30.0 dd
2JPH = 22.5
3JPH = 22.5
83.1 d
2JPH = 31.8
617.70 dd, 1H, N–CH=
3JPH = 20.3, 3JHH = 14.8
7.22 m, 5H, Ph
5.70 d, 1H, P–CH=
2JPH = 18.0
5.47 dd, 1H, P1–CH=
2JPH = 25.0, 3JHH = 14.8
33.4 dd
2JPH = 25.0
3JPH = 20.3
26.0 d
2JPH = 18.0
57EtPhCH231.3 dd
3JPH = 28.6
CH2 (Et)
3JPH = 19.7
75.5 d
2JPH = 30.5
627.49 d, 1H, N–CH=
3JPH = 21.6
7.20 m, 5H, Ph
5.62 d, 1H, P–CH=
2JPH = 18.7
4.23 s, 2H, NCH2
3.28 d, 2H, CH2 3JPH = 28.6
1.70 3H, CH3

38.7 dd
3JPH = 28.6
CH2 (Et)
3JPH = 21.6
24.5 d
2JPH = 18.7
a The phosphorus signal of PCl 6   anion in the 31P NMR spectra of hexachlorophosphorates 53–57 appears in the region −(297–298) ppm.
Table 7. The 1H and 31P NMR chemical shifts (δ, ppm) and coupling constants (J, Hz) of hexachlorophosphorates 63–67 (CH3NO2) and 1,4-azaphosphorines 68–72 (CDCl3).
Table 7. The 1H and 31P NMR chemical shifts (δ, ppm) and coupling constants (J, Hz) of hexachlorophosphorates 63–67 (CH3NO2) and 1,4-azaphosphorines 68–72 (CDCl3).
No.R1H31P, POCl231P a, PCl 2 + /POCl 2JPP
63Pr9.65 dd, 1H, N–CH=
3JPH = 45.1, 3JPH = 22.9
8.25 dd, 1H, P–CH=, 4JPH = 10.4
3.64 t, NCH2, 3JHH = 6.6
1.49 m, 2H, CH2
0.98 t, 3H, CH3, 3JHH = 6.8
19.8 ddd
2JPP = 25.9
3JPH = 22.9
4JPH = 10.4
46.8 dd
3JPH = 45.1
2JPP = 25.9
25.9
64 Bu 9.90 dd, 1H, N–CH=
3JPH = 45.5, 3JPH = 22.3
8.28 d, 1H, P–CH=, 4JPH = 10.1
4.18 t, NCH2, 3JHH = 6.6
1.44 m, 1.72 m, 4H, (CH2)2
0.90 t, 3H, CH3, 3JHH = 6.7
20.3 ddd
2JPP = 25.5
3JPH = 22.3
4JPH = 10.1
46.2 dd
3JPH = 45.5
2JPP = 25.5
25.5
65 PhCH29.85 dd, 1H, N–CH=
3JPH = 45.0, 3JPH = 22.3
8.02 d, 1H, P–CH=, 4JPH = 9.9
7.27 m, 5H, Ph
4.47 s, 2H, NCH2
19.5 ddd
2JPP = 25.6
3JPH = 22.3
4JPH = 9.9
46.4 dd
3JPH = 45.0
2JPP = 25.6
25.6
66 Ph 8.77 dd, 1H, N–CH=
3JPH = 44.6, 3JPH = 24.3
7.21 d, 1H, P–CH=, 4JPH = 9.6
7.34 m, 5H, Ph
20.0 ddd
2JPP = 25.0
3JPH = 24.8
4JPH = 9.6
48.1 dd
3JPH = 44.6
2JPP = 25.0
25.0
67 s-Bu 8.85 dd, 1H, N–CH=
3JPH = 45.0, 3JPH = 26.0
7.40 d, 1H, P–CH=, 4JPH = 10.5
3.34 m, 1H, NCH
1.5 m, 2H, CH2
1.15 m, 6H, CH3
20.6 ddd
2JPP = 26.0
3JPH = 26.0
4JPH = 10.5
45.5 dd
3JPH = 45.0
2JPP = 26.0
26.0
68 Pr 8.12 dd, 1H, N–CH=
3JPH = 33.0, 3JPH = 23.9
6.68 d, 1H, P–CH=, 4JPH = 10.9
3.40 t, NCH2, 3JHH = 6.4
1.53 m, 2H, CH2
1.04 t, 3H, CH3, 3JHH = 6.7
22.6 ddd
2JPP = 25.9
3JPH = 23.9
4JPH = 10.9
10.9 dd
3JPH = 33.0
2JPP = 25.9
25.9
69 Bu 8.07 dd, 1H, N–CH=
3JPH = 32.9, 3JPH = 23.5
6.22 d, 1H, P–CH=, 4JPH = 10.3
4.05 t, NCH2, 3JHH = 6.5
1.40 m, 1.75 m, 4H, (CH2)2
0.92 t, 3H, CH3, 3JHH = 6.4
22.1 ddd
2JPP = 24.3
3JPH = 23.5
4JPH = 10.3
10.2 dd
3JPH = 32.9
2JPP = 24.3
24.3
70 PhCH28.07 dd, 1H, N–CH=
3JPH = 32.4, 3JPH = 22.9
7.48 m, 5H, Ph
6.24 d, 1H, P–CH=, 4JPH = 10.2
4.24 s, 2H, NCH2
25.5 ddd
2JPP = 24.4
3JPH = 22.9
4JPH = 10.2
13.5 dd
2JPP = 24.4
3JPH = 32.4
24.4
71 Ph 7.95 dd, 1H, N–CH=
3JPH = 32.6, 3JPH = 22.8
7.44 m, 5H, Ph
6.18 d, 1H, P–CH=, 4JPH = 9.9
25.8 ddd
2JPP = 24.1
3JPH = 22.8
4JPH = 9.9
13.9 dd
2JPP = 24.1
3JPH = 32.6
24.1
72 s-Bu 7.90 dd, 1H, N–CH=
3JPH = 32.2, 3JPH = 23.4
6.47 d, 1H, P–CH=, 4JPH = 10.8
3.44 m, 1H, NCH
1.5 m, 2H, CH2
1.13 m, 6H, CH3
24.0 ddd
2JPP = 24.6
3JPH = 23.4
4JPH = 10.8
13.1 dd
2JPP = 24.6
3JPH = 32.2
24.6
a The PCl 6 anion in organyltrichlorophosphonium hexachlorophosphorates 63–67 has a broad signal in the 31P NMR spectra in the region −(296–298) ppm.
Table 8. The 1H and 31P NMR chemical shifts (δ, ppm) and coupling constants (J, Hz) of hexachlorophosphorates 73 and 75 (CH3NO2) and 1,4-azaphosphorines 74 and 76 (CDCl3).
Table 8. The 1H and 31P NMR chemical shifts (δ, ppm) and coupling constants (J, Hz) of hexachlorophosphorates 73 and 75 (CH3NO2) and 1,4-azaphosphorines 74 and 76 (CDCl3).
No.1H/13C31P, POCl231P, PCl 2 + / POCl
737.34 d, 1H, CH=, 4JPH = 9.3
4.32 s, CH3
140.8 d, CCl=, 2JPC = 14.0
108.47 dd, P-C-P, 1JPC = 152.6 (POCl2)
1JPC = 101.3 (POCl)
106.20 dd, CH=, 1JPC = 124.4, 3JPC = 8.5
42.15 s, CH3
21.8 dd
2JPP = 21.5
4JPH = 9.3
49.2 d
2JPP = 21.5
74 6.24 d, 1H, CH=, 4JPH = 8.5
3.80 s, CH3
148.8 d, CCl=, 2JPC = 12.0
110.24 dd, P-C-P, 1JPC = 149.6 (POCl2)
1JPC = 102.5 (POCl)
104.34 ddd, CH=, 1JCH = 180.6, 1JPC = 118.2, 3JPC = 9.2
43.67 s, CH3
17.2 dd
2JPP = 21.2
4JPH = 8.5
9.0 d
2JPP = 21.2
75 11.37 br s, 1H, NH
5.98 d, 2H, CH=, 2JPH = 2.2
145.20 d, CCl=, 2JPC = 16.1
82.31 dd, CH=, 1JCH = 187.0, 1JPC = 110.8
- 49.9 t
2JPH = 2.2
76 10.60 br s, 1H, NH
5.72 d, 2H, CH=, 2JPH = 1.1
143.70 d, CCl=, 2JPC = 14.2
80.44 dd, CH=, 1JCH = 180.0, 1JPC = 103.6
- 25.8 t
2JPH = 1.1
Table 9. The 1H, 13C and 31P NMR chemical shifts (δ, ppm) and coupling constants (J, Hz) of compounds 88 and 89 (CDCl3).
Table 9. The 1H, 13C and 31P NMR chemical shifts (δ, ppm) and coupling constants (J, Hz) of compounds 88 and 89 (CDCl3).
No. Structure 1H13C31P
88Ijms 24 09646 i0397.07–7.20 m, 5H, Ph
6.44 dd, 1H, H-6
3JPH = 23.9, 3JHH = 8.1
5.85 dd, 1H, H-3,
3JPH = 11.4, 4JHH = 2.2
5.36 ddd, 1H, H-5,
3JHH = 8.1, 4JHH = 2.2, 4JPH = 2.2
147.5, 129.8, 120.0, 127.2 Ph
139.5 d, C-6, 2JPC = 2.0
138.3 s, C-Cl,
107, d, C-3, 1JPC = 149.6
105.6 d, C-5, 3JPC = 11.6
21.6 dd
3JPH = 23.9
2JPH = 11.4
89 Ijms 24 09646 i04013.39 s, 1H, NH
8.22 d, 1H, H-5, 3JHH = 8.2
7.89 dd, 1H, H-7,
3JHH = 8.2, 3JHH = 7.6
7.78 d, 1H, H-8, 3JHH = 8.2
7.59 dd, 1H, H-6,
3JHH = 8.2, 3JHH = 7.6
3.8 br s, 2H, OH
2.83 s, 3H, CH3
176.03 d, C-4, 2JPC = 7.7
157.24 d, C-2, 2JPC = 13.8
138.66 s, C-9,
133.91 s, C-7,
126.05 s, C-5,
124.87 s, C-6,
121.17 d, C-10, 3JPC = 11.1
118.86 s, C-8,
109.39 d, C-3, 1JPC = 161.0
20.42 s, CH3
11.6 s
Table 10. The 1H, 13C and 31P NMR chemical shifts (δ, ppm) and coupling constants (J, Hz) of compounds 9094.
Table 10. The 1H, 13C and 31P NMR chemical shifts (δ, ppm) and coupling constants (J, Hz) of compounds 9094.
No. R1 R 1H13C31P
90 H Me 7.80 dd, 1H, H-3
3JHH = 15.2, 4JPH = 2.9
7.21 d, 1H, H-4, 3JHH = 15.2
5.30 d, 1H, H-1, 2JPH = 56.4
2.10 s, 3H, CH3
143.50 d, C-2, 2JPC = 27.8
133.85 s, C-4,
128.67, d, C-3, 3JPC = 9.8
118.20 d, C-1, 1JPC = 116.3
21.28 s CH3
84.3 dd
2JPH = 56.4
4JPH = 2.9
91 Me Me 7.73 d, 1H, H-3, 4JPH = 2.7
5.61 d, 1H, H-1, 2JPH = 56.2
2.08 s, 3H, CH3
1.98 s, 3H, CH3
142.05 d, C-2, 2JPC = 26.7
139.75 s, C-4,
124.51, d, C-3, 3JPC = 7.2
116.50 d, C-1, 1JPC = 114.9
29.27 s CH3
21.66 s CH3
85.7 dd
2JPH = 56.2
4JPH = 2.7
92 H Ph 7.75 dd, 1H, H-3
3JHH = 14.8, 4JPH = 3.1
7.55–7.40 m, 5H, Ph
7.15 d, 1H, H-4, 3JHH = 14.8
5.87 d, 1H, H-1, 2JPH = 41.0
144.65 d, C-2, 2JPC = 26.0
140.90 s, i-Ph
136.86 s, C-4
128.98 s, m-Ph, p-Ph
127.69 s, o-Ph
127.48 d, C-3, 3JPC = 8.9
117.27 d, C-1, 1JPC = 115.5
89.1 dd
2JPH = 41.0
4JPH = 3.1
93 H p-MeO-Ph 7.80 dd, 1H, H-3
3JHH = 14.9, 4JPH = 3.0
7.60, d 2H, m-Ph, 3JHH = 8.7
7.19 d, 1H, H-4, 3JHH = 14.9
6.92, d 2H, o-Ph, 3JHH = 8.7
5.00 d, 1H, H-1, 2JPH = 42.7
4.23 s, OMe
159.27 s, C-OMe
144.25 d, C-2, 2JPC = 26.9
136.95 s, C-4
133.10 s, i-Ph
129.36 s, o-Ph
128.44 d, C-3, 3JPC = 8.2
119.54 d, C-1, 1JPC = 117.1
114.32 s, m-Ph
49.36 s, OMe
84.6 dd
2JPH = 42.7
4JPH = 3.0
94 H p-Cl-Ph 7.77 dd, 1H, H-3
3JHH = 14.5, 4JPH = 2.7
7.63, d 2H, m-Ph, 3JHH = 8.3
7.16 d, 1H, H-4. 3JHH = 14.5
6.76, d 2H, o-Ph, 3JHH = 8.3
5.63 d, 1H, H-1, 2JPH = 41.2
143.09 d, C-2, 2JPC = 27.6
139.00 s, i-Ph
136.29 s, C-4
134.60 s, C-Cl
129.24 s, m-Ph
129.18 s, o-Ph
128.51 d, C-3, 3JPC = 7.8
119.34 d, C-1, 1JPC = 116.9
88.7 dd
2JPH = 41.2
4JPH = 2.7
Table 11. 1H, 13C and 31P NMR chemical shifts (δ, ppm) and coupling constants (J, Hz) of 1,5,2-diazaphosphorines 103107.
Table 11. 1H, 13C and 31P NMR chemical shifts (δ, ppm) and coupling constants (J, Hz) of 1,5,2-diazaphosphorines 103107.
No.Structure1H13C31P
103Ijms 24 09646 i0415.93 d, H-3
2JPH 14.0
163.15 d, C-4, JPC 21.2
160.90 d, C-6, 2JPC 6.2
88.02 d, C-3, JPC 112.0
53.4 d
2JPH 14.0
104Ijms 24 09646 i042-160.15 d, C-4, JPC 37.8
157.73 d, C-6, 2JPC 4.9
97.27 d, C-3, JPC 132.8
44.7 s
105Ijms 24 09646 i0435.88 d, H-3
2JPH 42.2
3.45 d, CH3
3JPH 6.9
158.15 C-4, 2JPC 23.4
156.90 C-6, 2JPC 6.8
82.20 C-3, 1JPC 116.0
64.4 dq
2JPH 42.2
3JPH 6.9
–297.5 br s
PCl 6
106Ijms 24 09646 i0446.08 d, H-3
2JPH 12.8
3.38 d, CH3
3JPH 6.6
166.45 C-4, 2JPC 24.2
159.70 C-6, 2JPC 6.2
85.50 C-3, 1JPC 112.0
21.5 dq
2JPH 12.8
3JPH 6.6
107Ijms 24 09646 i0456.84 d, CHCl2
4JPH 1.0
3.10 d, CH3
3JPH 24.4
147.7 d, C=O, 2JPC 17.5
95.8 d, C-Cl, 2JPC 6.1
76.5 d, CHCl, 3JPC 7.3
28.9 d, CH3, 2JPC 2.5
−69.1 qd
3JPH 24.4
4JPH 1.0
Table 12. 1H and 31P NMR chemical shifts (δ, ppm) and coupling constants (J, Hz) of 1,5,2-diazaphosphorines 108114.
Table 12. 1H and 31P NMR chemical shifts (δ, ppm) and coupling constants (J, Hz) of 1,5,2-diazaphosphorines 108114.
No.Structure1H31P
108Ijms 24 09646 i0466.83 d, H-3, 2JPH = 10.6
3.90 s, 5-CH3
3.78 d, 1-CH3
3JPH = 9.9
58.3 dq
2JPH = 10.6, 3JPH = 9.9
−294.7 br s, PCl 6
109Ijms 24 09646 i0473.68 s, 5-CH3
3.37 d, 1-CH3
3JPH = 9.6
1.85 d, 3-CH3
3JPH = 6.7
61.2 qq
3JPH = 9.6, 3JPH = 6.7
−297.6 br s, PCl 6
110Ijms 24 09646 i0483.57 s, 5-CH3
3.30 d, 1-CH3
3JPH = 12.2
2.85 dd, 3-CH2
3JPH = 10.0, 3JHH = 8.2
1.75 d, 3-CH3
3JHH = 8.2
60.8 qt
3JPH = 12.2, 3JPH = 10.0
−297.9 br s, PCl 6
111Ijms 24 09646 i0495.71 d, H-3
2JPH = 6.8
3.50 s, 5-CH3
3.19 d, 1-CH3
3JPH = 7.3
17.1 qd
3JPH = 7.3, 2JPH = 6.8
112Ijms 24 09646 i0503.35 s, 5-CH3
3.04 d, 1-CH3
3JPH = 7.6
1.96 d, 3-CH3
3JPH = 15.1
18.8 qq
3JPH = 15.1, 3JPH = 7.6
113Ijms 24 09646 i0513.48 s, 5-CH3
3.32 d, 1-CH3
3JPH = 10.0
2.85 dd, 3-CH2
3JPH = 9.0, 3JHH = 7.8
1.75 d, 3-CH3
3JHH = 7.8
19.2 qt
3JPH = 10.0, 3JPH = 9.0
114Ijms 24 09646 i0525.70 d, H-3
2JPH = 61.4
3.42 s, CH3
3.14 d, CH3
3JPH = 14.2
100.5 dq
2JPH = 61.4, 3JPH = 14.2
Table 13. 1H, 13C and 31P NMR chemical shifts (δ, ppm) and coupling constants (J, Hz) of 1,5,2-diazaphosphorines 117 and 118.
Table 13. 1H, 13C and 31P NMR chemical shifts (δ, ppm) and coupling constants (J, Hz) of 1,5,2-diazaphosphorines 117 and 118.
No.Structure1H31P13C
117Ijms 24 09646 i0537.14 d, = CH
2JPH = 21.1
6.38 br s, NH
4.10 t, 5-CH2
3JHH = 8.4
3.52 t, 4-CH2
3JHH = 8.4
27.9 d
2JPH = 21.1
155.98 s, C = O
143.89 d, C-Cl, 2JPC = 6.5
103.28 d, = CH
1JPC = 176.4
45.75 s, 5-CH2
36.38 s, 4-CH2
118Ijms 24 09646 i0547.64 br s, NH
3.75 t, 5-CH2
3JHH = 8.2
3.57 d, CH2
2JPH = 21.8
3.28 t, 4-CH2
3JHH = 8.2
16.0 t
2JPH = 21.8
165.70 d, C = O
2JPC = 6.9
155.80 s, 2-C = O
42.21 s, 5-CH2
35.33 s, 4-CH2
35.28 d CH2
1JPC = 126.3
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

Larina, L. C- and N-Phosphorylated Enamines—An Avenue to Heterocycles: NMR Spectroscopy. Int. J. Mol. Sci. 2023, 24, 9646. https://doi.org/10.3390/ijms24119646

AMA Style

Larina L. C- and N-Phosphorylated Enamines—An Avenue to Heterocycles: NMR Spectroscopy. International Journal of Molecular Sciences. 2023; 24(11):9646. https://doi.org/10.3390/ijms24119646

Chicago/Turabian Style

Larina, Lyudmila. 2023. "C- and N-Phosphorylated Enamines—An Avenue to Heterocycles: NMR Spectroscopy" International Journal of Molecular Sciences 24, no. 11: 9646. https://doi.org/10.3390/ijms24119646

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop