Next Article in Journal
The Effect of Hemp Cake (Cannabis sativa L.) on the Characteristics of Meatballs Stored in Refrigerated Conditions
Next Article in Special Issue
Photoinduced Bisphosphination of Alkynes with Phosphorus Interelement Compounds and Its Application to Double-Bond Isomerization
Previous Article in Journal
Chemical Physical Characterization and Profile of Fruit Volatile Compounds from Different Accesses of Myrciaria floribunda (H. West Ex Wild.) O. Berg through Polyacrylate Fiber
Previous Article in Special Issue
Synthesis and Antiproliferative Activity of Phosphorus Substituted 4-Cyanooxazolines, 2-Aminocyanooxazolines, 2-Iminocyanooxazolidines and 2-Aminocyanothiazolines by Rearrangement of Cyanoaziridines
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Nickel Complexes in C‒P Bond Formation

by
Almaz A. Zagidullin
,
Il’yas F. Sakhapov
,
Vasili A. Miluykov
and
Dmitry G. Yakhvarov
*
Arbuzov Institute of Organic and Physical Chemistry, FRC Kazan Scientific Center, Russian Academy of Sciences, Arbuzov Str. 8, 420088 Kazan, Russia
*
Author to whom correspondence should be addressed.
Submission received: 26 July 2021 / Revised: 23 August 2021 / Accepted: 26 August 2021 / Published: 31 August 2021
(This article belongs to the Special Issue Modern Organophosphorus Chemistry)

Abstract

:
This review is a comprehensive account of reactions with the participation of nickel complexes that result in the formation of carbon–phosphorus (C‒P) bonds. The catalytic and non-catalytic reactions with the participation of nickel complexes as the catalysts and the reagents are described. The various classes of starting compounds and the products formed are discussed individually. The several putative mechanisms of the nickel catalysed reactions are also included, thereby providing insights into both the synthetic and the mechanistic aspects of this phosphorus chemistry.

1. Introduction

Taking into account the use of organophosphorus compounds in organic synthesis and homogeneous catalysis, materials chemistry, agrochemical crop protection and pharmaceutical discovery, new methods for their synthesis hold particular significance [1,2,3,4,5]. Traditional methods to construct carbon–phosphorus (C−P) bonds—a key step in the synthesis of organophosphorus compounds—such as radical methods, anionic methods and the Arbuzov, Pudovik, Michaelis, Kabachnik–Fields, Abramov reactions are well known. These classical methodologies suffer from safety problems and limited scope, lack of selectivity and the use of protective groups that lead to additional stages of synthesis. Therefore, there has been growing interest in the development of transition-metal-catalyzed C−P bond construction as one of the most attractive method due to the safety, selectivity, high functional group tolerance and 100% atom economy provided by this approach [6,7,8,9].
Among the wide range of transition metals used for C−P bond construction, our attention has focused on nickel complexes. The development of organonickel chemistry has led to the discovery of several remarkable catalytic systems with excellent practical applications [10,11,12]. Actually, homogeneous Ni catalysis is currently experiencing a period of growing interest, resulting in numerous fascinating applications in synthetic organic and organophosphorus chemistry. Catalytic cycles with organonickel complexes as intermediates in many cases demonstrate high efficiency and include non-reactive organic and phosphorus substrates, but it is difficult to predict and control all reaction pathways [13]. From another point of view, cost-effectiveness is the undisputed driving force and great advantage behind the choice of nickel for catalytic applications, because comparing the cost of catalyst precursors for nickel and other noble metals shows a dramatic difference.
This review provides an overview of the nickel-catalyzed synthesis of phosphanes, phosphonium salts, phosphane oxides and phosphorus acid derivatives. This review focuses on the latest advances in applications of nickel complexes as an effective catalyst in C‒P bond formation, some aspects of the reaction mechanism and important advances in the asymmetric synthesis of organophosphorus compounds.

2. Synthesis of Tricoordinated Organophosphorus Compounds

2.1. Synthesis of Phosphanes Using Phosphane Chlorides

Phosphane chlorides are a useful class of P-coupling partners because the commonly used secondary phosphanes, their oxides or borane complexes are pyrophoric, require an additional reduction step or are incompatible with other functional groups. Cristau et al. reported initial efforts in the Ni-catalyzed cross-coupling of diphenylphosphane chloride with arylbromides [14]. Since the reaction yielded a mixture of phosphonium salts and triphenylphosphane oxides after workup, it has been evaluated that it is not synthetically useful.
An interesting version of cross-coupling with diphenylphosphane chloride catalyzed by NiCl2(dppe) (dppe—1,2-bis(diphenylphosphano)ethane) in the presence of stoichiometric amounts of metallic Zn was reported in [15] (Figure 1). In this reaction, zinc performs two functions: it reduces Ni(II) to Ni(0) and gives rise to zinc phosphide Ph2PZnCl, which reacts with ArylX. This methodology represents a convenient procedure for the preparation of different tertiary phosphanes, including the coupling of sterically hindered aryl halides or sulfonates that contain ortho-substituents as well as amide groups in the substrates. It should be noted that aryltrifluoromethanesulfonates provide higher yields (46–95%) than similar bromides.
Later, using this methodology, functionalized triarylphosphanes were obtained with good yields (55–86%) in a one-step reaction of an equimolar mixture of chlorodiphenylphosphane and aromatic bromides in NMP or DMF at 110 °C in the presence of zinc dust as a cheap reductant and NiBr2(bpy) (bpy—2,2′-bipyridine) as an efficient catalyst [16] (Figure 1). The main features of this versatile method are the simplicity of the reaction conditions and the compatibility with various functional groups.
Since then, (trimethylsilyl)diphenylphosphane was also employed as the phosphorus coupling partner. The optimized conditions for the Ni-catalyzed C–P cross-coupling reaction included NiCl2(PPh3)2 as the catalyst, tBuOK as the base and Me3SiPPh2 as the phosphane reagent at 90 °C [17] (Figure 2).
Nickel- as well as palladium- and copper-catalyzed couplings of terminal alkynes with chlorophosphanes were developed later [18,19]. The Ni-catalyzed coupling of diarylchloro-, dialkylchloro-, aryldichloro- and trichlorophosphane (PCl3) with terminal acetylenes is a smooth transformation leading to a corresponding C-P coupling product in high yield (Figure 3). It should be noted that the couplings of aryldichlorophosphanes and trichlorophosphane (PCl3) are not selective, resulting in a mixture of mono-, di- and trisubstituted products.
From the mechanistic point of view, the authors claim the reaction to be a heteroanalogue of the Sonogashira cross-coupling [20]. The initial step of the catalytic cycle is the oxidative addition of the chlorophosphane to the Ni complex, forming a Ni-phosphido complex. Subsequent ligand exchange from chloride to acetylide gives a Ni-organic sigma complex, which then liberates phosphanoalkyne after reductive elimination (Figure 4).

2.2. Arylation and Vinylation of Secondary and Primary Phosphanes

The first example of secondary phosphanes used in transition metal-catalyzed cross-coupling reactions was demonstrated by Cristau et al., who explored the arylation of diphenylphosphane Ph2PH by aryl bromides in the presence of NiBr2 salt [14]. Upon the reaction of bromobenzene with diphenylphosphane in the presence of NiBr2, a mixture of triphenylphosphane (31%) and tetraphenylphosphonium bromide (19%) salt was obtained (Figure 5).
Later, Shulyupin et al. further broadened the synthetic applicability by employing (Ph3P)2NiCl2 or Ni(acac)2 (acac—acetylacetone) as efficient catalyst precursors in the phosphination of vinyl bromides and chlorides with diphenylphosphane (Figure 6) [21]. The procedure uses a combination of up to 1 mol% of Ni complexes, triethylamine and DMF as a solvent, leading to products with 75–96% yields. The double bond geometry of the vinyl halides was retained under the reaction conditions.
Functionalized vinylphosphanes containing alkoxy- or amino- groups were synthesized by the reaction of diphenylphosphane or its trimethylsilyl derivative with the corresponding alkenyl bromides or chlorides under catalysis by nickel complexes. Ni complexes turned out to be more efficient than Pd complexes in reactions with less active alkenyl halides, such as 2-bromobutene and 1-bromovinylsilane [22]. Likewise, two diphenylphosphano groups were introduced into 1 and 4 positions of 1,4-diiodo-butadienes with 90–95% yields [23] (Figure 7).
A striking example of a Ni-catalyzed cross-coupling was developed by Cai et al. for the synthesis of BINAP [24], one of the most efficient and successful chiral ligands, which was synthesized for the first time by the Noyori group in the early 1980s. These authors anticipated the use of a Ni catalyst, instead of a Pd catalyst, to be more promising since the former binds to BINAP weaker than any other transition metal of the second or third row. Thus, catalyst poisoning was prevented. After the initial optimization of the reaction conditions, BINAP was synthesized without racemization. The best yield of 77% was obtained by NiCl2(dppe) cross-coupling of aryltriflate and diphenylphosphane in DMF in the presence of DABCO as a base [25], while other systems did not promote the reaction or led to side or oxidation reactions (Figure 8). Fortunately, Cai’s method is not restricted to diarylphosphanes, since dialkylphosphanes were also used in this reaction [26]. In addition, Wills and co-workers reported that, in some cases, the addition of zinc dust improves the yield [27].
Several research groups have adopted the Ni-catalyzed cross-coupling protocol developed by Cai et al. for the synthesis of a wide variety of chiral phosphanes: axially chiral Quinazolinap ligand [28,29], P-stereogenic BINAP [30] and other binaphthyl-based phosphane and phosphite ligands [31]. Figure 9 shows selected examples, such as steroidal [32] and pyrazinylnaphthyl derivatives of BINAP [33], PINAPs [34,35], or fluoroalkyl-tagged binaphthyls [36,37].
Recently, Zhao et al. disclosed a method for the cross-coupling of various aryl bromides with diphenylphosphane in the absence of external reductants and supporting ligands (Figure 10) [38]. The reaction gave a mixture of phosphanes and phosphane oxides with 64–99% yields. Several functional groups (ester, ether, ketone and cyano groups) remained intact under the conditions.
In addition to the above-mentioned methods, an electrochemically promoted nickel catalysed processes were also developed [39,40]. Organonickel sigma-complexes have been found as efficient key intermediates in various Ni-catalyzed C‒C coupling reactions [41], electrocatalytic processes and C‒P bond formation with participation of unfunctionalized organic arylhalides with elemental (white) phosphorus P4 [42], chlorophosphanes or various primary and secondary phosphanes (Figure 11) [43,44].

2.3. Hydrophosphination Reactions of Alkenes and Alkynes

Hydrophosphination reactions involve the addition of P–H to an unsaturated C–C bond and have gained great interest as an alternative to the classical phosphane syntheses involving a substitution that is incompatible with certain functional groups. In this reaction, phosphanes, silylphosphanes or phosphane–borane complexes are used as phosphinating agents to react with inactivated or activated alkenes, dienes and alkynes. Moreover, the addition of P–H to an unsaturated C–C bond is more efficient than substitution reactions when considering atom efficiency that makes it greener and more economical.
In recent years, great progress has been made in metal-complex-catalyzed hydrophosphination reactions [9,45]. It should be noted that reactions of P‒H and P‒E compounds with alkynes in the presence of transition metal complexes occur preferentially as syn-addition. It was shown that hydrophosphination reactions catalyzed by Ni-based complexes proceed more efficiently and allow inactivated alkenes to be employed [46,47,48,49]. Shulupin et al. first reported hydrophosphination with high yields up to 99% of weakly activated aryl olefins and their heterocyclic analogues in the presence of Ni(0) phosphite complexes (Figure 12) [50].
The reaction is regioselective: the only product is the corresponding β-phosphorylated adduct. The fact that no α-adduct is formed in the addition of Ph2PH to styrenes and vinylpyridines allows the formation of π-allyl intermediates to be excluded. A probable catalytic cycle includes the oxidative addition of phosphane to Ni(0) with the formation of a hydride phosphide complex, alkene insertion into the Ni‒H bond and subsequent reductive elimination (Figure 13). Later, Ganushevich et al. characterized, for the first time, the product of an oxidative addition of primary phosphane to a nickel(0) complex. The terminal phosphanido hydride nickel complex, [NiH{P(Dmp)(H)}(dtbpe)], where Dmp—2,6-dimesitylphenyl and dtbpe—1,2-bis(di-tert-butylphosphino)ethane, has been formed in this process [51,52,53].
Chiral metal complexes have been used to promote and control the asymmetric P–H addition reaction. A chiral pincer bisphosphane Ni complex was used in the first highly enantioselective catalytic synthesis of P-stereogenic secondary phosphane–boranes by the asymmetric hydrophosphination of electron-deficient alkenes with phenylphosphane. Various P-stereogenic secondary phosphane–boranes were obtained in 57–92% yields with up to 99% ee [54]. The Togni group developed the asymmetric hydrophosphination of vinyl nitriles catalyzed by a dicationic Ni-based complex yielding the desired phosphane product with 32–94% ee [55,56] (Figure 14).
The addition of P–H to a triple bond is a highly desirable method when considering the principles of atom economy. The first example of the hydrophosphination of terminal and internal alkynes, catalyzed by Pd and Ni complexes, was reported by Kazankova et al. (Figure 15) [57]. The regioselectivity was strongly dependent on the catalytic precursor and alkyne nature. In the presence of Pd(0) and Ni(0) complexes, the β-adduct was formed as the major product. By contrast, Pd(II) and Ni(II) complexes mainly gave rise to the α-adduct (α:β = 95:5) [58]. The different selectivity in the reactions catalyzed by Pd(0)/Ni(0) complexes and Pd(Ni)X2 is explained by the formation of catalytic amounts of HX (HOAc or HBr) in situ, which initiate the second catalytic cycle. The Ni-based catalyst was more effective than the Pd-based catalyst, and the reaction proceeded at lower temperatures. The relative reactivity of the metal complexes in the hydrophosphination of alkynes was studied theoretically by Ananikov et al., and it decreased in the order of Ni > Pd > Rh > Pt. [59]. The estimated relative reactivity order of the studied metals implies that nickel can not only be a cost-economic replacement of Pd, but also superior in terms of catalytic efficiency. In the reaction of diphenylphosphane with tert-butylacetylene, the corresponding β-adduct with 95% yield is formed as the only product for steric reasons. The addition of diphenylphosphane to other alkylacetylenes is characterized by lower selectivity, with both regio- and stereoselectivity strongly dependent on the reaction conditions.
Later, an efficient NCC pincer Ni(II)-catalyzed hydrophosphination of nitroalkenes with HPPH2 was developed. After the optimization of reaction conditions, (hetero)aromatic and aliphatic nitroalkenes were well tolerated, irrespective of electronic effects, to provide the products in up to 99% yield [60]. In addition to P(III) phosphanes, Montchamp and co-workers produced vinyl-H-phosphinates from alkynes and alkyl phosphinates using only 2–3 mol% NiCl2 [61]. Ananikov et al. have shown the Markovnikov-selective phosphorylation of internal and terminal alkynes using a range of phosphites in the presence of catalytic amounts of Ni(acac)2 and 1,2-bis(diphenylphosphano)ethane (dppe) [62]. Han et al. reported that nickel catalysts are more reactive than noble metal catalysts in the catalytic additions of a variety of P(O)−H bonds to alkynes (propargyl alcohols, 1-octyne), regioselectively affording both the Markovnikov and the anti-Markovnikov products in high yields (72–96%) [63]. A related five-coordinated hydrido nickel complex is successfully isolated in the catalysis, which can react readily with an alkyne to give the addition products (Figure 16) [64].

3. Synthesis of Tetracoordinated Organophosphorus Compounds

3.1. Synthesis of Phosphonium Salts in the Presence of Ni Salts

The reaction of aryl halides with PPh3 in the presence of Ni(II) salts is one of the oldest transition metal-catalyzed reactions known to form a C‒P bond (Figure 17). Iodobenzene worked best in the reaction (yield up to 90%), while the corresponding bromides and chlorides were less reactive and gave lower yields (60–90%). Electron-donating alkyl, amino or alkoxy groups facilitate the C‒P cross-coupling, while electron-withdrawing groups act as strong inhibitors [65]. The reaction is quite common for para- and meta-substituted aryl halides. Although ortho-substituted aryl halides are usually transferred due to the stabilization of the intermediately formed Csp2–Ni bond, [65] Allen et al. used the directing effect of an imine or a diazo nitrogen atom to perform a chemoselective cross-coupling reaction in the ortho-position under mild conditions [66].
Thiophene [67], furan [66] and pyrrole [68] halides also served as the substrates or as substituents at the phosphorus atom. In the case of trialkylphosphanes, high temperatures might be problematic for these sensitive substances, so further cross-coupling reactions were successfully performed in refluxing ethanol using (Ph3P)3Ni [65,69].
Later, the Charettea group reported a general and efficient Ni-catalyzed cross-coupling reaction between aryl halides (iodides, bromides, chlorides) or triflates and PPh3 generating tetraarylphosphonium salts in good to high yields (63–99%) (Figure 18). This Ni-catalyzed C‒P coupling is conducted in ethylene glycol using a readily available, cheap NiBr2 precatalyst and tolerates different functional groups such as alcohols, amides, ketones, aldehydes, phenols and amines [70].
Additionally, nickel-catalyzed C–P coupling polymerization of commercial aryl dihalides and diphenylphosphane was used for the convenient preparation of tetraarylphosphonium polyelectrolytes. A NiBr2-based catalyst was effective in C–P coupling reactions to yield tetraarylphosphonium polymers with degrees of polymerization up to about 30 [71].

3.2. Synthesis of Phosphane Oxides by C‒P Cross-Coupling

Transition metal-catalyzed C–P bond formation has been well explored [72]. In the last decade, there have been reports of the use of nickel-based catalysts for the synthesis of phosphane oxides. Yang and co-workers reported the synthesis of diphenylphosphoryl compounds through Ni-catalyzed cross-coupling of diphenylphosphane oxide Ph2P(O)H with heteroaryl chlorides (Figure 19) [73]. The reactions of various aryl halides with diphenylphosphane oxide were also carried out using the Ni(Zn) catalyst together with N-ligands in water, leading to the formation of diphenylarylphosphane oxides with 75–97% yields [74].
The Zhao group reported the Ni-catalyzed cross-coupling of functionalized arylboronic acids with H-phosphites, H-phosphane oxides and H-phosphinate esters to give various organophosphorus compounds with good to excellent yields (50–99%) (Figure 20) [75]. This strategy provided a generalized and substantial tool for the synthesis of triarylphosphane oxides and was the first example of a Ni-catalyzed C–P cross-coupling reaction utilizing >P(O)H substrates and arylboronic acids.
Later, Liu et al. demonstrated the reaction of 1,1-dibromo-1-alkenes with diphenylphosphane oxide using NiBr2(bpy) and magnesium in the presence of potassium phosphate at moderate temperatures (Figure 21) [76]. Mechanistic studies show that the reaction involves the «Hirao type» reduction to give alkenyl bromides, which undergo a reaction with diphenylphosphane oxide in the presence of Ni(0), which can be obtained from NiBr2 by reduction using magnesium.
The Xiao group reported the coupling of aryl iodides with diphenyl phosphane oxide at room temperature on the base of nickel and photoredox-based catalytic systems [77]. Photoredox catalysis using Ru(bpy)3Cl2(H2O)6 with 3W blue LED gives the P-centered radical, which reacts with the nickel(II)–aryl complex. The reaction takes place at room temperature and tolerates phenol, amide and ether functional groups.
Recently, the decarbonylative coupling of aryl esters with diphenylphosphane oxides in the presence of Ni(OAc)2 was demonstrated by Yamaguchi and co-workers (Figure 22) [78]. The key success of the transformation is the use of the 3,4-bis(dicyclohexylphosphano)thiophene (dcypt) ligand, and the yield was modest to good (48–82%).
A Ni-catalyzed asymmetric allylation of secondary phosphane oxides for the synthesis of tertiary phosphane oxides was realized in the Zhang group with high enantioselectivity (87–95% ee) (Figure 23) [79]. The protocol represents the first example of synthesizing P-stereogenic phosphane oxides by an allylation reaction. The finding of this research expands the applications of Ni-based catalysts and secondary phosphane oxides in the synthesis of P-stereogenic phosphanes.
Recently, a visible-light-induced Ni-based catalyst C–P coupling reaction of diarylphosphane oxides with aryl halides has been developed by the Zhu group. The Ni(I) species and chlorine atom radical were generated via the ligand to the metal charge transfer process of NiCl2(PPh3)2, which allows the formation of Ni(IV)–P species, giving various tertiary phosphane oxides under photocatalyst-free conditions at room temperature in good yields (40–75%) [80].
Aside from homogeneous systems, effort was made on the development of heterogeneous systems based on Ni/CeO2 or Ni/Al2O3 nanocatalysts for the coupling of aryl iodides or bromides with diphenyl phosphane oxides in the presence of K2CO3 [81]. The reaction is scalable and the nanocatalyst can be recycled without loss of activity.

3.3. Synthetic Routes of Phosphinates

Historically, one of the first methods of non-catalytic C‒P bond formation was reactions of alkyl(phenyl)phosphonous esters with aryl halides. One of the first catalytic versions of this reaction was performed with nickel complexes and aryl(alkyl)phosphonous esters with aryl bromides to give various phenyl-arylphosphinates with good to excellent yields (52–82%) (Figure 24) [82,83].
The carbon–phosphorus cross-coupling of aryl tosylates or mesylates has been accomplished with high yields up to 92%, with ethyl phenylphosphinate using NiCl2(dppf)/dppf (dppf—1,1′-bis(diphenylphosphino)ferrocene) at 100 °C in the presence of diethylisopropylamine (DIPEA) and zinc (Figure 25) [84]. The reaction can also be extended to the coupling of diarylphosphane oxides and diethylphosphonate with aryl sulfonates.
Later, Gao and co-workers achieved the coupling of phenylboronic acid with ethyl phenylphosphinate employing a NiBr2/pyridine system in the presence of K2CO3 (73% yield) (Figure 26) [75]. The reaction can be extended to the coupling of a series of phosphites and phosphane oxides.

3.4. Synthesis of Phosphonates by C‒P Cross-Coupling

3.4.1. Ni-Catalyzed Phosphonylation with Phosphites

The reaction of trialkyl phosphites with alkyl halides (Arbuzov reaction) leads to the formation of phosphonates RP(O)(OR’)2. The Ni-catalyzed procedure using trialkyl phosphites or dialkyl arylphosphonites was described fifty years ago by Tavs et al. and allowed the involvement of aryl and alkenyl halides in this reaction [85]. This work is the first example of a transition metal-catalyzed cross-coupling reaction of C–P bond formation. The reaction requires harsh conditions (150–200 °C), but the yields of aryl- and alkenylphosphonates are high. The reaction proceeded smoothly using aryl iodides or bromides and phosphite and phosphonite ethyl esters (Figure 27). Exceptions are reactions with ortho-substituted aryl halides, where the yield decreases to 15–40%. Even alkenyl chlorides, such as α- and β-chlorostyrenes and vinyl chloride, give rise to alkenylphosphonates in high yields under NiCl2 catalysis [86]. Phosphinites also reacted with good yields [87].
The catalytic arylation of tris(trimethysilyl)phosphite occurs with higher rates (Figure 28) [88]. The treatment of the obtained arylphosphonates with methanol at room temperature gives the corresponding phosphonic acids in quantitative yield.
The reactions of triethylphosphite with akenylhalides having an alkoxy or diethylamino group in position 1 occurred under milder conditions (Figure 29) [89]. Similar reactions with 2-halovinyl ethers or 2-haloenamines proceeded at a higher temperature, but the yields were also quite high [90].
Another example is the conversion of (E,E)-1,3-diiodobutadiene into bis-1,4-(diethoxyphosphanoyl)-1,3-butadiene in high yields [23]. In the above reactions, triethyl phosphite acts as both a phosphorylating and a reducing agent. Balthazor showed that the reaction involves the formation of a Ni(0) phosphite complex in the presence of a slight excess of the trialkyl phosphite [91]. This phosphite Ni(0) complex undergoes fast oxidative addition with the aryl(vinyl) iodide, followed by slow decomposition to give the quasi-phosphonium salt, which is finally transformed into arylphosphonate via classical Arbuzov rearrangement (Figure 30).
Heinicke et al. described the reaction of triethylphosphite with 2-haloanilides in the presence of NiX2 complexes. The resulting o-acylaminophenylphosphonates are useful intermediates in the synthesis of 1H-1,3-benzazaphospholes (Figure 31) [92].
An efficient method has been developed for the Ni-catalyzed phosphonylation of aryl triflates with triethylphosphite, in which KBr as an additive promotes the SN2 catalytic step (Figure 32). This is the first example of nickel-catalyzed Arbuzov-type reaction of aryl triflates. Most of the substrates showed good reactivity with the use of these catalytic systems and good to high yields (46–95%) [93].

3.4.2. Phosphonylation by Hirao Reaction

The Hirao reaction is an alternative synthetic method for obtaining a wide range of phosphonates, which proceeds under milder reaction conditions and with higher yields. It can be considered as a classic example of C‒P cross-coupling with the formation of a Csp2–P bond.
In the 1980s, Hirao et al. reported the first Pd-catalyzed cross-coupling reaction of dialkyl phosphites with aryl and vinyl bromides, resulting in dialkyl arylphosphonates and dialkyl vinylphosphonates, respectively [94]. Later, reactions of allyl acetates and allyl carbonates with dialkyl phosphites were investigated under Ni catalysis in the presence of bis(trimethylsilyl)acetamide (BSA) as a base [95,96]. A direct comparison of nickel- and palladium-catalyzed cross-coupling for vinyl halides was reported by Beletskaya and co-workers [90]. Aryl and vinyl iodides as well as bromides reacted smoothly under these reaction conditions, while the corresponding chlorides were unreactive.
Nowadays, nickel catalysis is often applied in Hirao C–P cross-coupling reactions, and a range of organic and organometallic compounds including organohalides, alcohol or phenol derivatives [97,98,99,100], aryl, benzyl or allyl ammonium salts [101], sulfides [102] and aryl nitriles [103,104] have been employed as the carbon coupling partners. Reductive procedures involve Ni(II) salts together with Zn/Mg as the reductant or without reductive agents or Ni(0)(cod)2 as the catalyst precursor [75].
Han and co-workers extended the substrate scope to aryl bromides and chlorides with dimethylphosphite and diphenylphosphane oxide using NiCl2(dppp) (where dppp—1,3-bis(diphenylphosphano)propane) in the presence of potassium phosphite K3PO4, leading to phosphonates with 50–96% yields (Figure 33) [38].
Challenging phenol derivatives could also be involved in C‒P coupling reactions after converting the hydroxyl function to a better leaving group by reaction with bromotripyrrolidinophosphonium hexafluorophosphate (PyBroP) (Figure 34) [105]. The method allows the C–P cross-coupling to be carried out in a one-pot procedure without the isolation of an activated phenol intermediate.
The substrate scope is further expanded for the coupling of aryl mesylates with dimethylphosphite and diphenylphosphane oxide utilizing NiCl2(dppf)/dppf at 100 °C in the presence of diisopropylamine and zinc [84]. Later, the decarboxylative coupling of alkenyl acids with H-phosphonates was shown to obtain (E)-1-alkenylphosphonates (Figure 35) [106]. The reaction utilizes NiCl2(dppf) with Ag2O at 100 °C under a nitrogen atmosphere. The substrate scope can be extended to the coupling of alkynyl acids to produce alkynyl phosphonates in moderate yields up to 92%.
A catalytic deamidative phosphorylation of a wide range of amides using a Ni catalyst giving aryl phosphonates in good to excellent yields was reported (Figure 36) [107]. This method tolerates a wide range of functional groups. The reaction constitutes the first example of the transition metal-catalyzed generation of a C‒P bond from amides.
Keglevich and co-workers found that NiCl2 may also be a suitable catalyst in the microwave-assisted C–P coupling of bromobenzene and different >P(O)H species [108]. The experiments were carried out at 150 °C under MW irradiation, applying K2CO3 in the absence of any solvent with 68–92% (Figure 37). The NiCl2-catalyzed phosphonylation of substituted bromoarenes led to similar results as in the presence of Pd(OAc)2, but the scope of the aryl bromides was somewhat limited.
Taking into account the reaction conditions, costs and safety concerns, it can be concluded that the application of Pd(OAc)2 is favorable, but the use of NiCl2 can also be a good alternative. Moreover, the C–P coupling reactions which apply Ni(II) salts in the absence of reductants have been investigated earlier, including theoretical calculations [109,110]. These latest developments of Hirao coupling mean a big step forward to “P-ligand-free” catalytic reactions, since there is no need for sensitive and expensive P-ligands.
Recently, Budnikova and co-workers have demonstrated the possibility of the electrochemical phosphorylation of aromatic compounds (benzene and coumarins [111], pyridines [112], azoles [113]) with dialkyl phosphites (Figure 38). This novel approach is based on the oxidation of a mixture of the aromatic compound and diethyl phosphite (1:1) under mild electrochemical conditions (room temperature, atmospheric pressure) in the presence of bimetallic catalytic systems: 1% of Mn(II)(bpy)/Ni(BF4)(bpy). This method allows one to obtain diethyl arylphosphonates in good yields (up to 70%) and 100% conversion of the phosphite [114].

4. Summary and Outlook

C‒P cross-coupling reactions have made significant progress in recent years. Although Pd-catalyzed reactions dominate, Ni-based catalytic systems have been considerably explored. This article may give a good overview on the present state of the art of the Ni-catalyzed synthesis of racemic and scalemic phosphanes, phosphonium salts, phosphane oxides and phosphorus acid derivatives. Additionally, some green chemical approaches, such as MW activation, solvent- and reducing agent-free and electrochemical methods, have been outlined. The renaissance in nickel catalysis has brought new life to well-known nickel salts NiX2 (X=Cl, Br, OAc, acac, etc.), which have been used as catalyst precursors. Solubility in organic solvents and the easy transformation of Ni(acac)2 to the catalyst active form ensure important preferences for practical applications.
The increasing price of Pd, Pt and other noble metals even further stimulates the search for inexpensive and easily available Ni-based catalysts. Although the field of nickel catalysis has rapidly expanded over the last decade, there are many challenges that remain to be overcome. Nickel catalysts retain significant synthetic potential, are very reactive and design/control of their catalytic systems requires much more effort. Indeed, in the majority of known Ni-mediated reactions, the active catalyst remains unknown. We expect to see further developments in the area of Ni-catalyzed C‒P bond formation, particularly in the expansion of substrate scope and the development of low-cost, air-stable and easy-to-handle sources of nickel for catalysis.

Author Contributions

Conceptualization and methodology, A.A.Z., I.F.S., V.A.M. and D.G.Y.; writing—original draft preparation, A.A.Z. and I.F.S.; writing—review and editing, V.A.M. and D.G.Y. All authors have read and agreed to the published version of the manuscript.

Funding

This research activity was funded by the Government assignment for FRC Kazan Scientific Center of RAS.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

Abbreviations

Acacetyl radical
acacacetylacetone
BDPP(2S,4S)-2,4-bis(diphenylphosphino)pentane
BINAP2,2′-bis(diphenylphosphino)-1,1′-binaphthyl
bpy2,2′-bipyridine
BSAbis(trimethylsilyl)acetamide
DABCO1,4-diazabicyclo [2.2.2]octane
DCE1,2-dichloroethane
dcyptbis(dicyclohexylphosphano)thiophene
DIPEAdiethylisopropylamine
DME1,2-dimethoxyethane
DMFN,N-dimethylformamide
Dmp2,6-dimesitylphenyl radical
dppe1,2-bis(diphenylphosphano)ethane
dppf1,1′-bis(diphenylphosphino)ferrocene
dppp1,3-bis(diphenylphosphano)propane).
dtbpe1,2-bis(di-tert-butylphosphino)ethane
MW irradiationmicrowave irradiation
naphthylnaphthyl radical
NMPN-methyl-2-pyrrolidone
Pigiphosbis{(R)-1-[(S)-2-(diphenylphosphino)ferrocenyl]ethyl}cyclohexylphosphane)
PINAP(R)-(+)-4-[2-(diphenylphosphino)-1-naphthalenyl]-N-[-1-phenylethyl]-1-phthalazinamine
Pypyridine
PyBroPbromotripyrrolidinophosphonium hexafluorophosphate

References

  1. Allen, D.W. Phosphines and related C–P bonded compounds. Organophosphorus Chem. 2016, 45, 1–50. [Google Scholar] [CrossRef] [Green Version]
  2. Peruzzini, M.; Gonsalvi, L. Phosphorus Compounds: Advanced Tools in Catalysis and Material Sciences; Springer: Dordrecht, The Netherlands, 2011. [Google Scholar]
  3. Borner, A. Phosphorus Ligands in Asymmetric Catalysis; Wiley: Weinheim, Germany, 2008. [Google Scholar]
  4. Rodriguez, J.B.; Gallo-Rodriguez, C. The role of the phosphorus atom in drug design. ChemMedChem 2018, 14, 190–216. [Google Scholar] [CrossRef] [Green Version]
  5. Zagidullin, A.A.; Bezkishko, I.A.; Miluykov, V.A.; Sinyashin, O.G. Phospholes–development and recent advances. Mendeleev Commun. 2013, 23, 117–130. [Google Scholar] [CrossRef]
  6. Schwan, A. Palladium catalyzed cross-coupling reactions for phosphorus–carbon bond formation. Chem. Soc. Rev. 2004, 33, 218–224. [Google Scholar] [CrossRef]
  7. Oestreich, M.; Tappe, F.; Trepohl, V. Transition-Metal-Catalyzed C-P Cross-Coupling Reactions. Synthesis 2010, 2010, 3037–3062. [Google Scholar] [CrossRef]
  8. Wauters, I.; Debrouwer, W.; Stevens, C.V. Preparation of phosphines through C–P bond formation. Beilstein J. Org. Chem. 2014, 10, 1064–1096. [Google Scholar] [CrossRef]
  9. Koshti, V.; Gaikwad, S.; Chikkali, S.H. Contemporary avenues in catalytic PH bond addition reaction: A case study of hydrophosphination. Coord. Chem. Rev. 2014, 265, 52–73. [Google Scholar] [CrossRef]
  10. Tasker, S.Z.; Standley, E.A.; Jamison, T.F. Recent advances in homogeneous nickel catalysis. Nat. Cell Biol. 2014, 509, 299–309. [Google Scholar] [CrossRef] [Green Version]
  11. Henrion, M.; Ritleng, V.; Chetcuti, M.J. Nickel N-Heterocyclic Carbene-Catalyzed C–C Bond Formation: Reactions and Mechanistic Aspects. ACS Catal. 2015, 5, 1283–1302. [Google Scholar] [CrossRef]
  12. Trotuş, I.-T.; Zimmermann, T.; Schüth, F. Catalytic reactions of acetylene: A feedstock for the chemical industry revisited. Chem. Rev. 2013, 114, 1761–1782. [Google Scholar] [CrossRef]
  13. Ananikov, V. Nickel: The “Spirited Horse” of Transition Metal Catalysis. ACS Catal. 2015, 5, 1964–1971. [Google Scholar] [CrossRef]
  14. Cristau, H.-J.; Chêne, A.; Christol, H. Arylation catalytique d’organophosphorés. Produits de l’arylation, catalysée par les sels de nickel (II), de composés du phosphore tricoordiné. J. Organomet. Chem. 1980, 185, 283–295. [Google Scholar] [CrossRef]
  15. Ager, D.J.; Laneman, S.A. Convenient and direct preparation of tertiary phosphines via nickel-catalysed cross-coupling. Chem. Commun. 1997, 2359–2360. [Google Scholar] [CrossRef]
  16. Le Gall, E.; Troupel, M.; Nédélec, J.-Y. Nickel-catalyzed reductive coupling of chlorodiphenylphosphine with aryl bromides into functionalized triarylphosphines. Tetrahedron 2003, 59, 7497–7500. [Google Scholar] [CrossRef]
  17. Sun, M.; Zang, Y.-S.; Hou, L.-K.; Chen, X.-X.; Sun, W.; Yang, S.-D. Convenient Formation of Triarylphosphines by Nickel-Catalyzed C-P Cross-Coupling with Aryl Chlorides. Eur. J. Org. Chem. 2014, 2014, 6796–6801. [Google Scholar] [CrossRef]
  18. Beletskaya, I.P.; Afanasiev, V.V.; Kazankova, M.A.; Efimova, I.V. New Approach to Phosphinoalkynes Based on Pd- and Ni-Catalyzed Cross-Coupling of Terminal Alkynes with Chlorophosphanes. Org. Lett. 2003, 5, 4309–4311. [Google Scholar] [CrossRef]
  19. Di Credico, B.; De Biani, F.F.; Gonsalvi, L.; Guerri, A.; Ienco, A.; Laschi, F.; Peruzzini, M.; Reginato, G.; Rossin, A.; Zanello, P. Cyclopentadienyl ruthenium(II) complexes with bridging alkynylphosphine ligands: Synthesis and electrochemical studies. Chem. Eur. J. 2009, 15, 11985–11998. [Google Scholar] [CrossRef] [PubMed]
  20. Beletskaya, I.P.; Afanasiev, V.V.; Kazankova, M.A.; Efimova, I.V.; Antipin, M.U. A convenient and direct route to phosphinoalkynes via copper-catalyzed cross-coupling of terminal alkynes with chlorophosphanes. Synthesis 2003, 2003, 2835–2838. [Google Scholar] [CrossRef]
  21. Kazankova, M.A.; Shulyupin, M.O.; Chirkov, E.A.; Beletskaya, I.P. Nickel-catalyzed cross-coupling of diphenylphosphine with vinyl bromides and chlorides as a route to diphenylvinylphosphines. Synlett 2005, 2005, 658–660. [Google Scholar] [CrossRef]
  22. Beletskaya, I.; Kazankova, M.A. Catalytic methods for building up phosphorus-carbon bond. Russ. J. Org. Chem. 2002, 38, 1391–1430. [Google Scholar] [CrossRef]
  23. Trostyanskaya, I.G.; Titskiy, D.Y.; Anufrieva, E.A.; Borisenko, A.A.; Kazankova, M.A.; Beletskaya, I.P. Cross-coupling of E,E-1,4-diiodobuta-1,3-diene with nucleophiles catalyzed by Pd or Ni complexes: A new route to functionalized dienes. Russ. Chem. Bull. 2001, 50, 2095–2100. [Google Scholar] [CrossRef]
  24. Berthod, M.; Mignani, G.; Woodward, G.; Lemaire, M. Modified BINAP: The how and the why. Chem. Rev. 2005, 105, 1801–1836. [Google Scholar] [CrossRef]
  25. Cai, D.; Payack, J.F.; Bender, D.R.; Hughes, D.L.; Verhoeven, T.R.; Reider, P.J. Synthesis of Chiral 2,2’-Bis(diphenylphosphino)-1,1’-binaphthyl (BINAP) via a Novel Nickel-Catalyzed Phosphine Insertion. J. Org. Chem. 1994, 59, 7180–7181. [Google Scholar] [CrossRef]
  26. Kerrigan, N.J.; Dunne, E.C.; Cunningham, D.; McArdle, P.; Gilligan, K.; Gilheany, D.G. Studies in the preparation of novel P-chirogenic binaphthyl monophosphanes (MOPs). Tetrahedron Lett. 2003, 44, 8461–8465. [Google Scholar] [CrossRef]
  27. Morris, D.J.; Docherty, G.; Woodward, G.; Wills, M. Modification of ligand properties of phosphine ligands for C–C and C–N bond-forming reactions. Tetrahedron Lett. 2007, 48, 949–953. [Google Scholar] [CrossRef]
  28. Fleming, W.J.; Müller-Bunz, H.; Lillo, V.; Fernandez, E.; Guiry, P.J. Axially chiral P-N ligands for the copper catalyzed β-borylation of α,β-unsaturated esters. Org. Biomol. Chem. 2009, 7, 2520–2524. [Google Scholar] [CrossRef]
  29. Fleming, W.J.; Müller-Bunz, H.; Guiry, P.J. Synthesis and post-resolution modification of new axially chiral ligands for asymmetric catalysis. Eur. J. Org. Chem. 2010, 2010, 5996–6004. [Google Scholar] [CrossRef]
  30. Rafter, E.; Muldoon, J.; Bunz, H.M.; Gilheany, D.G. Synthesis of P-stereogenic BINAP bissulfide analogues. Tetrahedron Asymmetry 2011, 22, 1680–1686. [Google Scholar] [CrossRef]
  31. Pereira, M.M.; Calvete, M.J.F.; Carrilho, R.M.B.; Abreu, A.R. Synthesis of binaphthyl based phosphine and phosphite ligands. Chem. Soc. Rev. 2013, 42, 6990–7027. [Google Scholar] [CrossRef]
  32. Enev, V.; Ewers, C.L.J.; Harre, M.; Nickisch, K.; Mohr, J.T. A Bis-steroidal phosphine as new chiral hydrogenation ligand. J. Org. Chem. 1997, 62, 7092–7093. [Google Scholar] [CrossRef]
  33. Lacey, P.M.; McDonnell, C.M.; Guiry, P.J. Synthesis and resolution of 2-methyl-Quinazolinap, a new atropisomeric phosphinamine ligand for asymmetric catalysis. Tetrahedron Lett. 2000, 41, 2475–2478. [Google Scholar] [CrossRef]
  34. Knöpfel, T.F.; Aschwanden, P.; Ichikawa, T.; Watanabe, T.; Carreira, E.M. Readily Available Biaryl P,N Ligands for Asymmetric Catalysis. Angew. Chem. Int. Ed. 2004, 43, 5971–5973. [Google Scholar] [CrossRef] [PubMed]
  35. Knöpfel, T.F.; Zarotti, P.; Ichikawa, T.; Carreira, E. Catalytic, enantioselective, conjugate alkyne addition. J. Am. Chem. Soc. 2005, 127, 9682–9683. [Google Scholar] [CrossRef]
  36. Birdsall, D.J.; Hope, E.G.; Stuart, A.M.; Chen, W.; Hu, Y.; Xiao, J. Synthesis of fluoroalkyl-derivatised BINAP ligands. Tetrahedron Lett. 2001, 42, 8551–8553. [Google Scholar] [CrossRef]
  37. Nakamura, Y.; Takeuchi, S.; Zhang, S.; Okumura, K.; Ohgo, Y. Preparation of a fluorous chiral BINAP and application to an asymmetric Heck reaction. Tetrahedron Lett. 2002, 43, 3053–3056. [Google Scholar] [CrossRef]
  38. 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. Chem. Eur. J. 2012, 18, 9622–9627. [Google Scholar] [CrossRef]
  39. Klein, A.; Budnikova, Y.H.; Sinyashin, O. Electron transfer in organonickel complexes of α-diimines: Versatile redox catalysts for C–C or C–P coupling reactions—A review. J. Organomet. Chem. 2007, 692, 3156–3166. [Google Scholar] [CrossRef]
  40. Budnikova, Y.H.; Perichon, J.; Yakhvarov, D.G.; Kargin, Y.M.; Sinyashin, O. Highly reactive σ-organonickel complexes in electrocatalytic processes. J. Organomet. Chem. 2001, 630, 185–192. [Google Scholar] [CrossRef]
  41. Yakhvarov, D.G.; Khusnuriyalova, A.F.; Sinyashin, O. Electrochemical synthesis and properties of organonickel σ-complexes. Organometallics 2014, 33, 4574–4589. [Google Scholar] [CrossRef]
  42. Yakhvarov, D.G.; Kvashennikova, S.V.; Sinyashin, O. Reactions of activated organonickel σ-complexes with elemental (white) phosphorus. Russ. Chem. Bull. 2013, 62, 2472–2476. [Google Scholar] [CrossRef]
  43. Gafurov, Z.; Musin, L.; Sakhapov, I.F.; Babaev, V.; Musina, E.I.; Karasik, A.A.; Sinyashin, O.; Yakhvarov, D.G. The formation of secondary arylphosphines in the reaction of organonickel sigma-complex [NiBr(Mes)(bpy)], where Mes = 2,4,6-trimethylphenyl, bpy = 2,2′-bipyridine, with phenylphosphine. Phosphorus Sulfur Silicon Relat. Elements 2016, 191, 1475–1477. [Google Scholar] [CrossRef]
  44. Gafurov, Z.; Sakhapov, I.F.; Kagilev, A.A.; Kantyukov, A.O.; Khayarov, K.R.; Sinyashin, O.; Yakhvarov, D.G. The formation of mesitylphosphine and dimesitylphosphine in the reaction of organonickel σ-complex [NiBr(Mes)(bpy)] (Mes = 2,4,6-trimethylphenyl, bpy = 2,2′-bipyridine) with phosphine PH3. Phosphorus Sulfur Silicon Relat. Elements 2020, 195, 726–729. [Google Scholar] [CrossRef]
  45. Bange, C.A.; Waterman, R. Challenges in Catalytic Hydrophosphination. Chem. Eur. J. 2016, 22, 12598–12605. [Google Scholar] [CrossRef]
  46. Kazankova, M.A.; Shulyupin, M.O.; Borisenko, A.A.; Beletskaya, I. Synthesis of alkyl(diphenyl)phosphines by hydrophosphination of vinylarenes catalyzed by transition metal complexes. Russ. J. Org. Chem. 2002, 38, 1479–1484. [Google Scholar] [CrossRef]
  47. Beletskaya, I.P.; Kazankova, M.A.; Shulyupin, M.O. Catalytic Hydrophosphination of Alkenylalkyl Ethers. Synlett 2003, 2003, 2155–2158. [Google Scholar] [CrossRef]
  48. Ananikov, V.; Gayduk, K.A.; Starikova, Z.A.; Beletskaya, I. Ni(acac)2/phosphine as an excellent precursor of Nickel(0) for catalytic systems. Organometallics 2010, 29, 5098–5102. [Google Scholar] [CrossRef]
  49. Webster, R.L. Room Temperature Ni(II) Catalyzed Hydrophosphination and Cyclotrimerization of Alkynes. Inorganics 2018, 6, 120. [Google Scholar] [CrossRef] [Green Version]
  50. Shulyupin, M.O.; Kazankova, M.A.; Beletskaya, I. Catalytic Hydrophosphination of Styrenes. Org. Lett. 2002, 4, 761–763. [Google Scholar] [CrossRef]
  51. Ganushevich, Y.S.; Miluykov, V.A.; Polyancev, F.M.; Latypov, S.; Lönnecke, P.; Hey-Hawkins, E.; Yakhvarov, D.G.; Sinyashin, O. Nickel Phosphanido Hydride Complex: An Intermediate in the Hydrophosphination of Unactivated Alkenes by Primary Phosphine. Organometallics 2013, 32, 3914–3919. [Google Scholar] [CrossRef]
  52. Latypov, S.K.; Kondrashova, S.A.; Polyancev, F.M.; Sinyashin, O.G. Quantum Chemical Calculations of 31P NMR Chemical Shifts in Nickel Complexes: Scope and Limitations. Organometallics 2020, 39, 1413–1422. [Google Scholar] [CrossRef]
  53. Latypov, S.K.; Polyancev, F.M.; Ganushevich, Y.S.; Miluykov, V.A.; Sinyashin, O.G. Mechanism of intramolecular transformations of nickel phosphanido hydride complexes. Dalton Trans. 2015, 45, 2053–2059. [Google Scholar] [CrossRef]
  54. Wang, C.; Huang, K.; Ye, J.; Duan, W.-L. Asymmetric Synthesis of P-Stereogenic Secondary Phosphine-Boranes by an Unsymmetric Bisphosphine Pincer-Nickel Complex. J. Am. Chem. Soc. 2021, 143, 5685–5690. [Google Scholar] [CrossRef]
  55. Sadow, A.D.; Haller, I.; Fadini, A.L.; Togni, A. Nickel(II)-Catalyzed Highly Enantioselective Hydrophosphination of Methacrylonitrile. J. Am. Chem. Soc. 2004, 126, 14704–14705. [Google Scholar] [CrossRef] [PubMed]
  56. Sadow, A.D.; Togni, A. Enantioselective addition of secondary phosphines to methacrylonitrile: Catalysis and mechanism. J. Am. Chem. Soc. 2005, 127, 17012–17024. [Google Scholar] [CrossRef]
  57. Kazankova, M.A.; Efimova, I.V.; Kochetkov, A.N.; Afanas’Ev, V.V.; Beletskaya, I.; Dixneuf, P.H. New Approach to Vinylphosphines Based on Pd- and Ni-Catalyzed Diphenylphosphine Addition to Alkynes. Synlett 2001, 2001, 0497–0500. [Google Scholar] [CrossRef]
  58. Kazankova, M.A.; Efimova, I.V.; Kochetkov, A.; Afanas’Ev, V.V.; Beletskaya, I.P. Synthesis of Vinylphosphines by Hydrophosphination of Alkynes in the Presence of Transition Metal Complexes. Russ. J. Org. Chem. 2002, 38, 1465–1474. [Google Scholar] [CrossRef]
  59. Ananikov, V.P.; Beletskaya, I. Alkyne insertion into the M-P and M-H bonds (M=Pd, Ni, Pt, and Rh): A theoretical mechanistic study of the C-P and C-H bond-formation steps. Chem. Asian J. 2011, 6, 1423–1430. [Google Scholar] [CrossRef]
  60. Yan, J.; Wang, Y.; Hou, S.; Shi, L.; Zhu, X.; Hao, X.; Song, M. NCC Pincer Ni (II) complexes Catalyzed hydrophosphination of Nitroalkenes with diphenylphosphine. Appl. Organomet. Chem. 2020, 34, 5954. [Google Scholar] [CrossRef]
  61. Ribiere, P.; Bravo-Altamirano, K.; Antczak, M.I.; Hawkins, J.D.; Montchamp, J.-L. NiCl2-Catalyzed Hydrophosphinylation. J. Org. Chem. 2005, 70, 4064–4072. [Google Scholar] [CrossRef]
  62. Ananikov, V.; Khemchyan, L.L.; Beletskaya, I.; Starikova, Z.A. Acid-Free Nickel Catalyst for Stereo- and Regioselective Hydrophosphorylation of Alkynes: Synthetic Procedure and Combined Experimental and Theoretical Mechanistic Study. Adv. Synth. Catal. 2010, 352, 2979–2992. [Google Scholar] [CrossRef]
  63. Han, L.-B.; Ono, Y.; Yazawa, H. Nickel-Catalyzed Addition of P(O)−H Bonds to Propargyl Alcohols: One-Pot Generation of Phosphinoyl 1,3-Butadienes. Org. Lett. 2005, 7, 2909–2911. [Google Scholar] [CrossRef]
  64. Han, L.-B.; Zhang, C.; Yazawa, A.H.; Shimada, S. Efficient and Selective Nickel-Catalyzed Addition of H−P(O) and H−S Bonds to Alkynes. J. Am. Chem. Soc. 2004, 126, 5080–5081. [Google Scholar] [CrossRef]
  65. Cassar, L.; Foà, M. Nickel-catalyzed synthesis of phosphonium salts. J. Organomet. Chem. 1974, 74, 75. [Google Scholar] [CrossRef]
  66. Allen, D.W.; Coles, S.J.; Light, M.E.; Hursthouse, M.B. Synthesis and X-ray crystal structures of organotri(2-furyl)phosphonium salts: Effects of 2-furyl substituents at phosphorus on intramolecular nitrogen to phosphorus hypervalent coordinative interactions. Inorganica Chim. Acta 2004, 357, 1558–1564. [Google Scholar] [CrossRef]
  67. Monkowius, U.V.; Nogai, S.; Schmidbaur, H. Unsuccessful/successful attempts to produce penta(heteroaryl)-phosphoranes/-arsoranes R5E (E = P, As; R = 2-furyl, 2-thienyl). Dalton Trans. 2004, 10, 1610–1617. [Google Scholar] [CrossRef] [PubMed]
  68. Asay, M.; Donnadieu, B.; Baceiredo, A.; Soleilhavoup, M.; Bertrand, G. Cyclic (Amino)[bis(ylide)]carbene as an Anionic Bidentate Ligand for Transition-Metal Complexes. Inorg. Chem. 2008, 47, 3949–3951. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  69. Manabe, K. Synthesis of novel chiral quaternary phosphonium salts with a multiple hydrogen-bonding site, and their application to asymmetric phase-transfer alkylation. Tetrahedron 1998, 54, 14465–14476. [Google Scholar] [CrossRef]
  70. Marcoux, D.; Charette, A.B. Nickel-Catalyzed Synthesis of Phosphonium Salts from Aryl Halides and Triphenylphosphine. Adv. Synth. Catal. 2008, 350, 2967–2974. [Google Scholar] [CrossRef]
  71. Wan, W.; Yang, X.; Smith, R.C. Convenient route to tetraarylphosphonium polyelectrolytes via metal-catalysed P–C coupling polymerisation of aryl dihalides and diphenylphosphine. Chem. Commun. 2016, 53, 252–254. [Google Scholar] [CrossRef]
  72. Budnikova, Y.H.; Sinyashin, O.G.; Budnikova, Y.H. Phosphorylation of C–H bonds of aromatic compounds using metals and metal complexes. Russ. Chem. Rev. 2015, 84, 917–951. [Google Scholar] [CrossRef]
  73. 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]
  74. Zhang, X.; Liu, H.; Hu, X.; Tang, G.; Zhu, J.; Zhao, Y. Ni(II)/Zn Catalyzed Reductive Coupling of Aryl Halides with Diphenylphosphine Oxide in Water. Org. Lett. 2011, 13, 3478–3481. [Google Scholar] [CrossRef]
  75. Hu, G.; Chen, W.; Fu, T.; Peng, Z.; Qiao, H.; Gao, Y.; Zhao, Y. Nickel-Catalyzed C–P Cross-Coupling of Arylboronic Acids with P(O)H Compounds. Org. Lett. 2013, 15, 5362–5365. [Google Scholar] [CrossRef]
  76. Liu, L.L.; Wang, Y.; Zeng, Z.; Xu, P.; Gao, Y.; Yin, Y.; Zhao, Y. Nickel(II)-Magnesium-Catalyzed Cross-Coupling of 1,1-Dibromo-1-alkenes with Diphenylphosphine Oxide: One-Pot Synthesis of (E)-1-Alkenylphosphine Oxides or Bisphosphine Oxides. Adv. Synth. Catal. 2013, 355, 659–666. [Google Scholar] [CrossRef]
  77. Xuan, J.; Zeng, T.-T.; Chen, J.-R.; Lu, L.-Q.; Xiao, W.-J. Room Temperature C-P Bond Formation Enabled by Merging Nickel Catalysis and Visible-Light-Induced Photoredox Catalysis. Chem. Eur. J. 2015, 21, 4962–4965. [Google Scholar] [CrossRef]
  78. 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]
  79. Liu, X.-T.; Zhang, Y.-Q.; Han, X.-Y.; Sun, S.-P.; Zhang, Q.-W. Ni-Catalyzed Asymmetric Allylation of Secondary Phosphine Oxides. J. Am. Chem. Soc. 2019, 141, 16584–16589. [Google Scholar] [CrossRef]
  80. Hou, H.; Zhou, B.; Wang, J.; Sun, D.; Yu, H.; Chen, X.; Han, Y.; Shi, Y.; Yan, C.; Zhu, S. Visible-light-induced ligand to metal charge transfer excitation enabled phosphorylation of aryl halides. Chem. Commun. 2021, 57, 5702–5705. [Google Scholar] [CrossRef]
  81. Łastawiecka, E.; Flis, A.; Stankevič, M.; Greluk, M.; Słowik, G.; Gac, W. P-Arylation of secondary phosphine oxides catalyzed by nickel-supported nanoparticles. Org. Chem. Front. 2018, 5, 2079–2085. [Google Scholar] [CrossRef]
  82. Balthazor, T.M. Phosphindolin-3-one. A useful intermediate for phosphindole synthesis. J. Org. Chem. 1980, 45, 2519–2522. [Google Scholar] [CrossRef]
  83. Hirao, T.; Masunaga, T.; Ohshiro, Y.; Agawa, T. Stereoselective synthesis of vinylphosphonate. Tetrahedron Lett. 1980, 21, 3595–3598. [Google Scholar] [CrossRef]
  84. Shen, C.; Yang, G.; Zhang, W. Nickel-catalyzed C–P coupling of aryl mesylates and tosylates with H(O)PR1R2. Org. Biomol. Chem. 2012, 10, 3500–3505. [Google Scholar] [CrossRef]
  85. Tavs, P. Reaktion von Arylhalogeniden mit Trialkylphosphiten und Benzolphosphonigsäure-Dialkylestern zu Aromatischen Phosphonsäureestern und Phosphinsäureestern unter Nickelsalzkatalyse. Eur. J. Inorg. Chem. 1970, 103, 2428–2436. [Google Scholar] [CrossRef]
  86. Tavs, P.; Weitkamp, H. Herstellung und KMR-spektren einiger α,β-ungesättigter phosphonsäureester: Nickelsalzkatalysierte reaktion von vinylhalogeniden mit trialkylphosphiten. Tetrahedron 1970, 26, 5529–5534. [Google Scholar] [CrossRef]
  87. Comins, D.L.; Ollinger, C.G. Inter- and intramolecular Horner–Wadsworth–Emmons reactions of 5-(diethoxyphosphoryl)-1-acyl-2-alkyl(aryl)-2,3-dihydro-4-pyridones. Tetrahedron Lett. 2001, 42, 4115–4118. [Google Scholar] [CrossRef]
  88. DeMik, N.N.; Kabachnik, M.M.; Novikova, Z.S.; Beletskaya, I.P. Preparation of arylphosphonates by the reaction of aryl halides with tris(trimethylsilyl) phosphite under homogeneous catalysis conditions. Russ. Chem. Bull. 1991, 40, 1300–1301. [Google Scholar] [CrossRef]
  89. Kazankova, M.A.; Trostyanskaya, I.G.; Lutsenko, S.V.; Efimova, I.V.; Beletskaya, I.P. Synthesis of 1- and 2-alkoxy- and dialkylaminoalkenyl-phosphonates, catalyzed by transition metal complexes. Russ. J. Org. Chem. 1999, 35, 452–458. [Google Scholar]
  90. Kazankova, M.A.; Trostyanskaya, I.G.; Lutsenko, S.V.; Beletskaya, I. Nickel- and palladium-catalyzed cross-coupling as a route to 1- and 2-alkoxy- or dialkylaminovinylphosphonates. Tetrahedron Lett. 1999, 40, 569–572. [Google Scholar] [CrossRef]
  91. Balthazor, T.M.; Grabiak, R.C. Nickel-catalyzed Arbuzov reaction: Mechanistic observations. J. Org. Chem. 1980, 45, 5425–5426. [Google Scholar] [CrossRef]
  92. Heinicke, J.; Gupta, N.; Surana, A.; Peulecke, N.; Witt, B.; Steinhauser, K.; Bansal, R.K.; Jones, P.G. Synthesis of 1H-1,3-benzazaphospholes: Substituent influence and mechanistical aspects. Tetrahedron 2001, 57, 9963–9972. [Google Scholar] [CrossRef]
  93. Yang, G.; Shen, C.; Zhang, L.; Zhang, W. Nickel-catalyzed Arbuzov reactions of aryl triflates with triethyl phosphite. Tetrahedron Lett. 2011, 52, 5032–5035. [Google Scholar] [CrossRef]
  94. Hirao, T.; Masunaga, T.; Yamada, N.; Ohshiro, Y.; Agawa, T. Palladium-catalyzed New Carbon-Phosphorus Bond Formation. Bull. Chem. Soc. Jpn. 1982, 55, 909–913. [Google Scholar] [CrossRef] [Green Version]
  95. Lu, X.; Zhu, J. Nickel(0)-catalyzed Reaction of O,O-Dialkyl Phosphonates with Allyl Acetates or Carbonates. A Novel Method of Preparing Allyl Phosphonates. Synthesis 1986, 1986, 563–564. [Google Scholar] [CrossRef]
  96. Lu, X.; Tao, X.; Zhu, J.; Sun, X.; Xu, J. Regio- and Stereoselective Synthesis of Symmetrical and Unsymmetrical 1,3-Diphosphoryl(Phosphonyl, Phosphinyl) Substituted (E)-Propenes. Synthesis 1989, 1989, 848–850. [Google Scholar] [CrossRef]
  97. Yang, J.; Xiao, J.; Chen, T.; Han, L.-B. Nickel-catalyzed phosphorylation of aryl triflates with P(O)H compounds. J. Organomet. Chem. 2016, 820, 120–124. [Google Scholar] [CrossRef]
  98. Yang, J.; Chen, T.; Han, L.-B. C–P Bond-Forming Reactions via C–O/P–H Cross-Coupling Catalyzed by Nickel. J. Am. Chem. Soc. 2015, 137, 1782–1785. [Google Scholar] [CrossRef]
  99. Yang, J.; Xiao, J.; Chen, T.; Han, L.-B. Nickel-Catalyzed Phosphorylation of Phenol Derivatives via C–O/P–H Cross-Coupling. J. Org. Chem. 2016, 81, 3911–3916. [Google Scholar] [CrossRef]
  100. Liao, L.-L.; Gui, Y.-Y.; Zhang, X.-B.; Shen, G.; Liu, H.-D.; Zhou, W.-J.; Li-Li, L.; Yong-Yuan, G. Phosphorylation of Alkenyl and Aryl C–O Bonds via Photoredox/Nickel Dual Catalysis. Org. Lett. 2017, 19, 3735–3738. [Google Scholar] [CrossRef] [PubMed]
  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. Yang, J.; Xiao, J.; Chen, T.; Yin, S.-F.; Han, L.-B. Efficient nickel-catalyzed phosphinylation of C–S bonds forming C–P bonds. Chem. Commun. 2016, 52, 12233–12236. [Google Scholar] [CrossRef]
  103. Zhang, J.-S.; Chen, T.; Yang, J.; Han, L.-B. Nickel-catalysed P–C bond formation via P–H/C–CN cross coupling reactions. Chem. Commun. 2015, 51, 7540–7542. [Google Scholar] [CrossRef]
  104. Sun, M.; Zhang, H.-Y.; Han, Q.; Yang, K.; Yang, S.-D. Nickel-Catalyzed C–P Cross-Coupling by C–CN Bond Cleavage. Chem. A Eur. J. 2011, 17, 9566–9570. [Google Scholar] [CrossRef]
  105. 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] [PubMed]
  106. Wu, Y.; Liu, L.L.; Yan, K.; Xu, P.; Gao, Y.; Zhao, Y. Nickel-Catalyzed Decarboxylative C–P cross-coupling of alkenyl acids with P(O)H compounds. J. Org. Chem. 2014, 79, 8118–8127. [Google Scholar] [CrossRef] [PubMed]
  107. Liu, C.; Szostak, M. Decarbonylative phosphorylation of amides by palladium and nickel catalysis: The Hirao cross-coupling of amide derivatives. Angew. Chem. Int. Ed. 2017, 56, 12718–12722. [Google Scholar] [CrossRef]
  108. Jablonkai, E.; Balazs, L.; Keglevich, G. A P-ligand-free nickel-catalyzed variation of the hirao reaction under microwave conditions. Curr. Org. Chem. 2015, 19, 197–202. [Google Scholar] [CrossRef]
  109. Keglevich, G.; Henyecz, R.; Mucsi, Z. Focusing on the Catal. of the Pd- and Ni-Catalyzed Hirao Reactions. Molecules 2020, 25, 3897. [Google Scholar] [CrossRef] [PubMed]
  110. Henyecz, R.; Mucsi, Z.; Keglevich, G. A surprising mechanism lacking the Ni(0) state during the Ni(II)-catalyzed P–C cross-coupling reaction performed in the absence of a reducing agent–An experimental and a theoretical study. Pure Appl. Chem. 2020, 92, 493–503. [Google Scholar] [CrossRef] [Green Version]
  111. Khrizanforov, M.; Strekalova, S.; Kholin, K.; Khrizanforova, V.; Kadirov, M.; Gryaznova, T.; Budnikova, Y. Novel approach to metal-induced oxidative phosphorylation of aromatic compounds. Catal. Today 2017, 279, 133–141. [Google Scholar] [CrossRef]
  112. Strekalova, S.; Khrizanforov, M.; Sinyashin, O.; Budnikova, Y. Catalytic Phosphorylation of Aromatic C-H Bonds: From Traditional Approaches to Electrochemistry. Curr. Org. Chem. 2019, 23, 1756–1770. [Google Scholar] [CrossRef]
  113. Gryaznova, T.V.; Khrizanforov, M.; Strekalova, S.O.; Budnikova, Y.H.; Sinyshin, O.G. Electrochemical oxidative phosphonation of azoles. Phosphorus Sulfur Silicon Relat. Elem. 2016, 191, 1658–1659. [Google Scholar] [CrossRef]
  114. Budnikova, Y.H.; Gryaznova, T.V.; Grinenko, V.; Dudkina, Y.; Khrizanforov, M. Eco-efficient electrocatalytic C–P bond formation. Pure Appl. Chem. 2017, 89, 311–330. [Google Scholar] [CrossRef]
Figure 1. Nickel-catalyzed cross-coupling in the presence of zinc.
Figure 1. Nickel-catalyzed cross-coupling in the presence of zinc.
Molecules 26 05283 g001
Figure 2. Nickel-catalyzed cross-coupling with (trimethylsilyl)diphenylphosphane.
Figure 2. Nickel-catalyzed cross-coupling with (trimethylsilyl)diphenylphosphane.
Molecules 26 05283 g002
Figure 3. General scheme for a nickel-catalyzed cross-coupling reaction with terminal alkynes.
Figure 3. General scheme for a nickel-catalyzed cross-coupling reaction with terminal alkynes.
Molecules 26 05283 g003
Figure 4. Proposed mechanism for the nickel- and palladium-catalyzed C(sp)-P cross-coupling (M = Ni, Pd).
Figure 4. Proposed mechanism for the nickel- and palladium-catalyzed C(sp)-P cross-coupling (M = Ni, Pd).
Molecules 26 05283 g004
Figure 5. Nickel-catalyzed arylation of diphenylphosphane.
Figure 5. Nickel-catalyzed arylation of diphenylphosphane.
Molecules 26 05283 g005
Figure 6. Nickel-catalyzed phosphination of vinyl bromides.
Figure 6. Nickel-catalyzed phosphination of vinyl bromides.
Molecules 26 05283 g006
Figure 7. Nickel-catalyzed phosphination of vinyl iodides.
Figure 7. Nickel-catalyzed phosphination of vinyl iodides.
Molecules 26 05283 g007
Figure 8. Ni(dppe)Cl2-catalyzed BINAP synthesis.
Figure 8. Ni(dppe)Cl2-catalyzed BINAP synthesis.
Molecules 26 05283 g008
Figure 9. Selected examples of phosphane ligands obtained by Cai’s method.
Figure 9. Selected examples of phosphane ligands obtained by Cai’s method.
Molecules 26 05283 g009
Figure 10. Nickel-catalyzed cross-coupling between aryl bromides and diphenylphosphane.
Figure 10. Nickel-catalyzed cross-coupling between aryl bromides and diphenylphosphane.
Molecules 26 05283 g010
Figure 11. Reactivity of organonickel sigma-complexes with phosphanes.
Figure 11. Reactivity of organonickel sigma-complexes with phosphanes.
Molecules 26 05283 g011
Figure 12. Hydrophosphination of weakly activated olefins and their heterocycle-containing analogues.
Figure 12. Hydrophosphination of weakly activated olefins and their heterocycle-containing analogues.
Molecules 26 05283 g012
Figure 13. Mechanism of a Ni-catalyzed hydrophosphination of alkenes.
Figure 13. Mechanism of a Ni-catalyzed hydrophosphination of alkenes.
Molecules 26 05283 g013
Figure 14. Ni-catalyzed asymmetric hydrophosphination of methacrylonitrile.
Figure 14. Ni-catalyzed asymmetric hydrophosphination of methacrylonitrile.
Molecules 26 05283 g014
Figure 15. Nickel-catalyzed addition of HPPh2 to alkynes.
Figure 15. Nickel-catalyzed addition of HPPh2 to alkynes.
Molecules 26 05283 g015
Figure 16. Nickel-catalyzed addition of R2P(O)H to alkynes.
Figure 16. Nickel-catalyzed addition of R2P(O)H to alkynes.
Molecules 26 05283 g016
Figure 17. Nickel-catalyzed quaternization of triphenylphosphane.
Figure 17. Nickel-catalyzed quaternization of triphenylphosphane.
Molecules 26 05283 g017
Figure 18. Scope of the Ni-catalyzed C‒P coupling.
Figure 18. Scope of the Ni-catalyzed C‒P coupling.
Molecules 26 05283 g018
Figure 19. Nickel-catalyzed coupling of aryl chlorides with a diphenyl phosphite.
Figure 19. Nickel-catalyzed coupling of aryl chlorides with a diphenyl phosphite.
Molecules 26 05283 g019
Figure 20. Nickel-catalyzed coupling of arylboronic acids with P-nucleophiles.
Figure 20. Nickel-catalyzed coupling of arylboronic acids with P-nucleophiles.
Molecules 26 05283 g020
Figure 21. Reaction of alkenyl bromides with diphenylphosphane oxide.
Figure 21. Reaction of alkenyl bromides with diphenylphosphane oxide.
Molecules 26 05283 g021
Figure 22. Decarbonylative coupling of esters with phosphane oxide.
Figure 22. Decarbonylative coupling of esters with phosphane oxide.
Molecules 26 05283 g022
Figure 23. Ni-catalyzed asymmetric allylation of secondary phosphane oxides.
Figure 23. Ni-catalyzed asymmetric allylation of secondary phosphane oxides.
Molecules 26 05283 g023
Figure 24. Ni-catalyzed reaction of phosphonic ethers with aryl bromides.
Figure 24. Ni-catalyzed reaction of phosphonic ethers with aryl bromides.
Molecules 26 05283 g024
Figure 25. C‒P cross-coupling using aryl tosylates or mesylates.
Figure 25. C‒P cross-coupling using aryl tosylates or mesylates.
Molecules 26 05283 g025
Figure 26. Reaction of phenylboronic acid with ethyl phenylphosphinate.
Figure 26. Reaction of phenylboronic acid with ethyl phenylphosphinate.
Molecules 26 05283 g026
Figure 27. Cross-coupling reactions of aryl halides with triethyl phosphites and diethyl phenylphosphonites.
Figure 27. Cross-coupling reactions of aryl halides with triethyl phosphites and diethyl phenylphosphonites.
Molecules 26 05283 g027
Figure 28. The catalytic arylation of tris(trimethylsilyl)phosphite.
Figure 28. The catalytic arylation of tris(trimethylsilyl)phosphite.
Molecules 26 05283 g028
Figure 29. The reactions of triethylphosphite with alkenyl halides having an alkoxy or diethylamino group.
Figure 29. The reactions of triethylphosphite with alkenyl halides having an alkoxy or diethylamino group.
Molecules 26 05283 g029
Figure 30. Proposed mechanism for the Ni-catalyzed cross-coupling of trialkyl phosphites with aryl iodides.
Figure 30. Proposed mechanism for the Ni-catalyzed cross-coupling of trialkyl phosphites with aryl iodides.
Molecules 26 05283 g030
Figure 31. Ni-catalyzed reaction of triethylphosphite with 2-haloanilides.
Figure 31. Ni-catalyzed reaction of triethylphosphite with 2-haloanilides.
Molecules 26 05283 g031
Figure 32. Nickel-catalyzed C‒P cross-coupling of aryl triflates with triethyl phosphite.
Figure 32. Nickel-catalyzed C‒P cross-coupling of aryl triflates with triethyl phosphite.
Molecules 26 05283 g032
Figure 33. Cross-coupling of dimethyl phosphite with various aryl halides.
Figure 33. Cross-coupling of dimethyl phosphite with various aryl halides.
Molecules 26 05283 g033
Figure 34. Hirao reaction after activating the hydroxyl group of phenols.
Figure 34. Hirao reaction after activating the hydroxyl group of phenols.
Molecules 26 05283 g034
Figure 35. Decarboxylative cross-coupling of alkenyl acids with P(O)H compounds.
Figure 35. Decarboxylative cross-coupling of alkenyl acids with P(O)H compounds.
Molecules 26 05283 g035
Figure 36. Phosphorylation of amides using Ni catalyst.
Figure 36. Phosphorylation of amides using Ni catalyst.
Molecules 26 05283 g036
Figure 37. NiCl2-catalyzed phosphonylation.
Figure 37. NiCl2-catalyzed phosphonylation.
Molecules 26 05283 g037
Figure 38. Phosphorylation of aromatic compounds in electrochemical conditions.
Figure 38. Phosphorylation of aromatic compounds in electrochemical conditions.
Molecules 26 05283 g038
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Zagidullin, A.A.; Sakhapov, I.F.; Miluykov, V.A.; Yakhvarov, D.G. Nickel Complexes in C‒P Bond Formation. Molecules 2021, 26, 5283. https://0-doi-org.brum.beds.ac.uk/10.3390/molecules26175283

AMA Style

Zagidullin AA, Sakhapov IF, Miluykov VA, Yakhvarov DG. Nickel Complexes in C‒P Bond Formation. Molecules. 2021; 26(17):5283. https://0-doi-org.brum.beds.ac.uk/10.3390/molecules26175283

Chicago/Turabian Style

Zagidullin, Almaz A., Il’yas F. Sakhapov, Vasili A. Miluykov, and Dmitry G. Yakhvarov. 2021. "Nickel Complexes in C‒P Bond Formation" Molecules 26, no. 17: 5283. https://0-doi-org.brum.beds.ac.uk/10.3390/molecules26175283

Article Metrics

Back to TopTop