Next Article in Journal
A Novel Drug Delivery System: Hyodeoxycholic Acid-Modified Metformin Liposomes for Type 2 Diabetes Treatment
Next Article in Special Issue
Bathochromic Shift of Fluorescence Peak in Dipyrrolo[1,2-a:2′,1′-c]quinoxaline by Introducing Each of Electron-Donating and Electron-Withdrawing Substituent
Previous Article in Journal
Food Plant Secondary Metabolites Antiviral Activity and Their Possible Roles in SARS-CoV-2 Treatment: An Overview
Previous Article in Special Issue
Mechanochemistry Frees Thiourea Dioxide (TDO) from the ‘Veils’ of Solvent, Exposing All Its Reactivity
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Synthesis of Phosphorus(V)-Substituted Six-Membered N-Heterocycles: Recent Progress and Challenges

N. D. Zelinsky Institute of Organic Chemistry, Russian Academy of Sciences, 47 Leninsky Prosp., 119991 Moscow, Russia
*
Author to whom correspondence should be addressed.
Molecules 2023, 28(6), 2472; https://doi.org/10.3390/molecules28062472
Submission received: 15 February 2023 / Revised: 2 March 2023 / Accepted: 2 March 2023 / Published: 8 March 2023

Abstract

:
Heterocycles functionalized with pentavalent phosphorus are of great importance since they include a great variety of biologically active compounds and pharmaceuticals, advanced materials, and valuable reactive intermediates for organic synthesis. Significant progress in synthesis of P(O)R2-substituted six-membered heterocycles has been made in the past decade. This review covers the synthetic strategies towards aromatic monocyclic six-membered N-heterocycles, such as pyridines, pyridazines, pyrimidines, and pyrazines bearing phosphonates and phosphine oxides, which were reported from 2012 to 2022.

1. Introduction

Pentavalent phosphorus-substituted heterocycles are organophosphorus compounds of great importance. In materials chemistry, they are used in design of dyes and polymers with outstanding characteristics [1]. They are of interest as Lewis acids and ligands in metal-catalyzed organic reactions [2,3,4,5]. Moreover, phosphorus(V)-substituted heterocycles exhibit a wide range of biological activities. In particular, pyridines were used to design anticancer [6,7,8], antidiabetic [9], and lusitropic agents [10], antioxidants [11], human glucokinase activators [12], NMDA receptor antagonists [13], and metalloprotein ligands [14]. Phosphorus-modified pyridazines are promising candidates as antimicrobial agents [15] and foliar herbicides [16]. Phosphorus-containing pyrimidines are used as inhibitors of dihydroorotate dehydrogenase and hepatitis C virus polymerase [17,18,19]. Pyrazines are of interest as modulators of the human glucagon-like peptide-1 receptor [20] and cyclin-dependent kinases [21].
A variety of methodologies for synthesis of pentavalent phosphorus-substituted heterocycles and their derivatives have been reported in past years. Meanwhile, development of novel methodologies is in continuous demand. Of particular interest are methods that provide a broad substrate scope and provide products with high atom economy, molecular complexity, and great efficiency under mild conditions. Among them, several eco-friendly approaches have been developed by us [22,23]. In general, the known methods for synthesis of phosphorus(V)-substituted aromatic heterocycle are classified into two synthetic strategies: (1) construction of a heterocyclic core using phosphorus-containing reagents and (2) functionalization of a pre-synthesized heterocyclic core with phosphorus reagents.
Several excellent reviews dealing with synthesis of phosphorus-containing heterocyclic compounds were published in the literature in recent years. A number of reviews highlighted phosphorylation as the most common approach [24,25] and specifically radical phosphorylation [26] and metal-catalyzed/promoted C–H bond phosphorylation [27]. Another array of reviews considered use of a specific kind of transformations, including photoredox-catalyzed reactions [28], electrocatalytic [29], and radical [30] C–P bond formation, multicomponent reactions [31], and functionalization of phosphorus-centered radicals [32]. Some reviews are focused on the type of P-functional groups [33] or type of heterocycles [34], e.g., phosphorylindoles [35,36], pyrazoles [37,38], or pyrrolidines [39]. However, systematic reviews of the published data on different methods for synthesis of most monocyclic phosphorus(V)-substituted six-membered N-heterocycles are lacking.
The main focus of this review is monocyclic six-membered aromatic P(O)R2-substituted N-heterocycles, including pyridines, pyridazines, pyrimidines, and pyrazines. The review is an attempt to systematically consider and classify methods for synthesis of corresponding heterocyclic phosphonates and phosphine oxides. Phosphonic acids are excluded from consideration since most of them are available by saponification of phosphonates [40,41,42,43,44,45]. Synthesis of six-membered N-heterocyclic phosphinamides, phosphonamide, and phosphonamidate has also not been described in recent years. The review is organized in a classical way and includes classification according to type of N-heterocycles using the following criteria: (a) the number of nitrogen atoms in heterocycles, (b) the nature of synthetic precursors, and (c) the type of chemical transformations involving these heterocycles.
The review addresses the scientific advances made over the past decade, in the period from 2012 up to the end of 2022. However, some earlier pioneering studies are cited where necessary. In many cases, the reaction mechanisms are briefly discussed in order to more completely describe the synthetic approaches to phosphorus-substituted N-heterocycles. In order to avoid overloading of the text with names of semantically similar processes, the term “phosphorylation” in this review deals with introduction of any P(O)R2 substituent, not only a phosphoryl moiety, into the heterocyclic core. To enhance the readability of the schemes, the following color coding of the reagents is used: the source of the nitrogen atom is shown in blue, the source of the phosphorus atom in red, and the extra components in green. In the case of intramolecular cyclizations of reagents containing both nitrogen and phosphorus, the nitrogen atom is marked in blue and the phosphorus moiety in red.

2. Pyridines

Phosphorus(V)-substituted pyridine was first synthesized by Plazek’s research group [46] in 1936 by a reaction of 2-dimethylaminopyridine with phosphorus trichloride under oxidative conditions. Later, related compounds were obtained by reactions of metalated pyridines with phosphorus-halogen compounds [47,48,49], of pyridinediazonium tetrafluoroborate with phosphorus trichloride [50], of N-alkoxypyridines with sodium diethyl phosphite and phosphines [51,52,53,54,55], and of N-pyridylpyridines with phosphonic acid [56], by the Michaelis–Arbuzov reaction [57,58], Pd(II)-catalyzed phosphorylation of halopyridines [45,59,60], cyclization of phosphorus-containing 3-azatrienes [61], and the Diels–Alder reaction involving 3-phosphoryl-1-azadienes [62]. This section of the review considers recent examples of synthesis of POR2-substituted pyridines, including the intramolecular cyclizations of Michael adducts, insertion of carbenoids, formal [2+2+2]-cycloaddition, and phosphorylation reactions (Figure 1).

2.1. Cyclizations Based on the Michael Reaction

In recent years, synthesis of pyridines via the Michael reaction using phosphoryl-substituted Michael acceptors and donors has gained significant attention. Allais et al. [63,64,65] described three-component condensation of 1,3-dicarbonyl compounds 3 with P(O)Et2-bearing vinyl ketones 2 and ammonium acetate (1) in the presence of oxygen, leading to pyridine-2-phosphonates 4 (Scheme 1). The authors suggested that the reaction occurs through successive addition of compound 3 to Michael acceptors 2, providing ketone 5 followed by enamination of the latter with ammonia (1′) to form intermediate 6, which undergoes intramolecular cyclization into dihydropyridine 7. Complete oxidation of intermediate 7 to pyridine 4 was achieved using oxygen in the presence of activated carbon. The reaction was general with respect to β-oxo esters and β-oxoamides, providing products in 49–80% yields.
Hanashalshahaby and Unaleroglue [66] showed that pyridine-3-phosphonates 10 can be obtained by three-component oxidative coupling of diethyl (2-oxobutyl)phosphonate (8) with Mannich bases 9 and ammonium acetate (1) in the presence of catalyst K-10 (Scheme 2). The product yields are reasonable both for aryl- and alkyl-substituted Mannich bases. The authors hypothesized that β-keto phosphonate 8 reacts with ammonia 1′ generated in situ from ammonium acetate (1) to form enamine 11, which is accompanied by thermal decomposition of Mannich base 9, giving α,β-unsaturated carbonyl compound 9′. These two intermediates are subjected to the Michael addition to form ketoamine 12, which undergoes intramolecular cyclization to dihydropyridine 13, followed by oxidation of the latter with atmospheric oxygen to provide final product 10.
Further, Abdou et al. [67] found that diethyl (2-amino-2-thioxoethyl)phosphonate acts as an efficient Michael donor in the addition reaction with β-(dimethylamino)vinyl ketone, thus leading to 2-thioxopyridin-3-ylphosphonate. Liao et al. [68] described one example of the aza-Michael reaction between diethyl (3-phenyl-3-oxopropyn-1-yl)phosphonate with methyl 3-aminocrotonate, providing phosphonate-ester-containing pyridine moiety under mild conditions.

2.2. Carbenoid-Mediated Reactions

Recently, several studies by Park and co-workers demonstrated the prospects of using metal carbenoids in synthesis of phosphoryl-substituted pyridines. They described [69] Rh(II)-catalyzed intramolecular cyclization of δ-diazo oximes to pyridines and showed that this is a facile method for synthesis of pyridine-2-phosphonate 15 (Scheme 3). Rh2(CF3CONH)4 was used as the catalyst of choice. The authors proposed a mechanism that involves reaction of the diazo group of oxime 14 with Rh(II), providing rhodium carbenoid 16, followed by insertion of the latter into the N–O bond to form dihydropyridinone 17, which undergoes aromatization to final pyridine 15 via elimination of methanol.
In another work by Park and co-workers [70], synthesis of pyridine-2-phosphonate 20 was accomplished using phosphorylated vinyl carbenoid 21 generated in situ from diazophosphonate 19 and dirhodium(II) catalyst Rh2(esp)2 (Scheme 4). Addition of compound 21 to 2H-azirine 18 affords intermediate 22. The latter undergoes three-membered ring opening accompanied by elimination of the Rh(II) catalyst to provide 3-azahexatriene 23, followed by cyclization to yield pyridine 20.

2.3. Formal [2+2+2]-Cycloaddition

An approach towards phosphoryl-substituted pyridines, which is probably one of the most versatile, is based on the formal [2+2+2]-cycloaddition. Tanaka’s research group [71,72] described synthesis of annulated pyridine-2-phosphonates 26 based on rhodium(I)/biaryl-bisphosphine-complex-catalyzed cycloaddition of 1,6- and 1,7-diynes 24 with diethyl phosphorocyanidate (25) (Scheme 5). The reaction has a broad scope with respect to diynes since quaternary carbon-, methylene-, nitrogen-, and oxygen-linked internal 1,6-diynes and terminal biaryl-linked 1,7-diynes can be involved in the heterocyclization. Steric and electronic variations in diynes had minimal impact on the efficacy of the reaction, but, in some cases, using unsymmetrically substituted diynes, the reactions afforded mixtures of regioisomers. The authors proposed rhodium cyclopentadiene 27 or rhodium azacyclopentadiene 28 as two possible key intermediates in the reaction (Scheme 5). Ring expansion of both cyclopentadienes can provide seven-membered intermediate 29, and reductive elimination of Rh(I)+ from the latter accomplishes formation of the final product 26. Based on the outcome of the enantioselective version of the reaction, the authors were inclined to believe that the main reaction pathway involves intermediate 28.

2.4. Phosphorylation of Pyridines

Phosphorylation is definitely the most general route to phosphorus-substituted pyridines. Due to a wide range of available phosphorylating agents and the possibility of performing the reaction in a nucleophilic, electrophilic, or radical manner, phosphorylation has attracted great attention. Such methods as the Arbuzov reaction, Hirao coupling, palladium-catalyzed cross-coupling of diethyl phosphonate with halogen-substituted heterocycles, and C(sp2)H-phosphorylation of heterocycles with diethyl phosphonates promoted by one-electron oxidants were extensively developed in past decades. Their specific applications in recent years are discussed in more detail later in the text.

2.4.1. Radical Phosphorylation of Pyridines

P-centered radicals can easily be generated through hydrogen atom transfer or single-electron transfer using peroxides, metal salts, and photocatalysts. Thus, radical phosphorylation has become an efficient strategy for synthesis of structurally diverse phosphorus(V)-substituted pyridines. Due to environmental friendliness and potential industrial application, photocatalytic radical phosphorylation of the pyridine ring received extra attention. In 2018, Yuan et al. [73] reported synthesis of 2- and 3-phosphine oxide-substituted pyridines 31 and 33 (Scheme 6, lines a,b) from 2- and 3-halopyridines 30 and 32 by photocatalytic functionalization with secondary phosphine oxides in the presence of tBuOK. The reaction occurred under mild conditions using irradiation with a blue-light-emitting diode. The scope of this transformation is somewhat limited to pyridines with electron-donating substituents and their analogs with an extended π-system. The plausible reaction mechanism involves formation of a complex of halopyridine with potassium tert-butoxide 34, absorption of a light quantum, transition to an excited state 35, and electron transfer from tert-butoxide to a pyridine ring to form the tert-butoxy radical and halopyridine radical anion 37. Elimination of halide from radical anion 36 affords aryl radical 37. Simultaneously, the tert-butoxy radical causes the proton abstraction from secondary phosphine to form a phosphorus-containing radical. Recombination of the latter radical with an aryl radical affords the final product. Recently, the approach proposed by Yuan et al. was expanded [74,75], including electron-primed photoredox [76] and visible light-induced nickel-catalyzed photoredox [77] conditions for radical generation.
In 2019, the photocatalytic radical reaction of pyridylazo sulfones 38 with triphenyl phosphite giving pyridine-3-phosphonates 39 was described by Qiu et al. [78] (Scheme 7). The reaction occurred in the presence of water. A variety of substituted pyridines can be efficiently employed in this reaction. The proposed mechanism for this transformation involves excitation of arylazo sulfone 38 under visible light to provide radical 40, its decomposition into sulfonyl and aryl radicals, along with extrusion of a nitrogen molecule. Then, aryl radical 41 reacts with triphenyl phosphate to form phosphorus-centered radical 42, followed by its oxidation with the sulfonyl radical and elimination of phenol to provide the final product 39.
Independently, in 2019, Kim et al. [79] described site-selective synthesis of pyridines 44 bearing phosphine oxide moieties at C-4 using radical coupling of N-ethoxypyridinium salts 43 with secondary phosphine oxides under photocatalytic conditions (Scheme 8). Further, 3-Diphenylphosphoryl-6-methoxy-1-methyl-2(1H)-quinolinone under blue-light-emitting diode illumination was used as a photocatalyst and potassium persulfate as an oxidant. Examination of the reaction scope revealed that a variety of electron-withdrawing and electron-donating substituted pyridines, as well various aryl-substituted phosphine oxides, were tolerated. The plausible mechanism of this reaction involves single-electron transfer from the photocatalyst to N-ethoxypyridinium 43, giving radical 45, which undergoes decomposition accompanied by elimination of pyridine. The remaining ethoxy radical abstracts a proton from phosphine, followed by addition of the resulting phosphinyl radical 46 to N-ethoxypyridinium 43. The subsequent deprotonation of intermediate 47 and elimination of the new ethoxy radical from intermediate 48 afford target product 44. The origin of reaction chemoselectivity was revealed by DFT calculations, showing that phosphinoyl radicals are too large for providing an electrostatic attraction between the its oxo functionality and the pyridine nitrogen crucial for ortho functionalization.
Apart from the approaches based on photocatalytic radical phosphorylation, several examples using metal salts for P-radical generation were described. Huang and co-workers [80] developed a CH-phosphorylation method for synthesis of pyridinyl-2-phosphonates 50 based on Ag(I)-catalyzed reaction of pyridines 49 with dialkyl phosphonates using potassium persulfate as an oxidant (Scheme 9). An interesting feature of this reaction is that it involves subsequent treatment of the reaction mixture with sodium thiosulfate, which makes it possible to significantly increase the product yield due to a reduction of pyridine N-oxide formed as a by-product. In the authors’ opinion, this transformation proceeds through a radical pathway and begins with oxidation of Ag(I) to Ag(II) with persulfate, followed by oxidation of dialkyl phosphite with Ag(II) to form radical cation 5. Addition of the latter at 2 position of the pyridine ring provides intermediate 52. Subsequent abstraction of two protons accompanied by oxidation of radical 52 affords the final product. Recently, Kittikool et al. [81] expanded the scope of this phosphorylation approach to 2-pyridones using Mn(OAc)2 as the catalyst.
Noteworthy also is a one-pot three-step protocol for synthesis of diphenyl(pyridin-2-yl)phosphine oxide based on the KOH-promoted oxidative radical phosphorylation of 2-bromopyridine with diphenylphosphine developed by Chen et al. [82].

2.4.2. Nucleophilic Phosphorylation of Pyridines

Studies by Trofimov’s research group [83,84,85] have contributed to recent progress in nucleophilic phosphorylation of pyridines at the 4 position. They accomplished synthesis of 4-phosphine oxide-substituted pyridines 55 by coupling of pyridines 53 with secondary phosphine oxides using diphenyl ethynyl ketone (54) as the oxidant (Scheme 10). A variety of substituted aryl and alkyl phosphine oxides can efficiently be employed in this reaction. The proposed mechanism for this transformation involves aza-Michael reaction of pyridine 53 with acetylene 54, providing intermediate 56, and deprotonation of phosphine, resulting in formation of the anion. Subsequent addition of a phosphine anion at the 4 position of the activated pyridine ring affords dihydropyridine 57. The latter undergoes isomerization to intermediate 58, followed by elimination of alkene 59 to form the final product. In this approach, 3-phenyl-2-propynenitrile can also be applied as an oxidant [84], The reaction can be stopped at the of step of 1,4-dihydropyridines 57 using terminal acylacetylenes [85]. It is interesting that 1,4-dihydropyridines are shown to be produced through 2-4 migration of the POR2 group in 1,2-dihydro adducts during vinylation/phosphorylation of pyridines [85]. DFT calculations supported the hypotheses that 1,4-dihydropyridines are the thermodynamic products while their 1,2-regioisomers are the kinetic ones [85].
Direct nucleophilic phosphorylation at the 2 position of the pyridine ring was described in 2012 by Oka et al. [86] A series of 2-phosphinate-substituted pyridines 61 were prepared from N-methoxypyridinium tosylates 60 by reaction with secondary phosphinates (Scheme 11). Presumably, the reaction proceeds via the SNAr mechanism. Further, 1,8-Diazabicyclo[5.4.0]undec-7-ene (DBU) in acetonitrile at low temperature was found to be the optimal base for chemoselective transformation. The procedure was general for thiophosphinate, providing 2-pyridyl thiophosphinate. The 2-pyridyl phosphinate derivatives produced by this method are P-chiral, and the reaction can be accomplished with high diastereoselectivity using chiral phosphines.
In 2014, Wang et al. [87] showed that synthesis of pyridine-2-phosphonates can be accomplished directly from pyridine N-oxides and dimethyl phosphate over long-term heating in toluene. Independently, Lee et al. [88] developed an approach to synthesize diethyl pyridine-2-phosphonates 63 by reaction of pyridine N-oxides 62 with triethyl phosphite in the presence of ethyl chloroformate (Scheme 12). This reaction presumably starts with activation of pyridine N-oxide 62 by chloroformate, followed by addition of triethyl phosphate to the resulting pyridinium salt 64 to form intermediate 65. The nucleophilic attack of the chloride anion at the ethyl moiety followed by elimination of ethyl carbonate from 1,2-dihydropyridine 66 affords the final product. This approach is general for preparation of quinolin-2-ylphosphonates. An apparent limitation of this approach is formation of phosphorylated isomeric mixtures in the case of 3-bromopyridine N-oxide.
Recently, Tsantrizos and co-workers [89] proposed a route to P-chiral 2-phosphine oxide-substituted pyridines (Scheme 13). Using the reaction of (R)-N-(1-(5-chloro-2-hydroxyphenyl)ethyl)-4-methylbenzenesulfonamide (67) with Grignard to generate chiral phosphorylating agent 68, they accomplished enantioselective synthesis of products 70 from 2-fluoropyridines 69.

2.4.3. Transition-Metal-Catalyzed Phosphorylation of Pyridines

Transition metals are widely used to increase efficiency of conventional non-catalyzed phosphorylation of heterocycles [27]. In the past decade, significant progress in the field of phosphorylation of pyridines was made due to application of palladium and nickel catalysis, which enabled expansion of the range of substrates active in phosphorylation reactions to involve halogen-, hydroxyl-, boronic acid-, triflate-, nonaflate-, ester-substituted, trimethylammonium pyridines, etc. It was also shown that transition-metal-catalyzed phosphorylation with phoshine oxides, phosphates, and phosphites could be efficiently used for construction of functionalized pyridines, including chiral structures. Meanwhile, transition-metal-catalyzed C–H phosphorylation of pyridines still remains an unsolved challenge.
Halopyridines (Hal = Cl, Br, and I) are efficient precursors for synthesis of 2,3,4-POR2-substituted pyridines under palladium catalysis conditions. In 2017, Han et al. [90] reported stereoselective palladium-catalyzed cross-coupling of 2-bromopyridine with chiral tert-butyl-containing phosphine oxides (Scheme 14). Use of Pd2(dba)3 with the dppp ligand enabled synthesis of tertiary pyridine-containing phosphine oxides 71 with excellent selectivity. In 2016, Dziuganowska et al. [13] used the same catalytic system to prepare a series of phosphonopyridinecarboxylic acid esters from appropriate bromides.
Mykhailiuk and co-workers [91] expanded the scope of this reaction using the Pd2(dba)3/Xantphos catalytic system in the reaction of bromo(iodo)pyridines with dimethylphosphine oxide (Scheme 15). The scope of this reaction was found to be quite general and enables facile preparation of 2-, 3-, and 4-dimethylphosphine-oxide-substituted pyridines 72 in up to good yields. A variety of electron-withdrawing and electron-donating aryl-substituted pyridines successfully participated in the reaction.
Borisova and co-workers [92] accomplished synthesis of bis(phosphoryl)pyridines and 2,2′-bipyridines using palladium acetate/dppf-catalyzed cross-coupling of chloropyridines with secondary phosphine oxides. Ligand-free microwave-assisted Pd(OAc)2-catalyzed phosphorylation of bromopyridines with diphenylphosphine oxide or diethyl phosphite was described by Henyecz et al. [93]. Catalyst Pd(OAc)2 in the presence of sodium iodide as the promoter was also shown to be an effective catalyst for ligand-free coupling of diphenylphosphine oxide with pyridinium nonaflate [94]. An example of Pd(PPh3)4-catalyzed coupling of 2-bromo-6-pyrrolylpyridine with diethyl phosphite was recently reported by Ti et al. [95].
The palladium-catalyzed Hirao coupling of bromopyridines with triethyl phosphite enables preparation of pyridinephosphonates. Adam et al. [96] used the palladium-catalyzed cross-coupling of 3-bromopyridines 73 with triethyl phosphite to synthesize 2-aminopyridine-3-phosphonates 74 (Scheme 16). It is worth noting that the reaction was performed in the absence of the solvent during short-term heating to 160–180 °C.
Apart from halopyridines, pyridine boronic acids and hydroxypyridines were involved in phosphorylation under palladium catalysis. Zou, Wu, and co-workers [97] constructed a C–P bond by means of cross-coupling of pyridine boronic acids 75 with dialkyl phosphonates under PdCl2 catalysis conditions (Scheme 17). The reaction requires the presence of Ag2O as the oxidant. The method can be used to prepare structurally diverse pyridine-3- and pyridine-4-phosphonates 76.
Recently, Ding and co-workers [98] described the general palladium-catalyzed one-pot procedure for synthesis of phosphonates, phosphinates, and phosphine oxides from phenols mediated by sulfuryl fluoride. The reaction was efficient for functionalization of 2- and 3-hydroxypyridines 77 with diethyl phosphonate (Scheme 18). According to the proposed mechanism, the fluorosulfates generated in situ are key intermediates in synthesis of pyridines 78.
Before these works, in 2012, direct replacement of an OH group of hydroxypyridines by a diphenylphosphoryl moiety was achieved by Zhao et al. [99] under Ni(II) catalysis. A C–O bond was activated using bromotripyrrolidinophosphonium hexafluorophosphate. The activated complex of hydroxypyridine 77 with this salt reacts with diphenylphosphine oxide in the presence of dichloro[1,3-bis(diphenylphosphino)propane]nickel salts as the catalyst to produce pyridines 79 (Scheme 19).
Later, the substrate scope of Ni-catalyzed phosphorylation of pyridines with phosphine oxides was expanded to pyridine carboxylic acid phenyl esters [100] and trimethylammonium triflate of pyridine [101]. Cross-coupling of pyridine tosylates [102] and aryltrimethylammonium tetrafluoroborate [103] with phosphites was also implemented.
A relatively general method was proposed by Yamaguchi’s research group [100]. They performed cross-coupling of pyridine carboxylic acid phenyl esters 80 with secondary phosphine oxides accompanied by decarbonylation (Scheme 20). The reaction is catalyzed by nickel acetate and requires high temperatures (150–170 °C). This approach is applicable to synthesis of 2- and 3-phosphine-oxide-substituted pyridines 81, the position of the POR2 substituent in the product being defined by the position of the carbonyl group in the starting compound. The authors suggested that this transformation initially proceeds through oxidative insertion of Ni(0) into the C–O bond, providing intermediate 82, and exchange of the phenoxide ligand by phosphine provides 83. Subsequent decarbonylation of intermediate 83 and reductive elimination from intermediate 84 afford the final product.
Of interest as well is Ni(II)-catalyzed reaction of tosylates 85 with H-phosphonate diesters proposed by Chun-Jing Li [102] as an approach to 2- and 3-phosphorylpyridines 86; NiCl2(cod)2 was an optimal catalyst for this transformation (Scheme 21).

3. Pyridazines

Phosphorus(V)-substituted pyridazines are poorly described heterocycles. Therefore, all known methods for synthesis of these compounds, including isomerization of (diazomethyl)cyclopropenes and bis-azirine, intramolecular cyclization of γ,δ-unsaturated α-diazo-β-ketone and hydrazone, as well as phosphorylation of the pyridazine ring (Figure 2), are considered below.

3.1. Isomerization Reactions

Regitz and co-workers [104,105,106,107] reported the only method for synthesis of POR2-containing pyridazines, the general character of which was demonstrated in relation to a representative series of compounds. The authors showed that (diazomethyl)cyclopropenes 89, derived from cyclopropenylium salts 87 and phosphorus-containing diazomethanes 88, undergo intramolecular isomerization to 3-P(O)R2-substituted pyridazines 90 (Scheme 22). When using secondary amino-substituted cyclopropenylium salts as substrates, the reaction can be performed in a one-pot fashion [107]. The method tolerates substrates containing phosphine oxide moieties and can be used to synthesize POR2-containg 4,5-diamino-substituted pyridazines in 16–68% yields. The authors suggested 3,4-diazabenzvalene I, product of the intramolecular 1,3-dipolar cycloaddition, or bicyclic zwitterions II, formed via intramolecular 1,5-dipolar cyclization, as two key intermediates in the isomerization reaction (Scheme 22). The results of the reactions with sterically hindered cyclopropenes inclined the authors to believe that the main reaction pathway involves intermediate II.
A single example of thermal rearrangement of phosphorus-substituted bis-azirine 91 into 4,5-bis(phosphine oxide)pyridazine 92 was reported by Banert et al. [108] (Scheme 23). Formation of pyridazine may be explained by multistep diradical pathways, supported by DFT calculations. However, this is one of several reaction pathways, and the yield of the target product was as low as 3%.

3.2. Intramolecular Cyclizations

Doutheau and co-workers [109] described one example of cyclization of γ,δ-unsaturated α-diazo-β-keto phosphonate 95 to pyridazine-2-phosphonate 94 (Scheme 24). This compound is generated in situ via diazo transfer from γ,δ-unsaturated β-keto phosphonate 93 in the presence of tosyl azide and immediately undergoes intramolecular electrophilic cyclization involving a diazo moiety and double bond to form intermediate 96, followed by rearrangement of the latter into final product.
Touil and Zantour [110] synthesized pyridazine-3-phosphonate 99 by reaction of phosphorylated 1,4-dicarbonyl compound 98 with hydrazine hydrate (97) under oxidative conditions (Scheme 25). The reaction presumably proceeds through intermediate hydrazones 100, which undergo spontaneous intramolecular cyclization to 4,5-dihydropyridazine 101. Oxidative aromatization of the latter with oxygen affords a product in high yield.

3.3. Phosphorylation of Pyridazines

There are only two known examples of phosphorylation of pyridazines. Thus, the Michaelis–Arbuzov reaction of 3,6-dichloropyridazine (102) with methoxydiphenylphosphine was proposed as an approach to synthesis of 3,6-bis(diphenylphosphine oxide)pyridazine (103) in the patent by Mrowca [111] (Scheme 26, line a). An example of synthesis of 2-diphenyl(pyridazin-3-yl)phosphine oxide 105 from N-ethoxypyridazinium 104 and diphenylphosphine oxide via photocatalysis was presented by Kim et al. [79] (Scheme 26, line b). This process follows a radical mechanism, which is considered in more detail in Section 2.4.1 (Scheme 8). Therefore, the known phosphorylation reactions of pyridazines are limited to functionalization of the 3 and 6 positions of the heterocyclic core.

4. Pyrimidines

Phosphorus(V)-substituted pyrimidines were first synthesized by Kosolapoff and Roy [112] in 1961 by reaction of chloropyrimidines with NaP(O)(OEt)2. This method was also applied in many other studies [111,113]. Further, related compounds were prepared by reaction of bis-electrophiles with amidines [114,115,116,117,118], reaction of β-keto vinylphosphonates with amidines [119], reaction of metalated pyrimidines with P-Cl compounds [120], and palladium-catalyzed phosphorylation [45]. This section deals with recent examples of synthesis of phosphorus-containing pyrimidines, including the reactions of α,β-unsaturated β-phosphoryl carbonyl compounds with guanidine and phosphorylation of the pyrimidine core (Figure 3).

4.1. Cyclizations Using Guanidine and Amidines

In early works, synthesis of pyrimidines was accomplished by heterocyclization of guanidine and amidines with phosphorus-containing bis-electrophiles, such as 3-(methylthio)acrylonitriles [114,118], 3-(dimethylamino)propenones [115,116], and acrylonitrile acetal [117]. Recently, Zhu et al. [121] reported one example of the reaction of phosphine oxide-substituted ketene dithioacetal 107 with guanidine (106) providing 5-phosphorus(V)-containing pyrimidine 108 (Scheme 27, line a). Liao et al. [68] described synthesis of pyrimidinyl-4-phosphonate 111 by a reaction between diethyl (3-phenyl-3-oxopropyn-1-yl)phosphonate 110 and benzamidine (109) (Scheme 27, line b).
In addition, it may be noted that, recently, three-component acidic condensation of β-keto phosphonates with aldehydes and urea (Biginelli reaction) was shown to be efficient for synthesis of 1,4-dihydro-2-oxopyrimidine-5-phosphonates [122,123,124]. In this regard, it can be assumed that, in the future, a similar approach using amidines can also be implemented to obtain pentavalent phosphorus-substituted pyrimidines.

4.2. Phosphorylation of Pyrimidines

Due to the significant limitations of synthetic approaches to structurally diverse phosphorus-substituted pyrimidines from phosphorus-containing precursors, phosphorylation of pyrimidines has gained great attention in recent years. Pioneering studies of nucleophilic and electrophilic phosphorylation of pyrimidines [111,112,113,120] were completed by the Michaelis–Arbuzov reaction, radical, and transition-metal-complex-catalyzed phosphorylation, which are discussed below.

4.2.1. Michaelis–Arbuzov Reaction in Synthesis of Phosphorus-Substituted Pyrimidines

The Michaelis–Arbuzov reaction can probably be considered the most general method for synthesis of phosphorylpyrimidines known at the moment. Examples of use of the Michaelis–Arbuzov reaction in synthesis of pyrimidinylphosphonates (Scheme 28) were described in several studies [125,126,127,128,129]. Pyrimidines 113, 115, 117, and 119 containing the phosphonate moiety at the 2, 4, 5, or 6 position were obtained by reaction of pyrimidinyl halides 112, 114, 116, and 118, respectively, with phosphites. Several procedures requiring either microwave activation (line a) or the presence of a catalyst, such as acidic ion-exchange resin (line b) or LaCl3 (line c), were developed and shown to be tolerant to electron-donating and electron-withdrawing substituents on the pyrimidine ring. The chemoselective Michaelis–Arbuzov reaction was accomplished by Varalakshmi et al. [129], who prepared 2,6-dichloropyrimidine-2-phosphonates by coupling of 2,4,6-trichloropyrimidine using silica-gel-supported BF3 as the catalyst (line d).
A photochemical version of the Arbuzov reaction was shown to be efficient for synthesis of pyrimidine-5-phosphonates 121 from 5-bromopyrimidines 120 and trialkyl phosphites (Scheme 29) [130,131]. This radical process can be accomplished either directly under mild UV activation or using the Ru(bpy)3 pyrene dyad irradiated with blue light (455 nm). The plausible mechanism of this transformation involves formation of pyrimidinyl radical 122 either by action of an excited pyrene molecule or directly upon absorption of a light quantum. This radical attacks trialkyl phosphite, and the resulting radical 123 is decomposed to form the product accompanied by release of the alkyl radical.

4.2.2. Radical Phosphorylation of Pyrimidines

Radical functionalization of pyrimidines with phosphine oxides and phosphonates is poorly known. Only the photocatalytic radical reaction of 2-chloropyrimidines 124 with secondary phosphine oxides giving products 125 (Scheme 30) was recently described by Yuan et al. [73] The mechanism of this transformation is similar to that for chloropyridines and is considered in Section 2.4.1 (Scheme 6). As a distant example of the radical phosphorylation of pyrimidines, let us mention the work by Zhang et al. [132], who accomplished phosphorylation of pyrimidin-4-ones with dimethyl phosphite in the presence of Mn(OAc)3.

4.2.3. Transition-Metal-Catalyzed Phosphorylation of Pyrimidines

Since the first work on palladium-catalyzed phosphorylation of pyrimidines published in 2008 by Belabassi et al. [45], transition-metal-catalyzed cross-coupling of halopyrimidines with secondary phosphine oxides and phosphonates proved to be an efficient route to phosphorylpyrimidines. Mykhailiuk and co-workers [91] published a procedure for Pd(II)-catalyzed coupling of halo(Br/I)pyrimidines 126 with dimethylphosphine oxide, showing remarkable generality (Scheme 31). Synthesis of 2-, 4-, and 6-phosphine-oxide-substituted pyrimidines 127 bearing electron-donating and electron-withdrawing substituents on the heterocyclic moiety was accomplished using Pd2(dba)3 in the presence of the Xantphos ligand.
Individual examples of metal-complex-catalyzed coupling of 2-chloropyrimidines 128 with phosphine oxides were published in studies by Montchamp, Yang, Montel, Zakirova, and their co-workers (Scheme 32) [133,134,135,136]. Catalysis was accomplished using Pd(II), Pd(0), or Ni(II) salts, providing 2-phosphorylated pyrimidines 129 in good yield.
Bai et al. [137] reported one example of electrochemical phosphorylation of 5-bromopyrimidine (120′) with diethyl phosphonate in the presence of NiBr2 (Scheme 33), giving diethylpyrimidine-5-phosphonate 130. The authors suggested that the reaction proceeds via a radical mechanism accompanied by anodic generation of a phosphorus-centered radical.

5. Pyrazines

Phosphorus(V)-containing pyrazines are poorly studied six-membered N-heterocycles. Synthesis of phosphorylpyrazines was described or mentioned in just over a dozen publications, and, what is more important, most of them are fragmentary in scope. The group of Palacios in the 2000s [138,139,140]. made the most significant contribution to the development of original synthetic approaches to POR2-substituted pyrazines. The authors proposed self-dimerization of nitrile ylides and 4-dimethylamino-3-phosphoryl-2-azadienes and the formal [4+2]-cycloaddition of 1,2-diaza-1,3-butadienes with 1,2-diamines, which are considered below along with works on phosphorylation of the pyrazine ring (Figure 4).

5.1. Dimerization Reactions

Palacios et al. [138] described synthesis of phosphorus-substituted pyrazines via dimerization of nitrile ylides 133 (Scheme 34), which can be generated in situ from POR2-bearing 2H-azirines 131 (pathway A) or tosylated 2-hydroxyiminophosphonates 132 (pathway B). According to the proposed mechanism, intermediate 133 undergoes self-dimerization to dihydropyrazine 134 followed by either oxidation of the latter compound or elimination of secondary phosphine oxide. An interesting feature of this transformation is that it allows targeted synthesis of either mono- or bisphosphorus(V)-substituted pyrazines 135 and 136. Thus, synthesis via pathway A at 80 °C affords products 136, whereas heating to 110 °C yields products 135. The outcome of the reaction via pathway B is determined by substituents on the phosphorus(V) moiety. Thus, presence of a phenyl group results in formation of the monophosphorylated product, whereas the bisphosphorylated product is generated in the presence of the ethoxy group.
In another study, the same group [139] demonstrated that bis-2,5-phosphorylpyrazine 138 can be obtained via dimerization of aminoaldehyde 139 derived from 4-dimethylamino-3-phosphoryl-2-azadiene 137 via acid hydrolysis of both enamine and imine groups (Scheme 35). The reaction was proposed to occur through formation of dihydropyrazine 140, the oxidation of which in the reaction medium provides the final pyrazine.

5.2. Formal [4+2]-Cycloaddition

In the subsequent study, Palacios and co-workers [140] used the reaction of phosphorus-containing 1,2-diaza-1,3-butadienes 142 with 1,2-diamines 141 as access to polysubstituted POR2-modified pyrazines 143 (Scheme 36). This formal [4+2]-cycloaddition is a general method for synthesis of alkyl- and arylpyrazines bearing phosphine oxides and phosphonates substituents. The plausible mechanism of this reaction involves the following steps: the Michael addition of diamine 141 to diazadiene system 142, giving adduct 144, intramolecular attack of the second amino group on the C=N double bond of hydrazone, and elimination of ethyl hydrazinecarboxylate from the resulting piperazine 145 followed by oxidation of 1,2,3,4-tetrahydropyrazine 146 to pyrazine.

5.3. Phosphorylation of Pyrazines

Synthesis of 2-phosphorus(V)-substituted pyrazine based on phosphorylation was first described by Seggio et al. [141] in 2007. They reported an efficient two-step procedure for preparation of (diphenylphosphino)pyrazine oxide based on deprotonation of pyrazine with a mixture of ZnCl2‚ TMEDA, and LiTMP followed by treatment of the resulting lithium complex of di(pyrazin-2-yl)zinc with chlorodiphenylphosphine. Most of the later studies are scattered and describe single examples of phosphorylation of the pyridazine core via the Michaelis–Arbuzov reaction, radical, and transition-metal-catalyzed phosphorylation reactions, which are discussed below.

5.3.1. Michaelis–Arbuzov Reaction in the Synthesis of Phosphorus-Substituted Pyrazines

Despite more than a century of history, the Michaelis–Arbuzov reaction was relatively recently applied for phosphorylation of pyridazines. Thus, Golla et al. [128] synthesized 2-phosphorylpyrazine 148 from 2-chloropyrazine (147) and dimethyl phenylphosphonite using LaCl3•7H2O as the catalyst under neat conditions (Scheme 37, line a). Goddard and co-workers [131] reported an example of using the photo-Arbuzov reaction to prepare dimethyl pyrazine-2-phosphonate 150 from 2-bromopyrazine (149) and trimethyl phosphite under mild UVA irradiation (Scheme 37, line b).

5.3.2. Radical Phosphorylation of Pyrazines

Recently, Yuan et al. [73] described three examples of photocatalytic radical phosphorylation of 2-chloropyrazines 151 with secondary phosphine oxides, giving 2-substituted pyrazines 152 (Scheme 38, line a). The reaction occurred in the presence of tBuOK under mild conditions using irradiation with a blue-light-emitting diode. The mechanism of transformation is similar to that discussed for pyridines in Chapter 2.4.1 (Scheme 6). Further, Berger and Montchamp [142] reported synthesis of single 2-phosphorylated pyrazine 154 by reaction of pyridazine (153) with phosphonite in the presence of MnO2 and Mn(OAc)3 (Scheme 38, line b). The authors suggested that the reaction proceeds through formation of a phosphorus-centered radical that adds to pyrazine.

5.3.3. Transition-Metal-Catalyzed Phosphorylation of Pyrazines

It was shown that 2-chloropyrimidines 155 have good synthetic potential as precursors for synthesis of 2-phosphorus-substituted pyrazines 156 under conditions of metal-catalyzed phosphorylation (Scheme 39). The Pd(II)-catalyzed phosphorylation of chloropyrazine was described in studies by Belabassi, Deal, Nikishin, and their co-workers [40,45,133,143] (conditions 1–3). The generality of the approach with respect to phosphine oxides and phosphonates was demonstrated using Pd(dppf)Cl3 as the catalyst (condition 3). Note that Zhao et al. [144] performed this transformation in the presence of NiCl2(dppp) complexes as the catalyst (condition 4). Further, Yamaguchi and co-workers [100] described one example of nickel(II)-acetate-catalyzed phosphorylation of phenyl pyrazine-2-carboxylate with diphenylphosphine oxide accompanied by decarbonylation.

6. Conclusions

In conclusion, in the past decade, significant progress was made in synthesis of phosphorus-substituted six-membered aromatic heterocycles. Numerous interesting synthetic approaches were developed based on construction of a heterocyclic core using phosphoryl-containing reagents, mainly formal cycloaddition and intra(inter)molecular cyclization/isomerization, using the Michael reaction, carbenoid intermediates, and classical nucleophilic–electrophilic interactions. Their apparent advantages are structurally diverse products and high tolerance to functional groups. Further, phosphorylation represents the most general route to phosphorus-substituted pyridines, pyridazines, pyrimidines, and pyrazines. Halogen-substituent in six-membered N-heterocycles has proven to be a universal leaving group providing phosphorylation, and some functional groups (OH, OTs, B(OH)2, CO2Ph, N=NSO2Me, etc.) were also efficient in this process. However, the necessity of finding conditions for each particular type of heterocycles, severe reaction conditions (high temperature, presence of strong bases and acids), and high cost of reagent/catalysts significantly limit practical application of this approach.
Evidently, development of synthetic methods based on phosphorylation of heterocycles and heterocyclization involving phosphorus-substituted reagents is of fundamental research interest. In point of fact, pyridines have received a great deal of attention among monocyclic six-membered N-heterocycles in the past decade, resulting in development of a number of approaches with good substrate scope and efficiency. Most of the methods proposed for pyridazines, pyrimidines, and pyrazines are fragmented and sporadic, while phosphorus-substituted triazines and tetrazines are not described. In this regard, further development of this research area is associated with the following issues: design of new versatile multifunctional reagents for various chemical transformations (detailed investigation of such new reagents would provide the basis for a new area in synthesis of nitrogen-containing heterocyclic compounds with different combinations of heteroatoms and also for preparation of new complex heterocyclic systems and assemblies); transition-metal-catalyzed C–H phosphorylation of six-memebered N-heterocycles (transition-metal-catalyzed CH phosphorylation is a straightforward and attractive approach to construct C–P bonds; however, it is difficult to execute this approach in practice because of the strong propensity of phosphorus reagents to induce catalyst poisoning through coordination); development of methods for synthesis of phosphorus-substituted six-memebered N-heterocycles with more than two nitrogen atoms in the ringe (no representatives of such compounds have been described to date). Since there are a few examples of synthesis of certain phosphorus(V)-substituted six-membered heterocycles while synthetic approaches to phosphorylated derivatives of a wide range of particular N-heterocycles are still absent, further research in this field is a long-term challenge.

Author Contributions

Conceptualization, writing, editing, and submitting, Y.V.; Conceptualization, I.Z. All authors have read and agreed to the published version of the manuscript.

Funding

This work was financially supported by the Russian Science Foundation (project No. 22-13-00161).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

We also acknowledge Mikhail Kozlov for technical assistance in preparing the manuscript.

Conflicts of Interest

The authors declare no conflict of interest.

Sample Availability

Not applicable.

Abbreviations

N-methyl-D-aspartate (NMDA); dibenzylideneacetone (dba); 1,8-diazabicyclo [5.4.0]undec-7-ene (DBU); density functional theory (DFT); 1,1′-bis(diphenylphosphino)ferrocene (dppf); 1,3-bis(diphenylphosphino)propane (dppp); 4,4’-dipyridyl (dpy); lithium tetramethylpiperidide (LiTMP); tetramethylethylenediamine (TMEDA); ultraviolet radiation (UVA); 4,5-bis(diphenylphosphino)-9,9-dimethylxanthene (Xantphos).

References

  1. Duffy, M.P.; Delaunay, W.; Bouit, P.-A.; Hissler, M. π-Conjugated phospholes and their incorporation into devices: Components with a great deal of potential. Chem. Soc. Rev. 2016, 45, 5296–5310. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  2. Tang, W.; Zhang, X. New chiral phosphorus ligands for enantioselective hydrogenation. Chem. Rev. 2003, 103, 3029–3070. [Google Scholar] [CrossRef] [PubMed]
  3. Surry, D.S.; Buchwald, S.L. Biaryl phosphane ligands in palladium-catalyzed amination. Angew. Chem. Int. Ed. 2008, 47, 6338–6361. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  4. Bayne, J.; Stephan, D. Phosphorus Lewis acids: Emerging reactivity and applications in catalysis. Chem. Soc. Rev. 2016, 45, 765–774. [Google Scholar] [CrossRef] [PubMed]
  5. Nixon, T.D.; Gamble, A.J.; Thatcher, R.J.; Whitwood, A.C.; Lynam, J.M. Synthesis and coordination chemistry of pyrimidine-substituted phosphine ligands. Inorg. Chim. Acta 2012, 380, 252–260. [Google Scholar] [CrossRef]
  6. Cross, J.M.; Gallagher, N.; Gill, J.H.; Jain, M.; McNeillis, A.W.; Rockley, K.L.; Tscherny, F.H.; Wirszycz, N.J.; Yufit, D.S.; Walton, J.W. Pyridylphosphinate metal complexes: Synthesis, structural characterisation and biological activity. Dalton Trans. 2016, 45, 12807–12813. [Google Scholar] [CrossRef] [Green Version]
  7. Artem’ev, A.V.; Eremina, J.A.; Lider, E.V.; Antonova, O.V.; Vorontsova, E.V.; Bagryanskaya, I.Y. Luminescent Ag (I) scorpionates based on tris(2-pyridyl)phosphine oxide: Synthesis and cytotoxic activity evaluation. Polyhedron 2017, 138, 218–224. [Google Scholar] [CrossRef]
  8. Shudo, N.; Mizoguchi, T.; Kiyosue, T.; Arita, M.; Yoshimura, A.; Seto, K.; Sakoda, R.; Akiyama, S.-I. Two pyridine analogues with more effective ability to reverse multidrug resistance and with lower calcium channel blocking activity than their dihydropyridine counterparts. Cancer Res. 1990, 50, 3055–3061. [Google Scholar]
  9. Pfefferkorn, J.A.; Guzman-Perez, A.; Litchfield, J.; Aiello, R.; Treadway, J.L.; Pettersen, J.; Minich, M.L.; Filipski, K.J.; Jones, C.S.; Tu, M. Discovery of (S)-6-(3-cyclopentyl-2-(4-(trifluoromethyl)-1H-imidazol-1-yl)propanamido) nicotinic acid as a hepatoselective glucokinase activator clinical candidate for treating type 2 diabetes mellitus. J. Med. Chem. 2012, 55, 1318–1333. [Google Scholar] [CrossRef] [PubMed]
  10. Murray, K.J.; Porter, R.A.; Warrington, B.H.; Lahouratate, P. Phenol and Pyridinol Derivatives as Lusitropic Agents. Patent WO1993019754A1, 23 March 1993. [Google Scholar]
  11. Ali, T.E.; Assiri, M.A.; El-Shaaer, H.M.; Abed-Kariem, S.M.; Abdel-Monem, W.R.; El-Edfawy, S.M.; Hassanin, N.M.; Shati, A.A.; Alfaifi, M.Y.; Elbehairi, S.E.I. Synthesis and biological activities of some new phosphorus compounds containing pyranopyrazole moiety. Heterocycles 2021, 102, 1119–1137. [Google Scholar] [CrossRef]
  12. de Assis, T.M.; Gajo, G.C.; de Assis, L.C.; Garcia, L.S.; Silva, D.R.; Ramalho, T.C.; da Cunha, E.F.F. QSAR models guided by molecular dynamics applied to human glucokinase activators. Chem. Biol. Drug Des. 2016, 87, 455–466. [Google Scholar] [CrossRef]
  13. Dziuganowska, Z.A.; Slepokura, K.; Volle, J.-N.; Virieux, D.; Pirat, J.-L.; Kafarski, P. Structural analogues of Selfotel. J. Org. Chem. 2016, 81, 4947–4954. [Google Scholar] [CrossRef] [PubMed]
  14. Khandelwal, A.; Lukacova, V.; Comez, D.; Kroll, D.M.; Raha, S.; Balaz, S. A combination of docking, QM/MM methods, and MD simulation for binding affinity estimation of metalloprotein ligands. J. Med. Chem. 2005, 48, 5437–5447. [Google Scholar] [CrossRef] [Green Version]
  15. Venkata Ramana, K.; Venkata Subbaiah, K.; Lokanatha, V.; Naga Raju, C. Synthesis and antimicrobial activity evaluation of new dialkyl heteroarylphosphonates. Der. Pharm. Chem. 2011, 3, 181–188. [Google Scholar]
  16. Gardner, G.; Steffens, J.J.; Grayson, B.T.; Kleier, D.A. 2-Methylcinnolinium herbicides: Effect of 2-methylcinnolinium-4-(O-methylphosphonate) on photosynthetic electron transport. J. Agric. Food Chem. 1992, 40, 318–321. [Google Scholar] [CrossRef]
  17. DeFrees, S.A.; Sawick, D.P.; Cunningham, B.; Heinstein, P.F.; Morré, D.J.; Cassady, J.M. Structure-activity relationships of pyrimidines as dihydroorotate dehydrogenase inhibitors. Biochem. Pharmacol. 1988, 37, 3807–3816. [Google Scholar] [CrossRef] [PubMed]
  18. Stansfield, I.; Avolio, S.; Colarusso, S.; Gennari, N.; Narjes, F.; Pacini, B.; Ponzi, S.; Harper, S. Active site inhibitors of HCV NS5B polymerase. The development and pharmacophore of 2-thienyl-5, 6-dihydroxypyrimidine-4-carboxylic acid. Bioorganic Med. Chem. Lett. 2004, 14, 5085–5088. [Google Scholar] [CrossRef] [PubMed]
  19. Zhang, H.-w.; Zhou, L.; Coats, S.J.; McBrayer, T.R.; Tharnish, P.M.; Bondada, L.; Detorio, M.; Amichai, S.A.; Johns, M.D.; Whitaker, T. Synthesis of purine modified 2′-C-methyl nucleosides as potential anti-HCV agents. Bioorganic Med. Chem. Lett. 2011, 21, 6788–6792. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  20. Teng, M.; Johnson, M.D.; Thomas, C.; Kiel, D.; Lakis, J.N.; Kercher, T.; Aytes, S.; Kostrowicki, J.; Bhumralkar, D.; Truesdale, L. Small molecule ago-allosteric modulators of the human glucagon-like peptide-1 (hGLP-1) receptor. Bioorganic Med. Chem. Lett. 2007, 17, 5472–5478. [Google Scholar] [CrossRef]
  21. Kumar, A.V.; Mohan, K. Insights into binding of potential antitumor quinoxaline analogues against cyclin dependent kinase 2 using docking studies. J. Chem. Biol. Phys. Sci. 2012, 2, 2419. [Google Scholar]
  22. Kozlov, M.; Komkov, A.; Losev, T.; Tyurin, A.; Dmitrenok, A.; Zavarzin, I.; Volkova, Y. Flexible synthesis of phosphoryl-substituted imidazolines, tetrahydropyrimidines, and thioamides by sulfur-mediated processes. J. Org. Chem. 2019, 84, 11533–11541. [Google Scholar] [CrossRef] [PubMed]
  23. Kozlov, M.; Kozlov, A.; Komkov, A.; Lyssenko, K.; Zavarzin, I.; Volkova, Y. Synthesis of phosphoryl thioamides via three-component reaction of phosphinic chlorides with amines and sulfur. Adv. Synth. Catal. 2019, 361, 2904–2915. [Google Scholar] [CrossRef]
  24. Van der Jeught, S.; Stevens, C.V. Direct phosphonylation of aromatic azaheterocycles. Chem. Rev. 2009, 109, 2672–2702. [Google Scholar] [CrossRef] [PubMed]
  25. Chen, L.; Liu, X.Y.; Zou, Y.X. Recent advances in the construction of phosphorus-substituted heterocycles, 2009–2019. Adv. Synth. Catal. 2020, 362, 1724–1818. [Google Scholar] [CrossRef]
  26. Liu, J.; Xiao, H.-Z.; Fu, Q.; Yu, D.-G. Advances in radical phosphorylation from 2016 to 2021. Chem. Synth. 2021, 1, 9. [Google Scholar] [CrossRef]
  27. Budnikova, Y.H.; Sinyashin, O.G.D. Phosphorylation of C–H bonds of aromatic compounds using metals and metal complexes. Russ. Chem. Rev. 2015, 84, 917–951. [Google Scholar] [CrossRef]
  28. Luo, K.; Yang, W.C.; Wu, L. Photoredox catalysis in organophosphorus chemistry. Asian J. Org. Chem. 2017, 6, 350–367. [Google Scholar] [CrossRef]
  29. Budnikova, Y.H.; Gryaznova, T.V.; Grinenko, V.V.; Dudkina, Y.B.; Khrizanforov, M.N. Eco-efficient electrocatalytic C–P bond formation. Pure Appl. Chem. 2017, 89, 311–330. [Google Scholar] [CrossRef]
  30. Pan, X.-Q.; Zou, J.-P.; Yi, W.-B.; Zhang, W. Recent advances in sulfur-and phosphorous-centered radical reactions for the formation of S–C and P–C bonds. Tetrahedron 2015, 40, 7481–7529. [Google Scholar] [CrossRef]
  31. Haji, M. Multicomponent reactions: A simple and efficient route to heterocyclic phosphonates. Beilstein J. Org. Chem. 2016, 12, 1269–1301. [Google Scholar] [CrossRef]
  32. Gao, Y.; Tang, G.; Zhao, Y. Recent progress toward organophosphorus compounds based on phosphorus-centered radical difunctionalizations. Phosphorus Sulfur Silicon Relat. Elem. 2017, 192, 589–596. [Google Scholar] [CrossRef]
  33. Demmer, C.S.; Krogsgaard-Larsen, N.; Bunch, L. Review on modern advances of chemical methods for the introduction of a phosphonic acid group. Chem. Rev. 2011, 111, 7981–8006. [Google Scholar] [CrossRef] [PubMed]
  34. Volkova, Y.; Kozlov, M.; Zavarzin, I. Synthetic Routes to P(O)-Substituted Five-Membered Aromatic Heterocycles (2010–2021). Targets Heterocycl. Syst. Chem. Prop. 2022, 26, 18–45. [Google Scholar]
  35. Egorova, A.V.; Svintsitskaya, N.I.; Dogadina, A.V. Synthesis of phosphorylated indoles. Russ. J. Gen. Chem. 2018, 88, 2276–2289. [Google Scholar] [CrossRef]
  36. Chen, L.; Zou, Y.X. Recent progress in the synthesis of phosphorus-containing indole derivatives. Org. Biomol. Chem. 2018, 16, 7544–7556. [Google Scholar] [CrossRef]
  37. Baiju, T.; Namboothiri, I.N. Synthesis of functionalized pyrazoles via 1,3-dipolar cycloaddition of α-diazo-β-ketophosphonates, sufones and esters with electron-deficient alkenes. Chem. Rec. 2017, 17, 939–955. [Google Scholar] [CrossRef]
  38. Goulioukina, N.S.; Makukhin, N.N.; Beletskaya, I.P. Synthetic routes to 3(5)-phosphonylated pyrazoles. Russ. Chem. Rev. 2016, 85, 667–683. [Google Scholar] [CrossRef]
  39. Gazizov, A.S.; Smolobochkin, A.V.; Turmanov, R.A.; Pudovik, M.A.; Burilov, A.R.; Sinyashin, O.G. Synthesis of phosphaproline derivatives: A short overview. Synthesis 2019, 51, 3397–3409. [Google Scholar] [CrossRef]
  40. Nikishkin, N.I.; Huskens, J.; Assenmacher, J.; Wilden, A.; Modolo, G.; Verboom, W. Palladium-catalyzed cross-coupling of various phosphorus pronucleophiles with chloropyrazines: Synthesis of novel Am(III)-selective extractants. Org. Biomol. Chem. 2012, 10, 5443–5451. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  41. Deng, L.; Diao, J.; Chen, P.; Pujari, V.; Yao, Y.; Cheng, G.; Crick, D.C.; Prasad, B.V.; Song, Y. Inhibition of 1-deoxy-D-xylulose-5-phosphate reductoisomerase by lipophilic phosphonates: SAR, QSAR, and crystallographic studies. J. Med. Chem. 2011, 54, 4721–4734. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  42. Perry, H.; Zoń, J.; Law, J.; Clearfield, A. Structural variations of SnII pyridylphosphonates influenced by an uncommon Sn–N interaction. J. Solid State Chem. 2010, 183, 1165–1173. [Google Scholar] [CrossRef]
  43. Zhang, J.; Wang, J.; Wei, C.; Wang, Y.; Xie, G.; Li, Y.; Li, M. Rapidly sequence-controlled electrosynthesis of organometallic polymers. Nat. Commun. 2020, 11, 2530. [Google Scholar] [CrossRef] [PubMed]
  44. Barycki, J.; Gancarz, R.; Milewska, M.; Tyka, R.; Sawka-Dobrowolska, W. 1-Phthalazinephosphonic acid-An unexpected product of the reaction between diisopropylphosphite and o-phthalazine. Phosphorus Sulfur Silicon Relat. Elem. 1998, 143, 167–178. [Google Scholar] [CrossRef]
  45. Belabassi, Y.; Alzghari, S.; Montchamp, J.-L. Revisiting the Hirao cross-coupling: Improved synthesis of aryl and heteroaryl phosphonates. J. Organomet. Chem. 2008, 693, 3171–3178. [Google Scholar] [CrossRef] [Green Version]
  46. Plazek, S.; Sasyk, Z. Uber einige phosphoe-pyridinverbinddungen. Chemishes Zent. 1935, 106, 2177. [Google Scholar]
  47. Collins, D.J.; Hetherington, J.W.; Swan, J.M. Organophosphorus compounds. XIV. A new synthesis of 3-pyridylphosphonic acid. Aust. J. Chem. 1974, 27, 1355–1360. [Google Scholar] [CrossRef]
  48. Wu, J.; Chen, H.; Zhou, Z.-Y.; Yeung, C.H.; Chan, A.S.C. Synthesis and structural characterization of a highly effective chiral dipyridylphosphine ligand and its application in the Ru-catalyzed asymmetric hydrogenation of β-ketoesters. Synlett 2001, 2001, 1050–1054. [Google Scholar] [CrossRef] [Green Version]
  49. Wu, J.; Chen, H.; Kwok, W.H.; Lam, K.H.; Zhou, Z.Y.; Yeung, C.H.; Chan, A.S.C. A new chiral dipyridylphosphine ligand Xyl-P-Phos and its application in the Ru-catalyzed asymmetric hydrogenation of β-ketoesters. Tetrahedron Lett. 2002, 43, 1539–1543. [Google Scholar] [CrossRef]
  50. Bennett, R.; Burger, A.; Volk, W. Communications—3-Pyridylphosphonic acid. J. Org. Chem. 1958, 23, 940. [Google Scholar] [CrossRef]
  51. Redmore, D. Phosphorus derivatives of nitrogen heterocycles. 2. Pyridinephosphonic acid derivatives. J. Org. Chem. 1970, 35, 4114–4117. [Google Scholar] [CrossRef]
  52. Eliseenkov, V.N.; Bureva, N.V.; Ivanov, B.E. Some transformations in a number of (2-pyridyl)phosphonates. Chem. Heterocycl. Compd. 1974, 10, 1182–1184. [Google Scholar] [CrossRef]
  53. Predvoditelev, D.A.; Chukbar, T.G.; Nifant’ev, E. Some new derivatives of pyridyl- and quinolylphosphonic acids. Chem. Heterocycl. Compd. 1975, 11, 330–333. [Google Scholar] [CrossRef]
  54. Bulot, J.J.; Aboujaoude, E.E.; Collignon, N.; Savignac, P. Preparation d’aminophenyl-, nitrophenyl-, pyridyl-, et quinolylphosphonates sous photostimulation ou assistance metallique; acces aux acides aminophosphoniques correspondants. Phosphorus Sulfur Relat. Elem. 2006, 21, 197–204. [Google Scholar] [CrossRef]
  55. Boduszek, B. Pyridine-2-phospho-carboxylic acids: Synthesis and properties. J. Prakt. Chem. Chem. Ztg. 1992, 334, 444–446. [Google Scholar] [CrossRef]
  56. Boduszek, B.; Wieczorek, J.S. A new method for the preparation of pyridine-4-phosphonic acids. Synthesis 1979, 1979, 452–453. [Google Scholar] [CrossRef]
  57. Markovskii, L.N.; Furin, G.G.; Shermolovich, Y.G.; Yakobson, G.G. Phosphorylation of polyfluoroaromatic compounds. Bull. Acad. Sci. USSR Div. Chem. Sci. 1981, 30, 646–648. [Google Scholar] [CrossRef]
  58. Boenigk, W.; Fischer, U.; Hägele, G. Über die michaelis-arbuzov-reaktion perhalogenierter pyridine. II1. Phosphorus Sulfur Relat. Elem. 2007, 16, 263–266. [Google Scholar] [CrossRef]
  59. Penicaud, V.; Odobel, F.; Bujoli, B. Facile and efficient syntheses of 2,2′-bipyridine-based bis(phosphonic) acids. Tetrahedron Lett. 1998, 39, 3689–3692. [Google Scholar] [CrossRef]
  60. Adam, M.S.S.; Kühl, O.; Kindermann, M.K.; Heinicke, J.W.; Jones, P.G. 3-Amino- and 3-acylamido-2-phosphonopyridines: Synthesis by Pd-catalyzed P–C coupling, structure and conversion to pyrido[b]-anellated P=C–N heterocycles. Tetrahedron 2008, 64, 7960–7967. [Google Scholar] [CrossRef]
  61. Palacios, F.; Gil, M.J.; de Marigorta, E.M.; Rodríguez, M. Synthesis and reactivity of imines derived from bisphosphonates and 3-phosphorylated 2-aza-1,3-dienes. Tetrahedron 2000, 56, 6319–6330. [Google Scholar] [CrossRef]
  62. Palacios, F.; Aparicio, D.; López, Y.; de los Santos, J.M.; Ezpeleta, J.M. Preparation and reactions of 3-phosphinyl-1-aza-1,3-butadienes. Synthesis of phosphorylated pyridine and pyrazole derivatives. Tetrahedron 2006, 62, 1095–1101. [Google Scholar] [CrossRef]
  63. Allais, C.; Liéby-Muller, F.; Rodriguez, J.; Constantieux, T. Metal-free michael-addition-initiated three-component reaction for the regioselective synthesis of highly functionalized pyridines: Scope, mechanistic investigations and applications. Eur. J. Org. Chem. 2013, 2013, 4131–4145. [Google Scholar] [CrossRef]
  64. Allais, C.; Constantieux, T.; Rodriguez, J. Use of beta,gamma-unsaturated alpha-ketocarbonyls for a totally regioselective oxidative multicomponent synthesis of polyfunctionalized pyridines. Chemistry 2009, 15, 12945–12948. [Google Scholar] [CrossRef] [PubMed]
  65. Allais, C.; Liéby-Muller, F.; Constantieux, T.; Rodriguez, J. Dual heterogeneous catalysis for a regioselective three-component synthesis of bi- and tri(hetero)arylpyridines. Adv. Synth. Catal. 2012, 354, 2537–2544. [Google Scholar] [CrossRef]
  66. Hanashalshahaby, E.H.; Unaleroglu, C. Mannich bases as enone precursors for water-mediated efficient synthesis of 2,3,6-trisubstituted pyridines and 5,6,7,8-tetrahydroquinolines. ACS Comb. Sci. 2015, 17, 374–380. [Google Scholar] [CrossRef]
  67. Abdou, W.M.; Shaddy, A.A.; Kamel, A.A. Structure-based design and synthesis of acyclic and substituted heterocyclic phosphonates linearly linked to thiazolobenzimidazoles as potent hydrophilic antineoplastic agents. Chem. Pap. 2017, 71, 1961–1973. [Google Scholar] [CrossRef]
  68. Liao, L.; Zhang, H.; Zhao, X. Selenium-π-acid catalyzed oxidative functionalization of alkynes: Facile access to ynones and multisubstituted oxazoles. ACS Catal. 2018, 8, 6745–6750. [Google Scholar] [CrossRef]
  69. Qi, X.; Dai, L.; Park, C.M. Carbenoid-mediated N-O bond insertion and its application in the synthesis of pyridines. Chem. Commun. 2012, 48, 11244–12246. [Google Scholar] [CrossRef]
  70. Loy, N.S.; Singh, A.; Xu, X.; Park, C.M. Synthesis of pyridines by carbenoid-mediated ring opening of 2H-azirines. Angew. Chem. Int. Ed. 2013, 52, 2212–2216. [Google Scholar] [CrossRef]
  71. Kashima, K.; Ishii, M.; Tanaka, K. Synthesis of pyridylphosphonates by rhodium-catalyzed [2+2+2] cycloaddition of 1,6- and 1,7-diynes with diethyl phosphorocyanidate. Eur. J. Org. Chem. 2015, 2015, 1092–1099. [Google Scholar] [CrossRef]
  72. Kashima, K.; Teraoka, K.; Uekusa, H.; Shibata, Y.; Tanaka, K. Rhodium-catalyzed atroposelective [2+2+2] cycloaddition of ortho-substituted phenyl diynes with nitriles: Effect of ortho substituents on regio- and enantioselectivity. Org. Lett. 2016, 18, 2170–2173. [Google Scholar] [CrossRef] [PubMed]
  73. Yuan, J.; To, W.P.; Zhang, Z.Y.; Yue, C.D.; Meng, S.; Chen, J.; Liu, Y.; Yu, G.A.; Che, C.M. Visible-light-promoted transition-metal-free phosphinylation of heteroaryl halides in the presence of potassium tert-butoxide. Org. Lett. 2018, 20, 7816–7820. [Google Scholar] [CrossRef]
  74. Zeng, H.; Dou, Q.; Li, C.-J. Photoinduced transition-metal-free cross-coupling of aryl halides with H-phosphonates. Org. Lett. 2019, 21, 1301–1305. [Google Scholar] [CrossRef] [PubMed]
  75. Liu, Y.; Guo, W. Visible-light driven C-P bond formation with recyclable carbon nitride photocatalyst. ChemCatChem 2022, 14, e202200449. [Google Scholar] [CrossRef]
  76. Cowper, N.G.; Chernowsky, C.P.; Williams, O.P.; Wickens, Z.K. Potent reductants via electron-primed photoredox catalysis: Unlocking aryl chlorides for radical coupling. J. Am. Chem. Soc. 2020, 142, 2093–2099. [Google Scholar] [CrossRef]
  77. Zhu, D.-L.; Jiang, S.; Wu, Q.; Wang, H.; Chai, L.-L.; Li, H.-Y.; Li, H.-X. Visible-light-induced nickel-catalyzed P(O)–C(sp2) coupling using thioxanthen-9-one as a photoredox catalysis. Org. Lett. 2020, 23, 160–165. [Google Scholar] [CrossRef]
  78. Qiu, D.; Lian, C.; Mao, J.; Ding, Y.; Liu, Z.; Wei, L.; Fagnoni, M.; Protti, S. Visible Light-Driven, Photocatalyst-free Arbuzov-like reaction via arylazo sulfones. Adv. Synth. Catal. 2019, 361, 5239–5244. [Google Scholar] [CrossRef]
  79. Kim, I.; Kang, G.; Lee, K.; Park, B.; Kang, D.; Jung, H.; He, Y.T.; Baik, M.H.; Hong, S. Site-selective functionalization of pyridinium derivatives via visible-light-driven photocatalysis with quinolinone. J. Am. Chem. Soc. 2019, 141, 9239–9248. [Google Scholar] [CrossRef] [PubMed]
  80. Xiang, C.B.; Bian, Y.J.; Mao, X.R.; Huang, Z.Z. Coupling reactions of heteroarenes with phosphites under silver catalysis. J. Org. Chem. 2012, 77, 7706–7710. [Google Scholar] [CrossRef]
  81. Kittikool, T.; Phakdeeyothin, K.; Chantarojsiri, T.; Yotphan, S. Manganese-promoted regioselective direct C3-phosphinoylation of 2-pyridones. Eur. J. Org. Chem. 2021, 2021, 3071–3078. [Google Scholar] [CrossRef]
  82. Zhou, Y.; Liu, C.; Shen, Z.; Dai, B.; Chen, J. Efficient potassium hydroxide promoted P-arylation of aryl halides with diphenylphosphine. J. Organomet. Chem. 2021, 949, 121932. [Google Scholar] [CrossRef]
  83. Trofimov, B.A.; Volkov, P.A.; Khrapova, K.O.; Telezhkin, A.A.; Ivanova, N.I.; Albanov, A.I.; Gusarova, N.K.; Chupakhin, O.N. Metal-free site selective cross-coupling of pyridines with secondary phosphine chalcogenides using acylacetylenes as oxidants. Chem. Commun. 2018, 54, 3371–3374. [Google Scholar] [CrossRef]
  84. Volkov, P.A.; Gusarova, N.K.; Khrapova, K.O.; Telezhkin, A.A.; Albanov, A.I.; Vasilevskiy, S.F.; Trofimov, B.A. A mechanistic insight into the chemoselectivity of the reaction between 3-phenyl-2-propynenitrile, secondary phosphine oxides and pyridinoids. Mendeleev Commun. 2021, 31, 670–672. [Google Scholar] [CrossRef]
  85. Volkov, P.A.; Telezhkin, A.A.; Khrapova, K.O.; Ivanova, N.I.; Albanov, A.I.; Gusarova, N.K.; Trofimov, B.A. Metal-free S HN cross-coupling of pyridines with phosphine chalcogenides: Polarization/deprotonation/oxidation effects of electron-deficient acetylenes. New J. Chem. 2021, 45, 6206–6219. [Google Scholar] [CrossRef]
  86. Oka, N.; Ito, K.; Tomita, F.; Ando, K. Synthesis of 2-pyridylphosphinate and thiophosphinate derivatives by nucleophilic aromatic substitution of N-methoxypyridinium tosylates. Chem. Lett. 2012, 41, 1630–1632. [Google Scholar] [CrossRef]
  87. Wang, H.; Cui, X.; Pei, Y.; Zhang, Q.; Bai, J.; Wei, D.; Wu, Y. Direct regioselective phosphonation of heteroaryl N-oxides with H-phosphonates under metal and external oxidant free conditions. Chem. Commun. 2014, 50, 14409–14411. [Google Scholar] [CrossRef] [PubMed]
  88. Lee, S.-J.; Kim, H.-S.; Yang, H.-W.; Yoo, B.-W.; Yoon, C.M. Synthesis of diethyl pyridin-2-ylphosphonates and quinolin-2-ylphosphonates by deoxygenative phosphorylation of the corresponding N-oxides. Bull. Korean Chem. Soc. 2014, 35, 2155–2158. [Google Scholar] [CrossRef] [Green Version]
  89. Li, S.G.; Yuan, M.; Topic, F.; Han, Z.S.; Senanayake, C.H.; Tsantrizos, Y.S. Asymmetric library synthesis of P-chiral t-butyl-substituted secondary and tertiary phosphine oxides. J. Org. Chem. 2019, 84, 7291–7302. [Google Scholar] [CrossRef] [PubMed]
  90. Han, Z.S.; Wu, H.; Xu, Y.; Zhang, Y.; Qu, B.; Li, Z.; Caldwell, D.R.; Fandrick, K.R.; Zhang, L.; Roschangar, F.; et al. General and stereoselective method for the synthesis of sterically congested and structurally diverse P-stereogenic secondary phosphine oxides. Org. Lett. 2017, 19, 1796–1799. [Google Scholar] [CrossRef]
  91. Stambirskyi, M.V.; Kostiuk, T.; Sirobaba, S.I.; Rudnichenko, A.; Titikaiev, D.L.; Dmytriv, Y.V.; Kuznietsova, H.; Pishel, I.; Borysko, P.; Mykhailiuk, P.K. Phosphine oxides (− POMe2) for medicinal chemistry: Synthesis, properties, and applications. J. Org. Chem. 2021, 86, 12783–12801. [Google Scholar] [CrossRef]
  92. Zakirova, G.G.; Mladentsev, D.Y.; Borisova, N.E. Synthesis of chelating tertiary phosphine oxides via palladium-catalysed C–P bond formation. Tetrahedron Lett. 2017, 58, 3415–3417. [Google Scholar] [CrossRef]
  93. Henyecz, R.; Oroszy, R.; Keglevich, G. Microwave-assisted Hirao reaction of heteroaryl bromides and >P(O)H reagents using Pd(OAc)2 as the catalyst precursor in the absence of added P-ligands. Curr. Org. Chem. 2019, 23, 1151–1157. [Google Scholar] [CrossRef]
  94. McErlain, H.; Riley, L.M.; Sutherland, A. Palladium-catalyzed C–P bond-forming reactions of aryl nonaflates accelerated by iodide. J. Org. Chem. 2021, 86, 17036–17049. [Google Scholar] [CrossRef]
  95. Ma, F.; Zhang, T.-T.; Zhang, Z.-H.; Tong, H.-X.; Yi, X.-Y. Photorelease of nitric oxide in water-soluble diruthenium nitrosyl complexes with phosphonate substituted pyridylpyrrole. Inorg. Chim. Acta 2022, 534, 120826. [Google Scholar] [CrossRef]
  96. Adam, M.S.S.; Kindermann, M.K.; Köckerling, M.; Heinicke, J.W. Phosphonylation of 2-amino- and 2-amido-3-bromopyridines and 2-amino-3-chloroquinoxalines with triethyl phosphite. Eur. J. Org. Chem. 2009, 2009, 4655–4665. [Google Scholar] [CrossRef]
  97. Geng, Z.; Zhang, Y.; Zheng, L.; Li, J.; Zou, D.; Wu, Y.; Wu, Y. Pd-catalyzed C–P coupling of heteroaryl boronic acid with H-phosphonate diester. Tetrahedron Lett. 2016, 57, 3063–3066. [Google Scholar] [CrossRef]
  98. Zhang, G.; Wang, J.; Guan, C.; Zhao, Y.; Ding, C. Nickel- and palladium-catalyzed cross-coupling of aryl fluorosulfonates and phosphites: Synthesis of aryl phosphonates. Eur. J. Org. Chem. 2021, 2021, 810–813. [Google Scholar] [CrossRef]
  99. Zhao, Y.L.; Wu, G.J.; Han, F.S. Ni-catalyzed construction of C-P bonds from electron-deficient phenols via the in situ aryl C-O activation by PyBroP. Chem. Commun. 2012, 48, 5868–5870. [Google Scholar] [CrossRef]
  100. Isshiki, R.; Muto, K.; Yamaguchi, J. Decarbonylative C-P bond formation using aromatic esters and organophosphorus compounds. Org. Lett. 2018, 20, 1150–1153. [Google Scholar] [CrossRef]
  101. Yang, B.; Wang, Z.-X. Ni-Catalyzed C–P coupling of aryl, benzyl, or allyl ammonium salts with P(O)H compounds. J. Org. Chem. 2019, 84, 1500–1509. [Google Scholar] [CrossRef]
  102. Li, C.-J. Nickel-catalyzed phosphorylation of tosylates. Russ. J. Gen. Chem. 2020, 90, 725–730. [Google Scholar] [CrossRef]
  103. Li, C.J. Aryltrimethylammonium tetrafluoroborates in nickel-catalyzed C–P bond-forming reactions. Russ. J. Org. Chem. 2021, 57, 954–960. [Google Scholar] [CrossRef]
  104. Regitz, M.; Heydt, A.; Weber, B. (Diazomethyl)cyclopropene—Synthese, isomerisierung und carbenreaktionen. Angew. Chem. 1979, 91, 566–568. [Google Scholar] [CrossRef]
  105. Heydt, A.; Heydt, H.; Weber, B.; Regitz, M. Untersuchungen an diazoverbindungen und aziden, XLII. (Diazomethyl)cyclopropene durch elektrophile diazoalkan-substitution. Chem. Ber. 1982, 115, 2965–2980. [Google Scholar] [CrossRef]
  106. Eisenbarth, P.; Regitz, M. Untersuchungen an diazoverbindungen und aziden, XLIII1). 1,2-Dewarpyridazine—Vorstufen zur erzeugung von azacyclobutadienen? Chem. Ber. 1984, 117, 445–454. [Google Scholar] [CrossRef]
  107. Heydt, H.; Eisenbarth, P.; Feith, K.; Regitz, M. Diaminopyridazine aus diaminocyclopropenyliumsalzen und (diazomethyl)- verbindungen. J. Heterocycl. Chem. 1986, 23, 385–391. [Google Scholar] [CrossRef]
  108. Banert, K.; Kohler, F.; Melzer, A.; Scharf, I.; Rheinwald, G.; Ruffer, T.; Lang, H.; Herges, R.; Hess, K.; Ghavtadze, N.; et al. Experimental and theoretical characterization of the aromatization, epimerization, and fragmentation reactions of bi-2H-azirin-2-yls prepared from 1,4-diazidobuta-1,3-dienes. Chemistry 2011, 17, 10071–10080. [Google Scholar] [CrossRef]
  109. Collomb, D.; Deshayes, C.; Doutheau, A. Synthesis of functionalized phenolic derivatives via the benzannulation of dienylketenes formed by a thermal wolff rearrangement of α-diazo-β-keto compounds. Tetrahedron 1996, 52, 6665–6684. [Google Scholar] [CrossRef]
  110. Touil, S.; Zantour, H. Action des hydrazines sur les γ,ß′-dicarbonylphosphonates et phosphineoxides: Synthese de 4-phosphopyridazines et pyridazin-3-ones. Phosphorus Sulfur Silicon Relat. Elem. 1998, 134, 493–502. [Google Scholar] [CrossRef]
  111. Mrowca, J.J. Phosphine Oxide-Substituted Pyrimidines. U.S. Patent 19790049481, 18 November 1980. [Google Scholar]
  112. Kosolapoff, G.M.; Roy, C.H. Synthesis of some pyrimidylphosphonates. J. Org. Chem. 1961, 26, 1895–1898. [Google Scholar] [CrossRef]
  113. Haas, A.; Lieb, M. Darstellung und eigenschaften trifluormethylmercapto-substituierter pyrimidine. J. Heterocycl. Chem. 1986, 23, 1079–1084. [Google Scholar] [CrossRef]
  114. Günther, O.; Hartke, K. Heterocyclische o-amino-phosphonester. Arch. Der Pharm. 1975, 308, 693–700. [Google Scholar] [CrossRef]
  115. Aboujaoude, E.E.; Collignon, N.; Savignac, P. Dialkyl formyl-1 methylphosphonates α-fonctionnels—II1. Tetrahedron 1985, 41, 427–433. [Google Scholar] [CrossRef]
  116. Aboujaoude, E.E.; Collignon, N.; Savignac, P. Synthèse d’hétérocycles α-phosphoniques. nouveaux développements. Phosphorous Sulfur Relat. Elem. 1987, 31, 231–243. [Google Scholar] [CrossRef]
  117. Neidlein, R.; Eichinger, T. [(1,3-Dioxolan-2-yliden)methyl]phosphonate und -phosphinate als (einfache) Synthone in der Heterocyclensynthese. Helv. Chim. Acta 1992, 75, 124–136. [Google Scholar] [CrossRef]
  118. Lu, R.; Yang, H. A novel approach to phosphonyl-substituted heterocyclic system(I). Tetrahedron Lett. 1997, 38, 5201–5204. [Google Scholar] [CrossRef]
  119. Penz, G.; Zbiral, E. 3-Oxo-2-tosyloxy-1-alkenylphosphonsauredialkylester als synthone zum aufbau von hetarylphosphonsauredialkylestern. Mon. Fur Chem. Chem. Mon. 1985, 116, 1041–1049. [Google Scholar] [CrossRef]
  120. Maruyama, T.; Taira, Z.; Horikawa, M.; Sato, Y.; Honjo, M. Synthesis of uracil-5- and adenine-8-phosphonic acids. Tetrahedron Lett. 1983, 24, 2571–2574. [Google Scholar] [CrossRef]
  121. Zhu, L.; Yu, H.; Guo, Q.; Chen, Q.; Xu, Z.; Wang, R. C-H bonds phosphorylation of ketene dithioacetals. Org. Lett. 2015, 17, 1978–1981. [Google Scholar] [CrossRef]
  122. Essid, I.; Lahbib, K.; Kaminsky, W.; Ben Nasr, C.; Touil, S. 5-Phosphonato-3,4-dihydropyrimidin-2(1H)-ones: Zinc triflate-catalyzed one-pot multi-component synthesis, X-ray crystal structure and anti-inflammatory activity. J. Mol. Struct. 2017, 1142, 130–138. [Google Scholar] [CrossRef]
  123. Essid, I.; Touil, S. β-Ketophosphonates as substrates in the Biginelli multicomponent reaction: An efficient and straightforward synthesis of phosphorylated dihydropyrimidinones. Arkivoc 2013, 4, 98–106. [Google Scholar] [CrossRef]
  124. Timoshenko, V.M.; Markitanov, Y.N.; Shermolovich, Y.G. Preparation of pyrimidine derivatives through three-component reactions of dialkyl(2-oxo-3,3,3-trifluoropropyl)phosphonates. Chem. Heterocycl. Compd. 2011, 47, 977–982. [Google Scholar] [CrossRef]
  125. Jansa, P.; Hradil, O.; Baszczyňski, O.; Dračínský, M.; Klepetářová, B.; Holý, A.; Balzarini, J.; Janeba, Z. An efficient microwave-assisted synthesis and biological properties of polysubstituted pyrimidinyl- and 1,3,5-triazinylphosphonic acids. Tetrahedron 2012, 68, 865–871. [Google Scholar] [CrossRef] [Green Version]
  126. Kunda, U.M.R.; Mudumala, V.N.R.; Reddy Gangireddy, C.S.; Nemallapudi, B.R.; Sandip, K.N.; Cirandur, S.R. Amberlyst-15 catalyzed synthesis of alkyl/aryl/heterocyclic phosphonates. Chin. Chem. Lett. 2011, 22, 895–898. [Google Scholar] [CrossRef]
  127. Mohan Naidu, K.R.; Dadapeer, E.; Reddy, C.B.; Rao, A.J.; Reddy, C.S.; Raju, C.N. Polyethylene glycol–promoted dialkyl, aryl/heteroaryl phosphonates. Synth. Commun. 2011, 41, 3462–3468. [Google Scholar] [CrossRef]
  128. Golla, M.; Syed, R.; Katla, V.R.; Devineni, S.R.; Kondapalli, N.; Chamarthi, N.R. LaCl3·7H2O: An efficient catalyst for the synthesis of phosphinates (Michaelis–Arbuzov reaction) under neat conditions and their potential antimicrobial activity. J. Chem. Sci. 2014, 126, 117–125. [Google Scholar] [CrossRef] [Green Version]
  129. Varalakshmi, M.; Srinivasulu, D.; Kotakadi, V.S. Nano-BF3*SiO2 catalyst-promoted Michaelis-Arbuzov reaction: Solvent-free synthesis and antimicrobial evaluation. Phosphorus Sulfur Silicon Relat. Elem. 2015, 190, 1518–1524. [Google Scholar] [CrossRef]
  130. Ghosh, I.; Shaikh, R.S.; Konig, B. Sensitization-initiated electron transfer for photoredox catalysis. Angew. Chem. Int. Ed. 2017, 56, 8544–8549. [Google Scholar] [CrossRef] [Green Version]
  131. Erbland, G.; Ruch, J.; Goddard, J.-P. Photochemical functionalization of diazines: Metal-free vinylation and phosphonylation. Tetrahedron 2016, 72, 7826–7831. [Google Scholar] [CrossRef]
  132. Zhang, W.; Zou, J.-P.; Sun, W.-B.; Ji, Y.-F.; Pan, X.-Q.; Zhou, S.-F.; Asekun, O. Mn(OAc)3-mediated selective free radical phosphonylation of pyridinones and pyrimidinones. Synthesis 2013, 45, 1529–1533. [Google Scholar]
  133. Deal, E.L.; Petit, C.; Montchamp, J.L. Palladium-catalyzed cross-coupling of H-phosphinate esters with chloroarenes. Org. Lett. 2011, 13, 3270–3273. [Google Scholar] [CrossRef] [PubMed]
  134. Zhang, H.Y.; Sun, M.; Ma, Y.N.; Tian, Q.P.; Yang, S.D. Nickel-catalyzed C-P cross-coupling of diphenylphosphine oxide with aryl chlorides. Org. Biomol. Chem. 2012, 10, 9627–9633. [Google Scholar] [CrossRef]
  135. Montel, S.; Midrier, C.; Volle, J.-N.; Braun, R.; Haaf, K.; Willms, L.; Pirat, J.-L.; Virieux, D. Functionalized phosphanyl-phosphonic acids as unusual complexing units as analogues of fosmidomycin. Eur. J. Org. Chem. 2012, 2012, 3237–3248. [Google Scholar] [CrossRef]
  136. Zakirova, G.; Mladentsev, D.; Borisova, N. Palladium-catalyzed C–P cross-coupling between (het)aryl halides and secondary phosphine oxides. Synthesis 2019, 51, 2379–2386. [Google Scholar] [CrossRef]
  137. Bai, Y.; Liu, N.; Wang, S.; Wang, S.; Ning, S.; Shi, L.; Cui, L.; Zhang, Z.; Xiang, J. Nickel-catalyzed electrochemical phosphorylation of aryl bromides. Org. Lett. 2019, 21, 6835–6838. [Google Scholar] [CrossRef]
  138. Palacios, F.; Ochoa de Retana, A.M.; Gil, J.I.; Lopez de Munain, R. Synthesis of pyrazine-phosphonates and -phosphine oxides from 2H-azirines or oximes. Org. Lett. 2002, 4, 2405–2408. [Google Scholar] [CrossRef]
  139. Palacios, F.; de Retana, A.M.a.O.; de Marigorta, E.M.n.; Rodriguez, M.; Pagalday, J. Aza-Wittig reaction of N-phosphorylalkyl phosphazenes with carbonyl compounds and phenylisocyanate. Synthesis of 4-amino-3-phosphoryl-2-azadienes and pyrazine-phosphonates. Tetrahedron 2003, 59, 2617–2623. [Google Scholar] [CrossRef]
  140. Aparicio, D.; Attanasi, O.A.; Filippone, P.; Ignacio, R.; Lillini, S.; Mantellini, F.; Palacios, F.; de Los Santos, J.M. Straightforward access to pyrazines, piperazinones, and quinoxalines by reactions of 1,2-diaza-1,3-butadienes with 1,2-diamines under solution, solvent-free, or solid-phase conditions. J. Org. Chem. 2006, 71, 5897–5905. [Google Scholar] [CrossRef]
  141. Seggio, A.; Chevallier, F.; Vaultier, M.; Mongin, F. Lithium-mediated zincation of pyrazine, pyridazine, pyrimidine, and quinoxaline. J. Org. Chem. 2007, 72, 6602–6605. [Google Scholar] [CrossRef]
  142. Berger, O.; Montchamp, J.L. Manganese-catalyzed and mediated synthesis of arylphosphinates and related compounds. J. Org. Chem. 2019, 84, 9239–9256. [Google Scholar] [CrossRef] [PubMed]
  143. Nikishkin, N.I.; Huskens, J.; Ansari, S.A.; Mohapatra, P.K.; Verboom, W. Pyrazine-functionalized calix[4]arenes: Synthesis by palladium-catalyzed cross-coupling with phosphorus pronucleophiles and metal ion extraction properties. New J. Chem. 2013, 37, 391–402. [Google Scholar] [CrossRef]
  144. Zhao, Y.L.; Wu, G.J.; Li, Y.; Gao, L.X.; Han, F.S. [NiCl2(dppp)]-catalyzed cross-coupling of aryl halides with dialkyl phosphite, diphenylphosphine oxide, and diphenylphosphine. Chemistry 2012, 18, 9622–9627. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Approaches to synthesis of POR2-containing pyridines.
Figure 1. Approaches to synthesis of POR2-containing pyridines.
Molecules 28 02472 g001
Scheme 1. Synthesis of pyridine-2-phosphonates.
Scheme 1. Synthesis of pyridine-2-phosphonates.
Molecules 28 02472 sch001
Scheme 2. Reaction of diethyl (2-oxobutyl)phosphonate with Mannich bases.
Scheme 2. Reaction of diethyl (2-oxobutyl)phosphonate with Mannich bases.
Molecules 28 02472 sch002
Scheme 3. Rh(II)-catalyzed cyclization of δ-diazo oxime ether.
Scheme 3. Rh(II)-catalyzed cyclization of δ-diazo oxime ether.
Molecules 28 02472 sch003
Scheme 4. Reaction of phosphorylated carbenoid with 2H-azirine.
Scheme 4. Reaction of phosphorylated carbenoid with 2H-azirine.
Molecules 28 02472 sch004
Scheme 5. Formal [2+2+2]-cycloaddition of 1,6- and 1,7-diynes with diethyl phosphorocyanidate.
Scheme 5. Formal [2+2+2]-cycloaddition of 1,6- and 1,7-diynes with diethyl phosphorocyanidate.
Molecules 28 02472 sch005
Scheme 6. Photocatalytic phosphorylation of halopyridines: (a) Synthesis of 2-phosphine oxide-substituted pyridines; (b) Synthesis of 3-phosphine oxide-substituted pyridines.
Scheme 6. Photocatalytic phosphorylation of halopyridines: (a) Synthesis of 2-phosphine oxide-substituted pyridines; (b) Synthesis of 3-phosphine oxide-substituted pyridines.
Molecules 28 02472 sch006
Scheme 7. Synthesis of pyridine-3- and pyridine-4-phosphonates from pyridylazo sulfones.
Scheme 7. Synthesis of pyridine-3- and pyridine-4-phosphonates from pyridylazo sulfones.
Molecules 28 02472 sch007
Scheme 8. Photocatalytic reaction of N-ethoxypyridinium salts with phosphine oxides.
Scheme 8. Photocatalytic reaction of N-ethoxypyridinium salts with phosphine oxides.
Molecules 28 02472 sch008
Scheme 9. Oxidative coupling of pyridines with dialkyl phosphites in the presence of Ag(I).
Scheme 9. Oxidative coupling of pyridines with dialkyl phosphites in the presence of Ag(I).
Molecules 28 02472 sch009
Scheme 10. Coupling of pyridines with secondary phosphine oxides in the presence of acetylenes.
Scheme 10. Coupling of pyridines with secondary phosphine oxides in the presence of acetylenes.
Molecules 28 02472 sch010
Scheme 11. Reaction between N-methoxypyridinium tosylates and secondary phosphines.
Scheme 11. Reaction between N-methoxypyridinium tosylates and secondary phosphines.
Molecules 28 02472 sch011
Scheme 12. Reaction of pyridine N-oxides with triethyl phosphite.
Scheme 12. Reaction of pyridine N-oxides with triethyl phosphite.
Molecules 28 02472 sch012
Scheme 13. Asymmetric synthesis of tertiary pyridine-containing phosphine oxides.
Scheme 13. Asymmetric synthesis of tertiary pyridine-containing phosphine oxides.
Molecules 28 02472 sch013
Scheme 14. Synthesis of chiral tertiary pyridine-containing phosphine oxides.
Scheme 14. Synthesis of chiral tertiary pyridine-containing phosphine oxides.
Molecules 28 02472 sch014
Scheme 15. Cross-coupling of halopyridines with dimethylphosphine oxide.
Scheme 15. Cross-coupling of halopyridines with dimethylphosphine oxide.
Molecules 28 02472 sch015
Scheme 16. Pd(II)-catalyzed cross-coupling of 3-bromopyridines with triethyl phosphite.
Scheme 16. Pd(II)-catalyzed cross-coupling of 3-bromopyridines with triethyl phosphite.
Molecules 28 02472 sch016
Scheme 17. Cross-coupling of pyridine carboxylic acids with dialkyl phosphites.
Scheme 17. Cross-coupling of pyridine carboxylic acids with dialkyl phosphites.
Molecules 28 02472 sch017
Scheme 18. Replacement of the OH group of hydroxypyridines mediated by sulfuryl fluoride.
Scheme 18. Replacement of the OH group of hydroxypyridines mediated by sulfuryl fluoride.
Molecules 28 02472 sch018
Scheme 19. Replacement of the OH group of hydroxypyridines by the diphenylphosphoryl moiety.
Scheme 19. Replacement of the OH group of hydroxypyridines by the diphenylphosphoryl moiety.
Molecules 28 02472 sch019
Scheme 20. Cross-coupling of aryl carboxylic acid esters with phosphine oxides.
Scheme 20. Cross-coupling of aryl carboxylic acid esters with phosphine oxides.
Molecules 28 02472 sch020
Scheme 21. Ni(II)-catalyzed replacement of the tosyl group in pyridines.
Scheme 21. Ni(II)-catalyzed replacement of the tosyl group in pyridines.
Molecules 28 02472 sch021
Figure 2. Approaches to synthesis of POR2-containing pyridazines.
Figure 2. Approaches to synthesis of POR2-containing pyridazines.
Molecules 28 02472 g002
Scheme 22. Reactions of cyclopropenylium salts with phosphorus-containing diazomethanes.
Scheme 22. Reactions of cyclopropenylium salts with phosphorus-containing diazomethanes.
Molecules 28 02472 sch022
Scheme 23. Rearrangement of phosphorus-containing bis-azirine.
Scheme 23. Rearrangement of phosphorus-containing bis-azirine.
Molecules 28 02472 sch023
Scheme 24. Cyclization of γ,δ-unsaturated α-diazo-β-keto phosphonate.
Scheme 24. Cyclization of γ,δ-unsaturated α-diazo-β-keto phosphonate.
Molecules 28 02472 sch024
Scheme 25. Synthesis of dimethyl 3,6-diphenylpyridazine-3-phosphonate.
Scheme 25. Synthesis of dimethyl 3,6-diphenylpyridazine-3-phosphonate.
Molecules 28 02472 sch025
Scheme 26. Phosphorylation of pyridazines: (a) Mrowca et al., 1980 [111]; (b) Kim et al., 2019 [79].
Scheme 26. Phosphorylation of pyridazines: (a) Mrowca et al., 1980 [111]; (b) Kim et al., 2019 [79].
Molecules 28 02472 sch026
Figure 3. Approaches to synthesis of POR2-containing pyrimidines.
Figure 3. Approaches to synthesis of POR2-containing pyrimidines.
Molecules 28 02472 g003
Scheme 27. Cyclization reactions of guanidine (a) and amidine (b).
Scheme 27. Cyclization reactions of guanidine (a) and amidine (b).
Molecules 28 02472 sch027
Scheme 28. Michaelis–Arbuzov reaction in synthesis of pyrimidine phosphonates: (a) Jansa et al., 2012 [125]; (b) Kunda et al., 2011 [126]; Mohan Naidu et al., 2011 [127]; (c) Golla et al., 2014 [128]; (d) Varalakshmi et al., 2015 [129].
Scheme 28. Michaelis–Arbuzov reaction in synthesis of pyrimidine phosphonates: (a) Jansa et al., 2012 [125]; (b) Kunda et al., 2011 [126]; Mohan Naidu et al., 2011 [127]; (c) Golla et al., 2014 [128]; (d) Varalakshmi et al., 2015 [129].
Molecules 28 02472 sch028
Scheme 29. Photo-Arbuzov reaction in synthesis of pyrimidine-5-phosphonates.
Scheme 29. Photo-Arbuzov reaction in synthesis of pyrimidine-5-phosphonates.
Molecules 28 02472 sch029
Scheme 30. Photocatalytic phosphorylation of 2-chloropyrimidines.
Scheme 30. Photocatalytic phosphorylation of 2-chloropyrimidines.
Molecules 28 02472 sch030
Scheme 31. Pd2(dba)3-catalyzed cross-coupling of halopyrimidines with dimethylphosphine oxide.
Scheme 31. Pd2(dba)3-catalyzed cross-coupling of halopyrimidines with dimethylphosphine oxide.
Molecules 28 02472 sch031
Scheme 32. Transition-metal-catalyzed cross-coupling of 2-chloropyrimidines with phosphine oxides.
Scheme 32. Transition-metal-catalyzed cross-coupling of 2-chloropyrimidines with phosphine oxides.
Molecules 28 02472 sch032
Scheme 33. NiBr2-catalyzed electrochemical synthesis of 5-phosphorylpyrimidine.
Scheme 33. NiBr2-catalyzed electrochemical synthesis of 5-phosphorylpyrimidine.
Molecules 28 02472 sch033
Figure 4. Approaches to synthesis of POR2-containing pyrazines.
Figure 4. Approaches to synthesis of POR2-containing pyrazines.
Molecules 28 02472 g004
Scheme 34. Synthesis of phosphorus-containing pyrazines via dimerization of nitrile ylides.
Scheme 34. Synthesis of phosphorus-containing pyrazines via dimerization of nitrile ylides.
Molecules 28 02472 sch034
Scheme 35. Dimerization of 4-dimethylamino-3-phosphoryl-2-azadiene.
Scheme 35. Dimerization of 4-dimethylamino-3-phosphoryl-2-azadiene.
Molecules 28 02472 sch035
Scheme 36. Synthesis of 2-POR2-substituted pyrazines from 1,2-diaza-1,3-butadienes and 1,2-diamines.
Scheme 36. Synthesis of 2-POR2-substituted pyrazines from 1,2-diaza-1,3-butadienes and 1,2-diamines.
Molecules 28 02472 sch036
Scheme 37. Synthesis of 2-phosphorylpyrazine by the Arbuzov reaction: (a) Golla et al., 2014 [128]; (b) Goddard et al., 2016 [131].
Scheme 37. Synthesis of 2-phosphorylpyrazine by the Arbuzov reaction: (a) Golla et al., 2014 [128]; (b) Goddard et al., 2016 [131].
Molecules 28 02472 sch037
Scheme 38. Synthesis of 2-POR2-substituted pyrazines by radical phosphorylation: (a) Yuan et al., 2018 [73]; (b) Berger and Montchamp, 2019 [142].
Scheme 38. Synthesis of 2-POR2-substituted pyrazines by radical phosphorylation: (a) Yuan et al., 2018 [73]; (b) Berger and Montchamp, 2019 [142].
Molecules 28 02472 sch038
Scheme 39. Synthesis of 2-phosphorylpyrazines under transition metal catalysis.
Scheme 39. Synthesis of 2-phosphorylpyrazines under transition metal catalysis.
Molecules 28 02472 sch039
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

Volkova, Y.; Zavarzin, I. Synthesis of Phosphorus(V)-Substituted Six-Membered N-Heterocycles: Recent Progress and Challenges. Molecules 2023, 28, 2472. https://doi.org/10.3390/molecules28062472

AMA Style

Volkova Y, Zavarzin I. Synthesis of Phosphorus(V)-Substituted Six-Membered N-Heterocycles: Recent Progress and Challenges. Molecules. 2023; 28(6):2472. https://doi.org/10.3390/molecules28062472

Chicago/Turabian Style

Volkova, Yulia, and Igor Zavarzin. 2023. "Synthesis of Phosphorus(V)-Substituted Six-Membered N-Heterocycles: Recent Progress and Challenges" Molecules 28, no. 6: 2472. https://doi.org/10.3390/molecules28062472

Article Metrics

Back to TopTop