Next Article in Journal
New 1-Aryl-3-Substituted Propanol Derivatives as Antimalarial Agents
Previous Article in Journal
A New Organofunctional Ethoxysilane Self-Assembly Monolayer for Promoting Adhesion of Rubber to Aluminum
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Addition and Cycloaddition Reactions of Phosphinyl- and Phosphonyl-2H-Azirines, Nitrosoalkenes and Azoalkenes

CIQA, Faculdade de Ciências e Tecnologia, Universidade do Algarve, Campus de Gambelas, 8005-137 Faro, Portugal
Molecules 2009, 14(10), 4098-4119; https://doi.org/10.3390/molecules14104098
Submission received: 25 August 2009 / Revised: 23 September 2009 / Accepted: 12 October 2009 / Published: 13 October 2009
(This article belongs to the Special Issue Alpha- or Beta-aminophosphonates and -phosphinates)

Abstract

:
An overview of the use of 2H-azirines, conjugated nitrosoalkenes and conjugated azoalkenes bearing phosphorus substituents in addition and cycloaddition reactions is presented, focused on strategies for the synthesis of aminophosphonate and aminophosphine oxide derivatives.

1. Introduction

Over recent years we and others have investigated the use of 2H-azirines, conjugated nitrosoalkenes and conjugated azoalkenes in nucleophilic addition and cycloaddition reactions. The structures I, II and III of these three classes are outlined in Figure 1.
A common feature of all three structures is that they possess a highly electrophilic carbon centre (C-3 in 2H-azirines, C-4 in conjugated nitroso- and azo-alkenes) that allows nucleophilic addition reactions to proceed very readily. In reactions of nucleophiles with 2H-azirines this often leads eventually to opening of the three membered ring. The electrophilic character of these structures also allows cycloaddition reactions, particularly those with nucleophilic olefins, to take place under very mild conditions. Such reactions have provided routes to a variety of novel heterocyclic structures which have proved to be very useful targets, not only due to their eventual biological and pharmacological properties, but especially to their wide and versatile use as synthetic intermediates or useful building blocks for the synthesis of amino acids, pyrroles, proline, indoles, pyrazines, and aza-sugars derivatives, amongst many other compounds [1,2,3,4,5,6,7,8,9,10,11].
Aminophosphonic and aminophosphinic acid derivatives can be considered as isosteres or surrogates of aminocarboxylic acids and they regulate various important biological functions [12,13,14,15,16]. In this context it is not surprising that organic chemists have been attracted to them and have paid particular attention to the synthesis of these types of compounds. The aim of this review is to illustrate the particular use of above reaction types of 2H-azirines, nitrosoalkenes and azoalkenes bearing phosphinyl or phosphonyl substituents, for the construction of alkyl α- and β-aminophosphonates and aminoalkylphosphine oxides.

2. 2H-Azirines

2H-Azirines are strained and activated imines. Their high reactivity makes them very useful synthetic intermediates for the synthesis of aziridines, amino acids, indoles, pyrazines, and other biologically active compounds through cycloaddition and nucleophilic addition reactions [3,5,6,8,9,17].

2.1. Synthesis

Despite all these potential applications, 2H-azirines bearing phosphorus substituents have received comparatively little attention. Photocyclization of vinyl azides [18], reaction of phosphites with β-nitrostyrenes [19] and carbene addition to aromatic nitriles [20,21] constituted the earlier examples of 2H-azirines with a phosphinyl or phosphonyl functional group. Afterwards, a diverse methodology based on Swern oxidation of chiral aziridines 1 and 2, produced regioisomeric mixtures of azirinyl phosphonates 3-5 (Scheme 1) [22,23].
Thermolysis of vinyl azide 6 [24] allowed the isolation of diphenylphosphinyl 2H-azirine 7 in good yield (Scheme 2), but this strategy was not suitable for the preparation of enantiopure azirines.
The asymmetric synthesis of 2H-azirines bearing phosphinyl [24] and phosphonyl [25] substituents was disclosed by alkaloid mediated Neber reactions of β-keto tosyloximes. Similarly, the use of chiral polymer-supported bases [26] led to 2H-azirines 9 regioselectively and in high yields (Scheme 3).
Another approach based on the treatment of phosphorylated allenes 11 with sodium azide was the basis of a convenient methodology for the synthesis of 3-vinyl- and 3-dihydroisoxazolyl-2H-azirines 13 and 16 (Scheme 4) [27,28].

2.2. Addition reactions

One of the earliest reported reactions of 2H-azirines bearing phosphorus substituents was hydride addition [24,25]. The treatment of azirines 9, 10 with sodium borohydride in ethanol produced cis-aziridines exclusively (Scheme 5). The stereochemical assignment was based on the large coupling constant observed for the ring protons and further established by the transformation into enantiopure cis-N-(p-toluenesulfinyl)-aziridines by treatment with (-)-(S)-menthyl p-toluenesulfinate [24].
For the synthesis of β-amino-phosphine oxide and –phosphonate derivatives 22 from tosyloximes 19, a similar addition of hydride takes place with 3-fluoroalkyl-2H-azirines 20–postulated as plausible intermediates – producing regioselectively cis-aziridines 21 which then lead to compounds 22 by ring opening [29] (Scheme 6).
Nitrogen heterocycles, in the presence or absence of base, add regioselectively to the azirine nucleus following the general pattern – the attack being from the less hindered face of the azirine - yielding functionalized aziridines [29,30] (Scheme 7 and Scheme 8).
Oxygen [30] (Scheme 9) and sulfur [29] (Scheme 10) nucleophiles also add in a similar and regioselective mode, to 2H-azirines bearing phosphorus substituents. In the case of the reaction with benzenethiol, if a methyl group is present in the ring of the resulting aziridines, subsequent ring opening reaction leads to a-aminophospine oxide and –phosphonates 33.
The addition of Grignard reagents to 2-phosphinyl- and 2-phosphonyl-2H-azirines is less simple. Early reports with 2,3-diphenyl-2H-azirine revealed that the reaction followed the general pattern of addition of nucleophiles, i.e., the obtained aziridines arise from the attack at the less hindered face of the azirine [31]. These findings are in clear contrast with those obtained with alkyl 2H-azirine-2-carboxylates, in which the syn addition –to the more hindered face– is preferred (Scheme 11) [32,33]. These facts have been ascribed to a prechelating effect of the Grignard reagents with carboxylate substituents.
When the carboxylate group was replaced by a phosphoryl group, the reverse preference was observed, i.e., an exclusive attack at the least hindered side was encountered [30] (Table 1).
This behaviour has been ascribed to the high exocyclic dihedral angle of the saturated carbon and to the presence of a bulky tetrahedral phosphorus group. But if a chelating substituent, such as alkylfluoromethyl or perfluoroalkylmethyl is present beside the phosphorus group, the former may play a major role in the mode of addition and in the reaction outcome, as demonstrated by the production of mixtures of cis/trans aziridines 43/44 [29] (Scheme 12).
Carboxylic acids [26], N-protected aminoacids and peptide residues [34] also add to the carbon nitrogen double bond of phosphinyl- and phosphonyl-2H-azirines. The concomitant ring opening leads to ketamides 48 (Scheme 13).
Due to their ambident character, phosphinyl- and phosphonyl-2H-azirines also react as nucleophiles with carboxylic acid derivatives, such as acid chlorides, producing exclusively trans-aziridines 50 [35]. The scope of the reaction is not limited to simple chlorides since other functionalized acyl chlorides will react similarly in good overall yields.

2.3. Cycloadditions

Simple alkyl- and aryl-2H-azirines, although being more reactive than acyclic imines, participate in Diels-Alder reactions only with highly reactive dienes, such as cyclopentadienones and 1,3-diphenylisobenzofuran in refluxing toluene [36], or with acyclic dienes and cyclopentadiene under Lewis acid catalysis [37,38]. 2H-Azirines with an alkoxy-, aryl-, amino-carbonyl [8,39] or heteroaromatic [40] substituent on the C=N bond are particularly good dienophiles in Diels–Alder reactions with a great variety of dienes, as a consequence of the conjugated effect of ring strain and extra activation by the electron-withdrawing group.
Similarly enantiomerically enriched 2H-azirine-3-phosphonates 51 when stirred with 100 equiv of 2,3-dimethylbutadiene or trans-piperylene for 2-4 days at room temperature or with Danishefsky’s diene for 8 hours, afforded bicyclic aziridines 53 as single stereoisomers in good yields [23] (Scheme 15).
The stereochemistry of cycloadducts 53 was consistent with exclusive addition of the diene to the less hindered face of the azirine 51. The longer reaction times, when compared with 2H-azirine-3-carboxylates, may suggest that 2H-azirine-3-phosphonates are less reactive than carboxylates.
Recently an azirine bearing both ethoxycarbonyl and phosphonate groups, was generated in situ and intercepted with a number of nucleophilic dienes [41]. With open chain dienes bicyclic functionalized six-membered ring fused aziridines were produced; although cyclic dienes afforded tryciclic structures. The presence of a trimethylsilyloxy group at the conjugated system, induced hydrolysis of cycloadduct 56f and 56c to 57 and 58 respectively (Scheme 16).
The products were isolated as single isomers, presumed to be formed by endo selective processes, as clearly indicated by the low field resonance of H-3 in the tricyclic structure 56d, attributed to the anisotropy of the backside double bond over H-3, due to constrain of the tricyclic structure [42]. To the poor stability of azirine 55 was ascribed the low to moderate yields of cycloadducts (Scheme 16).

3. Nitroso- and Azo-alkenes

Nitrosoalkenes and azoalkenes, used either as Michael-type acceptors in conjugate 1,4-additions, or as heterodienes in cycloaddition reactions with a range of nucleophiles, alkenes and heterocycles, have proved to be invaluable tools for the synthetic organic chemists.

3.1. Synthesis or generation

Electron-deficient nitrosoalkenes are, generally, very unstable species and for this reason, they are usually generated and intercepted in situ. Depending on the substituents, azoalkenes are sometimes stable enough to be isolated. Anyway, being isolated or generated “in situ”, the most common, general and broad scope method for the obtention of nitroso- and azo-alkenes is the base induced 1,4-dehydro elimination from oximes and hydrazones bearing a suitable halogen or ester leaving group at the α-position [1,2,4,7,10,11] (Scheme 17).

3.2. Reactions with nucleophiles

Although in reactions involving nitroso- and azo-alkenes generated and intercepted in situ, the distinction between nucleophilic substitution of the original α-halogenated oxime or hydrazone and 1,4-conjugate (or Michael type) addition, is sometimes an intricate decision, the following reactions are thought to proceed via this latter process.
The primary literature reference to azoalkenes bearing phosphorus substituents [43] reported their use at the synthesis of 1-aminopyrroles substituted with a phosphine oxide or phosphonate group in the 3-position.
Achiral and chiral phosphinyl- and phosphonyl-1,2-diaza-1,3-butadienes obtained from hydrazonoalkyl-phosphine oxides and –phosphonates, were reported to add ammonia, aminoesters and aminoalcohols giving functionalized a-amino-phosphine oxides and –phosphonates [44,45] (Scheme 18). Very low diastereoselection with optically active amines was encountered - the adducts were isolated as nonseparable diastereoisomeric mixtures. Better diastereoselection was found when the bulky (S)-tert-leucinol substituent was used.
The diastereoselective addition of optically active amino esters gave slightly improved results when an optically active group was introduced at position 3 of the azoalkene. Thus the azovinylphosphonate 68 produced hydrazono derivatives 70 in good yields but moderate diastereoselectivity [46] (Scheme 19).
Amino- and alkoxy-carbonyl 1,2-diaza-1,3-dienes underwent 1,4-addtion with sulfur nucleophiles such as 1,2-aminothiols [47] (Scheme 20), thiourea [48] (Scheme 21) and 3-mercapto-2-ketones [49] (Scheme 22) producing either functionalized hydrazones or ensuing cyclised heterocycles.
Similarly, hydrazones 78a,b arised from the 1,4-addition of oxygen nucleophiles to 1,2-diaza-1,3-dienes [48] (Scheme 21).
Under solvent free or solid-phase conditions, a Michael type addition of 1,2-diamines to phosphinyl- and phosphonyl-azoalkenes was observed, furnishing substituted hydrazones which subsequently were cyclised into pyridazines [50] (Scheme 23).
The nitrogen heteroaromatic N-methylimidazole was reported to promote dehydrohalogenation of hydrazone 87 and to add regioselectively to the azo-alkene 89, conducting to functionalized α-hydrazonophosphonate 91 [51] (Scheme 24).
Several examples of nucleophilic conjugate addition to nitrosoalkenes are reported in the literature [4,10,52,53,54], whereas reports of related reactions of phosphinyl- and phosphonyl-nitrosoalkenes are scarcely described. Nitrososalkenes 92, generated “in situ” from the corresponding α-bromooxime, were reported to act as Michael acceptors towards ammonia, amines and optically active amino esters, affording α-amino phosphine oxides and α-amino phosphonates [55] (Scheme 25). The reactions, although being regioselective, lacked stereoselectivity with optically active compounds.
There are various works reporting reactions of conjugated nitrosoalkenes with enamines [52,56,57,58,59,60] affording 1,2-oxazines in good to excellent yields via postulated Diels-Alder reactions with inverse electron demand. A more recent work [61] using density functional theory (DFT) studies, showed a polar character of the Diels–Alder reaction of nitrosoalkenes with enamines, being the two σ bonds formed in these polar cycloaddition reactions in a highly asynchronous concerted process. On the other hand reactions of azoalkenes with enamines are solvent and structure – of enamine and azoalkene - dependent [1,62,63,64,65,66,67,68,69], proceeding either via [4+2] cycloaddition or conjugate addition.
In this context, Palacios and co-workers [70] found that the reactions of nitrosoalkenes, bearing a phosphinyl or phosphonyl substituent at the terminal carbon, with enamines 99 did not proceed by [4+2] cycloaddition reactions producing the expected 1,2-oxazines, but N-hydroxypyrroles were instead isolated, presumably by a mechanism involving an initial conjugate addition of the enamine followed by a formal [3+2] dipolar cycloaddition (Scheme 26).

3.3. Cycloadditions

Cycloaddition reactions of nitroso- and azo-alkenes with electron rich carbon-carbon double bonds and heterocycles are important and powerful synthetic tools, as demonstrated by the enormous impetus of the chemistry of these compounds over the last decades.
Azoalkenes with a phosphinyl or phosphonyl substituent at the 4-position were generated by triethylamine induced dehydrohalogenation of chlorohydrazones [71]. Its interception with acyclic— styrene and cyclic—cyclopentadiene and norbornadiene alkenes and di-hydrofuran gave regio- and stereo-selectivelly endo cycloadducts, except in the case of the strained norbornadiene, which produced an exo cycloadduct (Scheme 27).
Attempts to bring about some degree of diastereoselectivity to the reaction, induced by the use of an optically active (-)-menthyl ester, were unsuccessful, since a diasteroisomeric ratio of 1:1 was found.
Cycloadducts were also obtained in regioselective fashion in reactions of azoalkenes bearing a phosphonate substituent at the 3-position with electron rich olefins [51] (Scheme 28).
Reactions with electron rich heterocycles such furan and dihydrofuran gave rise to the formation of the corresponding cycloadducts. With pyrrole and indole open chain hydrazones were isolated (Table 2), assumed to be the result of rearomatization of the primarily formed cycloadducts [72]. The yields were, in general, lower than those obtained in similar reactions with azoalkenes bearing the same alkoxycarbonylazo substituent, but having an ethoxycarbonyl group at the 3-position [73], pointing out that eventually the ethoxycarbonyl group may play a more effective role than the phosphonate moiety.
Nitrosovinylphosphonates generated from the corresponding chlorooximes were intercepted by electron-rich alkenes and heterocycles [74]. The yields, although not optimised, were fair and the reactions were found to be completely regioselective: no other isomers were detected or isolated from the reaction medium (Scheme 29).

4. Concluding Remarks

Synthetic applications of 2H-azirines, nitroso- and azo-alkenes bearing phosphinyl or phosphonyl substituents, emphasised towards the synthesis of α- and β-amino-phosphonates and –phosphine oxides have been addressed in this review. Established methods of aziridine ring opening [17,75] and reductive transformations at the C=N bond [76-78] will further broaden the scope and importance of these strategies that surely will be further developed and attract the attention and interest of synthetic chemists.

Acknowledgments

Special thanks are due to T. L. Gilchrist for helpful suggestions and discussions.

References and Notes

  1. Attanasi, O.A.; De Crescentini, L.; Favi, G.; Filippone, P.; Mantellini, F.; Perrulli, F.R.; Santeusanio, S. Cultivating the passion to build heterocycles from 1,2-diaza-1,3-dienes: The force of imagination. Eur. J. Org. Chem. 2009, 3109–3127. [Google Scholar] [CrossRef]
  2. Rai, K.M.L. Heterocycles via Oxime Cycloadditions. Top. Heterocycl. Chem. 2008, 13, 1–69. [Google Scholar]
  3. Palacios, F.; Ochoa de Retana, A.M.; Marigorta, E.M.; de los Santos, J.M. Addition Reactions to the Imine Bond of 2H-Azirines. Synthesis of Heterocyclic Compounds. In Recent Developments in Heterocyclic Chemistry; Pinho e Melo, T.M.V.D., Rocha-Gonsalves, A.M.A., Eds.; Research Signpost: Trivandrum, India, 2007; pp. 27–58. [Google Scholar]
  4. Reissig, H.U.; Zimmer, R. 1-Nitrosoalkenes. In Science of Synthesis; Molander, G.A., Ed.; Thieme: Stuttgard, Germany, 2006; Vol. 33, pp. 371–389. [Google Scholar]
  5. Pinho e Melo, T.M.V.D.; Rocha Gonsalves, A.M.A. Exploiting 2-halo-2H-Azirine chemistry. Curr. Org. Synth. 2004, 1, 275–292. [Google Scholar] [CrossRef]
  6. Palacios, F.; Ochoa de Retana, A.M.; Marigorta, E.M.; de los Santos, J.M. Preparation, properties and synthetic applications of 2H-Azirines. A review. Org. Prep. Proced. Int. 2002, 34, 219–269. [Google Scholar] [CrossRef]
  7. Attanasi, O.A.; De Crescentini, L.; Filippone, P.; Mantelilini, F.; Santeusanio, S. 1,2-Diaza-1,3-butadienes; just a nice class of compounds, or powerful tools in organic chemistry? Reviewing an experience. ARKIVOC 2002, 274–292. [Google Scholar]
  8. Gilchrist, T.L. Activated 2H-Azirines as Dienophiles and Electrophiles. Aldrichim. Acta 2001, 34, 51–55. [Google Scholar]
  9. Palacios, F.; Ochoa de Retana, A.M.; Marigorta, E.M.; de los Santos, J.M. 2H-Azirines as synthetic tools in organic chemistry. Eur. J. Org. Chem. 2001, 2401–2414. [Google Scholar] [CrossRef]
  10. Lyapkalo, I.M.; Ioffe, S.L. Conjugated Nitrosoalkenes. Russ. Chem. Rev. 1998, 67, 467–484. [Google Scholar] [CrossRef]
  11. Attanasi, O.A.; Filippone, P. Working twenty years on conjugated azo-alkenes (and environs) to find new entries in organic synthesis. Synlett 1997, 1128–1140. [Google Scholar] [CrossRef]
  12. Ordóñez, M.; Rojas-Cabrera, H.; Cativiela, C. An overview of stereoselective synthesis of α-aminophosphonic acids and derivatives. Tetrahedron 2009, 65, 17–49. [Google Scholar] [CrossRef] [PubMed]
  13. Palacios, F.; Alonso, C.; de los Santos, J.M. Synthesis of β-Aminophosphonates and -Phosphinates. Chem. Rev. 2005, 105, 899–931. [Google Scholar] [CrossRef] [PubMed]
  14. Moonen, K.; Laureyn, I.; Stevens, C.V. Synthetic methods for azaheterocyclic phosphonates and their biological activity. Chem. Rev. 2004, 104, 6177–6215. [Google Scholar] [CrossRef] [PubMed]
  15. Kafarski, P.; Lejczak, B. The Biological Activity of Phosphono- and Phosphino-peptides. In Aminophosphinic and Aminophosphonic Acids. Chemistry and Biological Activity; Kukhar, V.P., Hudson, H.R., Eds.; John Wiley & Sons Ltd: Chichester, UK, 2000; pp. 173–203, 407–442. [Google Scholar]
  16. Kafarski, P.; Lejczak, B. Biological activity of aminophosphonic acids. Phosphor. Sulfur Silicon 1991, 63, 193–215. [Google Scholar] [CrossRef]
  17. Singh, G.S.; D’hooghe, M.; De Kimpe, N. Synthesis and reactivity of C-heteroatom-substituted aziridines. Chem. Rev. 2007, 107, 2080–2135. [Google Scholar] [CrossRef] [PubMed]
  18. Christensen, B.G.; Beattie, T.R. Herstellung von Phosphonsäuren. Ger. Offen. 2,011,092, 1970. [Chem. Abstr. 1971, 74, 42491]. [Google Scholar]
  19. Russel, G.A.; Yao, C.F. Reactions of ethyl phosphites with β-nitrostyrenes. The role of nitrosoalkenes as intermediates. J. Org. Chem. 1992, 57, 6508–6513. [Google Scholar] [CrossRef]
  20. Alcaraz, G.; Wecker, U.; Baceiredo, A.; Dahan, F.; Bertrand, G. Synthesis of a 2H-Azirine by [1+2] cycloaddition of a phosphinocarbene with a nitrile and its ring-expansion to a 1,2λ5-Azaphosphete. Angew. Chem. Int. Ed. Engl. 1995, 34, 1246–1248. [Google Scholar] [CrossRef]
  21. Piquet, V.; Baceiredo, A.; Gornitzka, H.; Dahan, F.; Bertrand, G. The Stable (phosphino)(silyl)carbene as a useful building block: Synthesis and reactivity of 2-phosphorus-substituted 2H-Azirines. Chem. Eur. J. 1997, 3, 1757–1764. [Google Scholar] [CrossRef]
  22. Davis, F.A.; McCoull, W. Asymmetric synthesis of aziridine 2-phosphonates and azirinyl PHOSPHONATES from enantiopure sulfinimines. Tetrahedron Lett. 1999, 40, 249–252. [Google Scholar] [CrossRef]
  23. Davis, F.A.; Wu, Y.; Yan, H.; Prasad, K.R.; McCoull, W. 2H-Azirine 3-phosphonates: A new class of chiral iminodienophiles. asymmetric synthesis of quaternary piperidine phosphonates. Org. Lett. 2002, 4, 655–658. [Google Scholar] [CrossRef] [PubMed]
  24. Palacios, F.; Ochoa de Retana, A.M.; Gil, J.I.; Ezpeleta, J.M. Simple asymmetric synthesis of 2H-Azirines derived from phosphine oxides. J. Org. Chem. 2000, 65, 3213–3217. [Google Scholar] [CrossRef] [PubMed]
  25. Palacios, F.; Ochoa de Retana, A.M.; Gil, J.I. Easy and efficient synthesis of enantiomerically enriched 2H-azirines derived from phosphonates. Tetrahedron Lett. 2000, 41, 5363–5366. [Google Scholar] [CrossRef]
  26. Palacios, F.; Aparicio, D.; Ochoa de Retana, A.M.; de los Santos, J.M.; Gil, J.I.; Alonso, J.M. Asymmetric synthesis of 2H-Azirines derived from phosphine oxides using solid-supported amines. Ring opening of azirines with carboxylic acids. J. Org. Chem. 2002, 67, 7283–7288. [Google Scholar] [CrossRef] [PubMed]
  27. Brel, V.K. New synthetic route to 2-(Diethylphosphono)-2H-Azirines. Synthesis 2002, 1829–1832. [Google Scholar] [CrossRef]
  28. Brel, V.K. Synthesis of 3-azido-4-(diethoxyphosphoryl)alka-1,3-dienes and their transformation to derivatives of 2H-azirine. Synthesis 2007, 2674–2680. [Google Scholar] [CrossRef]
  29. Palacios, F.; Ochoa de Retana, A.M.; Alonso, J.M. Regioselective synthesis of fluoroalkylated β-aminophosphorus derivatives and aziridines from phosphorylated oximes and nucleophilic reagents. J. Org. Chem. 2006, 71, 6141–6148. [Google Scholar] [CrossRef] [PubMed]
  30. Palacios, F.; Ochoa de Retana, A.M.; Alonso, J.M. Reaction of 2H-Azirine phosphine oxide and -phosphonates with nucleophiles. Stereoselective synthesis of functionalized aziridines and α- and β-aminophosphorus derivatives. J. Org. Chem. 2005, 70, 8895–8901. [Google Scholar] [CrossRef] [PubMed]
  31. Carlson, R.M.; Lee, S.Y. Olefin synthesis via the stereospecific deamination of aziridines. Tetrahedron Lett. 1969, 10, 4001–4004. [Google Scholar] [CrossRef]
  32. Davis, F.A.; Liang, C.H.; Liu, H. Asymmetric synthesis of β-substituted α-amino acids using 2H-azirine-2-carboxylate esters. Synthesis of 3,3-disubstituted aziridine-2-carboxylate esters. J. Org. Chem. 1997, 62, 3796–3797. [Google Scholar] [CrossRef]
  33. Davis, F.A.; Deng, J.; Zhang, Y.; Haltiwanger, R.C. Aziridine-mediated asymmetric synthesis of quaternary β-amino acids using 2H-azirine 2-carboxylate esters. Tetrahedron 2002, 58, 7135–7143. [Google Scholar] [CrossRef]
  34. Palacios, F.; Ochoa de Retana, A.M.; Gil, J.I.; Alonso, J.M. Synthesis of optically active oxazoles and N-protected amino acids or peptides. Tetrahedron Asymmetry 2002, 13, 2541–2552. [Google Scholar] [CrossRef]
  35. Palacios, F.; Ochoa de Retana, A.M.; Gil, J.I.; Alonso, J.M. Regioselective synthesis of 4- and 5-oxazole-phosphine oxides and –phosphonates from 2H-azirines and acyl chlorides. Tetrahedron 2004, 60, 8937–8947. [Google Scholar] [CrossRef]
  36. Anderson, D.J.; Hassner, A. Synthesis of heterocycles via cycloadditions to 1-azirines. Synthesis 1975, 483–495. [Google Scholar] [CrossRef]
  37. Ray, C.A.; Risberg, E.; Somfai, P. Lewis acid-catalyzed hetero Diels–Alder cycloadditions of 3-alkyl, 3-phenyl and 3-carboxylated 2H-azirine. Tetrahedron Lett. 2001, 42, 9289–9291. [Google Scholar] [CrossRef]
  38. Ray, C.A.; Risberg, E.; Somfai, P. Diastereoselective Lewis acid-catalysed [4+2] cycloadditions of 3-alkyl-, 3-aryl- and 3-carboxyl-2H-azirines: A route to aziridine containing azabicyclo[4.1.0]heptanes and azatricyclo[2.2.1.0]nonanes. Tetrahedron 2002, 58, 5983–5987. [Google Scholar] [CrossRef]
  39. Gilchrist, T.L.; Mendonça, R. Synthesis and diels–alder reactions of 2H-azirine-3-carboxamides. ARKIVOC 2000, 769–778. [Google Scholar]
  40. Alves, M.J.; Fortes, A.G.; Lemos, A.; Martins, C. Ethyl 3-(2-pyridyl)-2H-azirine-2-carboxylate: synthesis and reaction with dienes. Synthesis 2005, 555–558. [Google Scholar]
  41. Alves, M.J.; Lemos, A.; Rodriguez-Borges, J.E.; García-Mera, X.; Fortes, A.G. Ethyl 2-(diisopropoxyphosphoryl)-2H-azirine-3-carboxylate: Reactions with nucleophilic 1,3-dienes. Synthesis 2009, 3263–3266. [Google Scholar] [CrossRef]
  42. Alves, M.J.; Gilchrist, T.L. Methyl 2-aryl-2H-azirine-3-carboxylates as dienophiles. J. Chem. Soc. Perkin 1 1998, 299–303. [Google Scholar] [CrossRef]
  43. Palacios, F.; Aparicio, D.; de 1os Santos, J.M. An efficient strategy for the regioselective synthesis of 3-phosphorylated-1-aminopyrroles from β-hydrazono phosphine oxides and phosphonates. Tetrahedron 1999, 55, 13767–13778. [Google Scholar] [CrossRef]
  44. Palacios, F.; Aparicio, D.; López, Y.; de los Santos, J.M. Addition of amine derivatives to phosphorylated 1,2-diaza-1,3-butadienes. Synthesis of α-aminophosphonates. Tetrahedron Lett. 2004, 45, 4345–4348. [Google Scholar] [CrossRef]
  45. Palacios, F.; Aparicio, D.; López, Y.; de los Santos, J.M.; Ignacio, R. An efficient synthesis of functionalized α-amino-phosphine oxides and -phosphonates by addition of aminoalcohols to 4-phosphorylated-1,2-diaza-1,3-butadienes. Arkivoc 2005, 153–161. [Google Scholar]
  46. Palacios, F.; Aparicio, D.; López, Y.; de los Santos, J.M. Synthesis of functionalized a-amino-phosphine oxides and -phosphonates by addition of amines and aminoesters to 4-phosphinyl- and 4-phosphonyl-1,2-diaza-1,3-butadienes. Tetrahedron 2005, 61, 2815–2830. [Google Scholar] [CrossRef]
  47. Attanasi, O.A.; Filippone, P.; Lillini, S.; Mantellini, F.; Nicolini, S.; de los Santos, J.M.; Ignacio, R.; Aparicio, D.; Palacios, F. Reactions of 1,2-diaza-1,3-dienes with thiol derivatives: A versatile construction of nitrogen/sulfur heterocycles. Tetrahedron 2008, 64, 9264–9274. [Google Scholar] [CrossRef]
  48. Attanasi, O.A.; Baccolini, G.; Boga, C.; De Crescentini, L.; Favi, G.; Filippone, P.; Mantellini, F. Carbon-phosphorus bond formation and transformation in the reaction of 1,2-diaza-1,3-butadienes with alkyl phenylphosphonites. Tetrahedron 2008, 64, 6724–6732. [Google Scholar] [CrossRef]
  49. Attanasi, O.A.; Filippone, P.; Lillini, S.; Mantellini, F.; de los Santos, J.M.; Ignacio, R.; Palacios, F. Domino reaction for the construction of new 2-Oxo[1,2,4]triazolo[5,1-c]thiazines. Synlett 2009, 735–738. [Google Scholar]
  50. 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] [PubMed]
  51. Lemos, A.; Lopes, M. Reactions of azovinylphosphonates with nucleophilic alkenes and heterocycles: Synthesis of tetrahydropyridazine-3-phosphonate and 2-substituted-1-hydrazonoethylphosphonate derivatives. Phosphor. Sulfur Silicon 2008, 183, 2882–2890. [Google Scholar] [CrossRef]
  52. Gilchrist, T.L; Lingham, D.A.; Roberts, T.G. Ethyl 3-bromo-2-hydroxyiminopropanoate, a reagent for the preparation of ethyl esters of α-amino acids. J. Chem. Soc. Chem. Commun. 1979, 1089–1090. [Google Scholar] [CrossRef]
  53. Gilchrist, T.L.; Stretch, W.; Chrystal, E.J.T. Reaction of azoles with ethyl bromopyruvate oxime: Alkylation by substitution and by elimination-addition. J. Chem. Soc. Perkin Trans. 1 1987, 2235–2239. [Google Scholar] [CrossRef]
  54. Zimmer, R.; Reissig, H.U. Efficient synthesis of trifluoromethyl-substituted 5,6-dihydro-4H-1,2-oxazines by the hetero-diels-alder reaction of l,l,l-trifluoro-2-nitroso-2-propene and electron-rich olefins. J. Org. Chem. 1992, 57, 339–347. [Google Scholar] [CrossRef]
  55. De los Santos, J.M.; Ignacio, R.; Aparicio, D.; Palacios, F. Michael addition of amine derivatives to conjugate phosphinyl and phosphonyl nitrosoalkenes. Preparation of α-amino phosphine oxide and phosphonate derivatives. J. Org. Chem. 2007, 72, 5202–5206. [Google Scholar] [CrossRef] [PubMed]
  56. Henning, R.; Lerch, U.; Urbach, H. Synthesis of esters of lipophilic proline analogs by reduction of ethyl 5,6- dihydro-4H-1,2-oxazine-3-carboxylates. Synthesis 1989, 265–268. [Google Scholar] [CrossRef]
  57. Paulini, K.; Reissig, H.U.; Rademacher, P. On the reactivity of N,N-bis(trimethyl)silylated enamines: Model studies on hetero Diels-Alder reactions and other cycloadditions. J. Prakt. Chem. 1995, 337, 209–215. [Google Scholar] [CrossRef]
  58. Gallos, J.K.; Sarli, V.C.; Varvogli, A.C; Papadoyanni, C.Z.; Papaspyrou, S.D.; Argyropoulos, N.G. The hetero-Diels–Alder addition of ethyl 2-nitrosoacrylate to electron-rich alkenes as a route to unnatural α-amino acids. Tetrahedron Lett. 2003, 44, 3905–3909. [Google Scholar] [CrossRef]
  59. Wabnitz, T.C.; Saaby, S.; Jorgensen, K.A. The first catalytic inverse-electron demand hetero-Diels–Alder reaction of nitroso alkenes using pyrrolidine as an organocatalyst. Org. Biomol. Chem. 2004, 2, 828–834. [Google Scholar] [CrossRef] [PubMed]
  60. Gallos, J.K.; Sarli, V.C.; Massen, Z.S.; Varvogli, A.C.; Papadoyanni, C.Z; Papaspyrou, S.D.; Argyropoulos, N.G. A new strategy for the stereoselective synthesis of unnatural a-amino acids. Tetrahedron 2005, 61, 565–574. [Google Scholar] [CrossRef]
  61. Domingo, L.R.; Picher, M.T.; Arroyo, P. Towards an understanding of the polar Diels–Alder reactions of nitrosoalkenes with enamines: A theoretical study. Eur. J. Org. Chem. 2006, 2570–2580. [Google Scholar] [CrossRef]
  62. Rossi, E.; Abbiati, G.; Attanasi, O.A.; Rizzato, S.; Santeusanio, S. Divergent and solvent dependent reactions of 4-ethoxycarbonyl-3-methyl-1-tert-butoxycarbonyl-1,2-diaza-1,3-diene with enamines. Tetrahedron 2007, 63, 11055–11065. [Google Scholar] [CrossRef]
  63. Attanasi, O.A.; Favi, G.; Filippone, P.; Forzato, C.; Giorgi, G.; Morganti, S.; Nitti, P.; Pitacco, G.; Rizzato, E.; Spinelli, S.; Valentin, E. On the reactivity of some 2-methyleneindolines with β-nitroenamines, α-nitroalkenes, and 1,2-diaza-1,3-butadienes. Tetrahedron 2006, 62, 6420–6434. [Google Scholar] [CrossRef]
  64. South, M.S. Novel cyclization reactions of mono- and dichloroazodienes. A new synthesis of substituted pyridazines. J. Heterocycl. Chem. 1999, 36, 301–319. [Google Scholar] [CrossRef]
  65. South, M.S.; Jakuboski, T.L.; Westmeyer, M.D.; Dukesherer, D.R. Synthesis and reactions of haloazodienes. A new and general synthesis of substituted pyridazines. J. Org. Chem. 1996, 61, 8921–8934. [Google Scholar] [CrossRef] [PubMed]
  66. South, M.S.; Jakuboski, T.L.; Westmeyer, M.D.; Dukesherer, D.R. Novel cyclization reactions of dichloroazodienes. Tetrahedron Lett. 1996, 37, 1351–1354. [Google Scholar] [CrossRef]
  67. South, M.S.; Jakuboski, T.L. Synthesis and reactions of chloroazodienes. A new and general synthesis of pyridazines. Tetrahedron Lett. 1995, 36, 5703–5706. [Google Scholar]
  68. Clarke, S.J.; Davies, D.E.; Gilchrist, T.L. Competing [4+2] and [3+2] cycloaddition in the reactions of nucleophilic olefins with ethyl 3-(toluene-p-sulphonylazo)but-2-enoate. J. Chem. Soc. Perkin Trans. 1 1983, 1803–1807. [Google Scholar] [CrossRef]
  69. Sommer, S. [3+2]-cycloadditionen von azoalkenen an enamine—“criss-cross”—cycloadditionen an azoalkene. Angew.Chem. 1979, 91, 756–757. [Google Scholar] [CrossRef]
  70. De los Santos, J.M.; Ignacio, R.; Aparicio, D.; Palacios, F.; Ezpeleta, J.M. Reactions of conjugate phosphinyl- and phosphonyl-nitroso alkenes with enamines. Preparation of N-hydroxypyrrole derivatives. J. Org. Chem. 2009, 74, 3444–3448. [Google Scholar] [CrossRef] [PubMed]
  71. Palacios, F.; Aparicio, D.; López, Y.; De los Santos, J.M.; Alonso, C. Cycloaddition reactions of phosphorylated 1,2-diaza-1,3-butadienes with olefins: Regioselective synthesis of pyridazine derivatives. Eur. J. Org. Chem. 2005, 1142–1147. [Google Scholar] [CrossRef]
  72. Gilchrist, T.L.; Lemos, A. Reaction of pyrroles with ethyl 2-nitroso- and 2-azo-propenoates, and with ethyl cyanoformate N-oxide: A comparison of the reaction pathways. J. Chem. Soc. Perkin Trans. 1 1993, 1391–1395. [Google Scholar] [CrossRef]
  73. Clarke, S.J.; Gilchrist, T.L.; Lemos, A.; Roberts, T.G. Reactions of azoalkenes derived from hydrazones of ethyl bromopyruvate with electron rich alkenes and heterocycles. Tetrahedron 1991, 47, 5615–5624. [Google Scholar] [CrossRef]
  74. Lemos, A.; Lopes, M. Reactions of nitrosovinylphosphonates with electron-rich alkenes and heterocycles. Phosphor. Sulfur Silicon 2007, 182, 2149–2155. [Google Scholar]
  75. Hu, E.X. Nucleophilic ring opening of aziridines. Tetrahedron 2004, 60, 2701–2743. [Google Scholar] [CrossRef]
  76. Rogers, R.S.; Stern, M.K. An improved synthesis of the phosphonic acid analog of tryptophan. Synlett 1992, 708. [Google Scholar] [CrossRef]
  77. Burk, M.J.; Martinez, J.P.; Feaster, J.E.; Cosford, N. Catalytic asymmetric reductive amination of ketones via highly enantioselective hydrogenation of the C=N double bond. Tetrahedron 1994, 50, 4399–4428. [Google Scholar] [CrossRef]
  78. Demir, A.S.; Tanyeli, C.; Sesenoglu, O.; Demiç, S. A simple synthesis of 1-aminophosphonic acids from 1-hydroxyiminophosphonates with NaBH4 in the presence of transition metal compounds. Tetrahedron Lett. 1996, 37, 407–410. [Google Scholar] [CrossRef]
Sample Availability: Not applicable.
Figure 1. Name, structure and numbering of compounds.
Figure 1. Name, structure and numbering of compounds.
Molecules 14 04098 g001
Scheme 1. Synthesis of 2H-azirines by Swern oxidation.
Scheme 1. Synthesis of 2H-azirines by Swern oxidation.
Molecules 14 04098 sch001
Scheme 2. Thermolysis of vinyl azide 6.
Scheme 2. Thermolysis of vinyl azide 6.
Molecules 14 04098 sch002
Scheme 3. 2H-Azirines by Neber reactions of β-keto tosyloximes.
Scheme 3. 2H-Azirines by Neber reactions of β-keto tosyloximes.
Molecules 14 04098 sch003
Scheme 4. Synthesis of 3-vinyl- and 3-dihydroisoxazolyl-2H-azirines.
Scheme 4. Synthesis of 3-vinyl- and 3-dihydroisoxazolyl-2H-azirines.
Molecules 14 04098 sch004
Scheme 5. Hydride addition to 2H-azirines 9 and 10.
Scheme 5. Hydride addition to 2H-azirines 9 and 10.
Molecules 14 04098 sch005
Scheme 6. β-Amino-phosphine oxides and –phosphonates 22 from tosyloximes 19.
Scheme 6. β-Amino-phosphine oxides and –phosphonates 22 from tosyloximes 19.
Molecules 14 04098 sch006
Scheme 7. Addition of nitrogen heterocycles to 2H-azirines 23.
Scheme 7. Addition of nitrogen heterocycles to 2H-azirines 23.
Molecules 14 04098 sch007
Scheme 8. Imidazole mediated generation of 2H-azirine 20a and nucleophilic addition.
Scheme 8. Imidazole mediated generation of 2H-azirine 20a and nucleophilic addition.
Molecules 14 04098 sch008
Scheme 9. Methanol addition to 2H-azirines 20.
Scheme 9. Methanol addition to 2H-azirines 20.
Molecules 14 04098 sch009
Scheme 10. Benzenethiol addition to 2H-azirines 23 and 29.
Scheme 10. Benzenethiol addition to 2H-azirines 23 and 29.
Molecules 14 04098 sch010
Scheme 11. Addition of Grignard reagents to 2H-azirine-2-carboxylates.
Scheme 11. Addition of Grignard reagents to 2H-azirine-2-carboxylates.
Molecules 14 04098 sch011
Scheme 12. Generation of 2H-azirines 42 by Grignard reagents and nucleophilic addition.
Scheme 12. Generation of 2H-azirines 42 by Grignard reagents and nucleophilic addition.
Molecules 14 04098 sch012
Scheme 13. Ketamides 48 from 2H-azirines 23 and 29.
Scheme 13. Ketamides 48 from 2H-azirines 23 and 29.
Molecules 14 04098 sch013
Scheme 14. Reactions of acyl chlorides with phosphinyl- and phosphonyl-2H-azirines.
Scheme 14. Reactions of acyl chlorides with phosphinyl- and phosphonyl-2H-azirines.
Molecules 14 04098 sch014
Scheme 15. Cycloaddition reactions of 2H-azirine 51.
Scheme 15. Cycloaddition reactions of 2H-azirine 51.
Molecules 14 04098 sch015
Scheme 16. Cycloaddition reactions of 2H-azirine 55 with nucleophilic dienes.
Scheme 16. Cycloaddition reactions of 2H-azirine 55 with nucleophilic dienes.
Molecules 14 04098 sch016
Scheme 17. General method for the obtention of nitroso- and azo-alkenes.
Scheme 17. General method for the obtention of nitroso- and azo-alkenes.
Molecules 14 04098 sch017
Scheme 18. Nucleophilic addition of ammonia, aminoesters and aminoalcohols to azoalkene 64.
Scheme 18. Nucleophilic addition of ammonia, aminoesters and aminoalcohols to azoalkene 64.
Molecules 14 04098 sch018
Scheme 19. Addition of amino esters to optically active azoalkene 68.
Scheme 19. Addition of amino esters to optically active azoalkene 68.
Molecules 14 04098 sch019
Scheme 20. 1,4-Addtion of 1,2-aminothiols 72 to nitrosoalkenes 71.
Scheme 20. 1,4-Addtion of 1,2-aminothiols 72 to nitrosoalkenes 71.
Molecules 14 04098 sch020
Scheme 21. Addition of thiourea and methanol to 1,2-diaza-1,3-butadienes 75.
Scheme 21. Addition of thiourea and methanol to 1,2-diaza-1,3-butadienes 75.
Molecules 14 04098 sch021
Scheme 22. Addition of 3-mercapto-2-ketones 80 to azoalkenes 79.
Scheme 22. Addition of 3-mercapto-2-ketones 80 to azoalkenes 79.
Molecules 14 04098 sch022
Scheme 23. Addition of 1,2-diamines to phosphinyl- and phosphonyl-azoalkenes 83.
Scheme 23. Addition of 1,2-diamines to phosphinyl- and phosphonyl-azoalkenes 83.
Molecules 14 04098 sch023
Scheme 24. N-methylimidazole generation of azoalkene 89 and nucleophilic addition.
Scheme 24. N-methylimidazole generation of azoalkene 89 and nucleophilic addition.
Molecules 14 04098 sch024
Scheme 25. Nitrosoalkenes 92 as Michael acceptors.
Scheme 25. Nitrosoalkenes 92 as Michael acceptors.
Molecules 14 04098 sch025
Scheme 26. Reaction of nitrosoalkenes 92 with enamines 99.
Scheme 26. Reaction of nitrosoalkenes 92 with enamines 99.
Molecules 14 04098 sch026
Scheme 27. Cycloaddition reactions of azoalkenes 64.
Scheme 27. Cycloaddition reactions of azoalkenes 64.
Molecules 14 04098 sch027
Scheme 28. Cycloaddition reactions of azoalkenes 89 and 115 with electron rich olefins.
Scheme 28. Cycloaddition reactions of azoalkenes 89 and 115 with electron rich olefins.
Molecules 14 04098 sch028
Scheme 29. Cycloaddition reactions of nitrosovinylphosphonates 128.
Scheme 29. Cycloaddition reactions of nitrosovinylphosphonates 128.
Molecules 14 04098 sch029
Table 1. Addition of Grignard reagents.
Table 1. Addition of Grignard reagents.
Molecules 14 04098 i001
Table 2. Reactions of azoalkene 89 with electron-rich heterocycles.
Table 2. Reactions of azoalkene 89 with electron-rich heterocycles.
Molecules 14 04098 i002

Share and Cite

MDPI and ACS Style

Lemos, A. Addition and Cycloaddition Reactions of Phosphinyl- and Phosphonyl-2H-Azirines, Nitrosoalkenes and Azoalkenes. Molecules 2009, 14, 4098-4119. https://doi.org/10.3390/molecules14104098

AMA Style

Lemos A. Addition and Cycloaddition Reactions of Phosphinyl- and Phosphonyl-2H-Azirines, Nitrosoalkenes and Azoalkenes. Molecules. 2009; 14(10):4098-4119. https://doi.org/10.3390/molecules14104098

Chicago/Turabian Style

Lemos, Américo. 2009. "Addition and Cycloaddition Reactions of Phosphinyl- and Phosphonyl-2H-Azirines, Nitrosoalkenes and Azoalkenes" Molecules 14, no. 10: 4098-4119. https://doi.org/10.3390/molecules14104098

Article Metrics

Back to TopTop