O,P-Acetals (Update 2016)

30.2.3 O,P-Acetals (Update 2016)


K. Murai and H. Fujioka


General Introduction


Syntheses of O,P-acetals, defined as phosphorus compounds containing a P—C—O moiety, reported after 2006 are introduced in this review. For previously published material, see Section 30.2.1.


30.2.3.1 Method 1: Addition of Phosphorus Compounds to Ketones or Aldehydes


α-Acetoxy phosphonates are prepared by reacting aldehydes with diethyl phosphite in the presence of acetic anhydride under solvent-free conditions using solid bases. The reactions of dialkyl phosphites with aldehydes in the presence of acetic anhydride and potassium carbonate under solvent-free conditions produce α-acetoxy phosphonates in one pot. This method is operationally simple, rapid, and generally gives high yields. Alkali metal carbonates (K2CO3 and Na2CO3) give better results than other solid bases such as magnesium oxide, calcium oxide, barium oxide, and alumina.[1,2] Catalysts such as molybdenum(VI) dichloride dioxide (MoO2Cl2) or potassium phosphate (K3PO4) also effectively promote the hydrophosphonylation of aldehydes under solvent-free conditions to give α-hydroxy phosphonates in good yields ( Scheme 1).[3,4]



The formation of α-hydroxy phosphonates from aldehydes and triethyl phosphite is accelerated by ultrasound under solvent-free conditions in the presence of potassium dihydrogen phosphate (KH2PO4) ( Scheme 2). A variety of aromatic, heteroaromatic, and α,β-unsaturated aldehydes can be used for this method, whereas low yields are obtained with aliphatic aldehydes, and no conversion occurs with ketones.[5]























R1 Conditions Time (min) Yield (%) Ref
Ph ultrasound 5 86 [5]
Ph no ultrasound 45 80 [5]

α-Hydroxy phosphonates 1, derived from heterocyclic aldehydes or six-membered heterocyclic ketones, are effectively synthesized under solvent-free conditions using microwave irradiation ( Scheme 3). A wide range of heterocyclic compounds, including thiazoles, pyridines, and diazines, are tolerated on solid magnesium oxide supports and give satisfactory yields. The reaction yield is affected by the steric bulk of the phosphite used; see, for example, the reactions of aldehydes 2 and 3 ( Scheme 3).[6]



































Ar1 R1 Yield (%) Ref
missing link H 83 [6]
missing link H 87 [6]
missing link H 84 [6]
missing link Me 80 [6]
missing link Me 80 [6]

missing link






















R1 R2 Yield (%) Ref
Me Me 74 [6]
iPr iPr 72 [6]
CH2CMe2CHPh 57 [6]























R1 R2 Yield (%) Ref
Me Me 76 [6]
iPr iPr 61 [6]
CH2CMe2HPh 0 [6]

Hydrophosphonylation of ketones is promoted by titanium(IV) isopropoxide under solvent-free conditions to produce quaternary α-hydroxy phosphonates efficiently in high yields ( Scheme 4).[7] The substrate scope is broad, and functional groups of various types are tolerated. These are the first examples of ketone hydrophosphonylation using Lewis acid catalysts, although such catalysts are effective for hydrophosphonylation of aldehydes. Retro-hydrophosphonylation reactions and phospha-Brook reactions, which are often problematic with base-catalyzed ketone hydrophosphonylation, are avoided under these conditions.






















































R1 R2 Yield (%) Ref
Ph Me 98 [7]
missing link 95 [7]
missing link Me 93 [7]
Cy Me 81 [7]
4-ClC6H4 CH(OEt)2 87 [7]
Ph (CH2)2Cl 86 [7]
Ph CO2Me 86 [7]
CH(Bn)CO2Et Me 93 [7]
missing link Me 89 [7]

Butyllithium serves as an efficient precatalyst for the hydrophosphonylation of aldehydes and unactivated ketones with dialkyl phosphites ( Scheme 5).[8] Reduction of the catalyst loading (0.1 mol%) suppresses the reverse reaction (elimination of the dialkyl phosphite from the product) with ketone substrates. A broad range of substrates are tolerated to generate a series of α-hydroxy phosphonates in high yields in shortreaction times (generally within 5 min).

































































































































R1 R2 R3 Yield (%) Ref
Ph H Et >99 [8]
2-MeOC6H4 H Et >99 [8]
2-ClC6H4 H Et >99 [8]
3-O2NC6H4 H Et >99 [8]
1-naphthyl H Et >99 [8]
2-furyl H Et 90 [8]
Pr H Et 97 [8]
Ph H iPr 99 [8]
Ph Me Et 95 [8]
2-ClC6H4 Me Et 79 [8]
2-BrC6H4 Me Et 59 [8]
3-O2NC6H4 Me Et 99 [8]
3-pyridyl Me Et 96 [8]
Ph Ph Et 97 [8]
2-furyl Me Et 69 [8]
Ph (CH2)10Me Et 40 [8]
Ph Bz Et 55 [8]
missing link Et 93 [8]
(CH2)5 Et 95 [8]
Ph Me iPr 85 [8]

Several lanthanide complexes have been used as highly efficient catalysts for hydrophosphonylation reactions ( Scheme 6).[911] A lanthanide amide {[(TMS)2N)]3La(μ-Cl)Li(THF)3}[9]and amino-coordinated lithium-bridged bis(indolyl)lanthanide amide 4[10] catalyze the hydrophosphonylation of aldehydes with diethyl or diphenyl phosphite. Reactions using 4 are performed neat and high yields (ca. 99%) are achieved. Lanthanide anilido complex 5 catalyzes the hydrophosphonylation of ketones, as well as aldehydes, with diethyl phosphite in hexane.[11] Under these reaction conditions, a low catalyst loading (0.1 mol%) is sufficient and excellent yields (ca. 99%) are obtained.



Ammonium metavanadate (NH4VO3), which is a water-soluble inorganic acid, also promotes the hydrophosphonylation of aryl- or hetarylaldehydes with triethyl phosphite ( Scheme 7; 18 examples, up to 94% yield). This method involves solvent-free conditions at room temperature and short reaction times. Aliphatic aldehydes and aromatic ketones are not converted by this reaction. In these cases no reaction occurs, even on increasing the concentration of the catalyst.[12]



Immobilized catalysts have been developed for hydrophosphonylation reactions. Such methods are environmentally benign because of the solvent-free conditions and simple workup.


The first example of this type of catalyst is polymer-bound phosphazene base 6, 2-(tert-butylimino)-2-(diethylamino)-1,3-dimethyl-1,3,2-diazaphosphinane supported on polystyrene (PS-BEMP); it promotes hydrophosphonylation of aromatic and aliphatic aldehydes. The process can be performed on a large scale using a continuous-flow procedure ( Scheme 8).[13]



The second example is magnetic nanoparticle (γ-Fe2O3) immobilized 1,5,7-triazabicyclo[4.4.0]dec-5-ene 8 (MNP-TBD). Oxindolyl α-hydroxy phosphonates are synthesized from isatins 7 and dimethyl or diethyl phosphite; the substrate scope is wide ( Scheme 9). The catalyst is easily recovered using an external magnet and can be reused (6 times) without any significant loss of activity.[14]



Diethyl Hydroxy(4-methylthiazol-5-yl)methylphosphonate (1, Ar1 = 4-Methylthiazol-5-yl; R1 = H); Typical Procedure:[6]

4-Methylthiazole-5-carbaldehyde (0.381 g, 3 mmol), diethyl phosphite (0.414 g, 3 mmol), and MgO (0.484 g, 12 mmol) were mixed and exposed to microwave irradiation. The heating was intermittent (every 2 min), and the mixture was stirred. The reaction was monitored by TLC, and was stopped at 6 min. Then, CHCl3 (2 × 10 mL) was poured into the container and MgO was removed by filtration. The solvent was evaporated under reduced pressure and the residue was recrystallized (EtOAc/petroleum ether) to afford the pure product as a white solid; yield: 83%.


30.2.3.1.1 Variation 1: Diastereoselective Hydrophosphonylation


The reactions of optically pure chiral 2-azanorbornane aldehydes exo9 and endo9 with silylated phosphorus esters are highly diastereoselective. (S)-α-Hydroxy phosphonic acid 10A is obtained from exo9, and the R-diastereomer 10B is obtained from endo9 ( Scheme 10).[15]



Highly diastereoselective hydrophosphonylation is observed in the reactions of tetraaryl-2,2-dimethyldioxolane-4,5-dimethanol (TADDOL) derived H-phosphonate 11 and aldehydes. The chiral auxiliary can be removed by two different procedures to give (α-hydroxyalkyl)phosphonic acids in good yields ( Scheme 11).[16]















































R1 Conditions dr Yield (%) Ref
Me Et2Zn (1.1 equiv), TMEDA (1.2 equiv) 92:8 90 [16]
Et Et2Zn (1.1 equiv), TMEDA (1.2 equiv) 95:5 89 [16]
Ph LDA (1 equiv) 92:8 91 [16]
4-MeOC6H4 LDA (1 equiv) 92:8 87 [16]
3-pyridyl LDA (1 equiv) 89:11 86 [16]
2-thienyl LDA (1 equiv) 90:10 87 [16]

missing link




























R1 Conditions Yield (%) Ref
Et 1. TMSBr, rt 2. MeOH 90 [16]
Ph 1. TMSBr, rt 2. MeOH 91 [16]
Et aq HCl, toluene, reflux 92 [16]
Ph aq HCl, toluene, reflux 95 [16]

30.2.3.1.2 Variation 2: Enantioselective, Metal-Catalyzed Addition of Phosphites to Aldehydes (Pudovik Reaction)


The optically active aluminum(salalen) complex 12, bearing a new optically active salalen ligand derived from (1R,2R)-cyclohexane-1,2-diamine, catalyzes enantioselective hydrophosphonylation of aldehydes to give the corresponding α-hydroxy phosphonates 14 with high enantioselectivities; however, the reaction is slow, and a high catalyst loading of 10 mol% and long reaction times are required.[17] Aluminum(salalen) 13, which has tert-hexyl groups in place of tert-butyl groups, was developed in the same laboratory. Catalyst 13 needs a lower catalyst loading and gives improved yields, product enantioselectivities, and reaction times ( Scheme 12).[18]



missing link






























































































R1 Catalyst (mol%) Conditions ee(%) Yield (%) Ref
Ph 12 (10) THF, –15 °C, 48 h 90 87 [17]
Ph 13 (1) Et2O,–30°C, 24 h 97 99 [18]
4-O2NC6H4 12 (10) THF, –15 °C, 48 h 94 95 [17]
4-O2NC6H4 13 (2) Et2O, –30°C, 24 h 98 98 [18]
4-MeOC6H4 12 (10) THF, –15°C, 48 h 81 87 [17]
4-MeOC6H4 13 (2) Et2O, –30 °C, 24 h 93 98 [18]
(E)-CH=CHPh 12 (10) THF, –15°C, 48 h 83 77 [17]
(E)-CH=CHPh 13 (2) Et2O, –30 °C, 24 h 95 97 [18]
(CH2)2Ph 12 (10) THF, –15 °C, 48 h 91 94 [17]
(CH2)2Ph 13 (2) Et2O, –30 °C, 24 h 97 93 [18]
iPr 12 (10) THF, –15 °C, 48 h 89 89 [17]
iPr 13 (2) Et2, –30 °C, 24 h 96 96 [18]

Catalysts with chiral binaphthyl skeletons also have significant potential. The aluminum–binaphthyl Schiff base complex 15 obtained from 1,1′-binaphthalene-2,2′-diamine is a good catalyst for enantioselective hydrophosphonylation of aldehydes, and high selectivities are obtained in the reactions of both aromatic and aliphatic aldehydes ( Scheme 13).[19]



A bifunctional chiral aluminum complex of 1,1′-binaphthalene-2,2′-diol (BINOL) derivative 16, which has tertiary amine groups at the 3- and 3′-positions of BINOL, promotes enantioselective hydrophosphonylation of aldehydes. The enantioselectivity is usually moderate to good (up to 87% ee), and a wide variety of aromatic, heteroaromatic, α,β-unsaturated, and aliphatic aldehydes can be used in this method ( Scheme 14).[20]



































R1 ee (%) Yield (%) Ref
Ph 75 88 [20]
4-MeOC6H4 87 74 [20]
2-thienyl 45 75 [20]
(E)-CH=CHPh 63 86 [20]
Bu 77 85 [20]

The metal–organic self-assembly of substituted binaphthalenediols 17/18 and cinchona alkaloid 19 in combination with titanium(IV) isopropoxide catalyzes asymmetric hydrophosphonylation of aldehydes with high enantioselectivities. BINOL ligand 17 is the best choice for aromatic aldehydes, whereas BINOL 18 gives the best results for aliphatic aldehydes ( Scheme 15).[21]



missing link





































































R1 BINOL Ligand ee (%) Yield (%) Ref
Ph 17 99 99 [21]
4-Tol 17 97 96 [21]
4-ClC6H4 17 92 99 [21]
4-NCC6H4 17 91 99 [21]
4-O2NC6H4 17 93 99 [21]
(E)-CH=CHPh 17 89 97 [21]
(CH2)2Ph 18 92 95 [21]
Cy 18 92 95 [21]
(CH2)7Me 18 94 97 [21]
iPr 18 94 94 [21]

Enantioselective Pudovik reactions of bis(2,2,2-trifluoroethyl)phosphite with aldehydes are catalyzed by the tethered bis(8-quinolinolato) (TBOx) aluminum complex 20 ( Scheme 16).[22] The reaction proceeds even with a low catalyst loading (0.5–1 mol%). The obtained chiral α-hydroxy phosphonates 21 can be converted into α-hydroxy phosphonic acids without loss of enantiopurity.


Only gold members can continue reading. Log In or Register to continue

Stay updated, free articles. Join our Telegram channel

Jun 14, 2017 | Posted by in GENERAL SURGERY | Comments Off on O,P-Acetals (Update 2016)

Full access? Get Clinical Tree

Get Clinical Tree app for offline access