Stable phosphorus radicals

From Wikipedia, the free encyclopedia
Spin Density map on phosphinyl radical found by NBO analysis.

Stable and persistent phosphorus radicals are phosphorus-centred radicals that are isolable and can exist for at least short periods of time.[1] Radicals consisting of main group elements are often very reactive and undergo uncontrollable reactions, notably dimerization and polymerization.[2] The common strategies for stabilising these phosphorus radicals usually include the delocalisation of the unpaired electron over a pi system or nearby electronegative atoms, and kinetic stabilisation with bulky ligands. Stable and persistent phosphorus radicals can be classified into three categories: neutral, cationic, and anionic radicals. Each of these classes involve various sub-classes, with neutral phosphorus radicals being the most extensively studied. Phosphorus exists as one isotope 31P (I = 1/2) with large hyperfine couplings relative to other spin active nuclei, making phosphorus radicals particularly attractive for spin-labelling experiments.[1]

Neutral phosphorus radicals[edit]

Neutral phosphorus radicals include a large range of conformations with varying spin densities at the phosphorus. Generally, they can categorised as mono- and bi/di-radicals (also referred to as bisradicals and biradicaloids) for species containing one or two radical phosphorus centres respectively.[2]

Monoradicals[edit]

In 1966, Muller et. al published the first electron paramagnetic resonance (EPR/ESR) spectra displaying evidence for the existence of phosphorus-containing radicals.[3] Since then a variety of phosphorus monoradicals have been synthesised and isolated. Common ones include phosphinyl (R2P), phosphonyl (R2PO), and phosphoranyl (R4P) radicals.[1]

Synthesis[edit]

Synthetic methods for obtaining neutral phosphorus mondoradicals include photolytic reduction of trivalent phosphorus chlorides, P-P homolytic cleavage, single electron oxidation of phosphines, and cleavage of P-S or P-Se bonds.

Photolysis of three-coordinate phosphorus chloride for the synthesis of [(Me3Si)2N]2P by Lappert and co-workers.[4]

The first persistent two-coordinate phosphorus-centred radicals [(Me3Si)2N]2P and [(Me3Si)2CH]2P were reported in 1976 by Lappert and co-workers. They are prepared by photolysis of the corresponding three-coordinate phosphorus chlorides in toluene in the presence of an electron-rich olifin. [4] In 2000, the Power group found that this species can be synthesised from the dissolution, melting or evaporation of the dimer.[5]

Synthesis of the first stable diphosphanyl radical [Mes*MeP-PMes*] by Grützmacher and co-workers via reduction of phosphonium salt.[6]

In 2001, Grützmacher et al. reported the first stable diphosphanyl radical [Mes*MeP-PMes*] (Mes = 1,3,5-trimethylbenzene) from the reduction of the phosphonium salt [Mes*MeP-PMes*]+(O3SCF3) in an acetonitrile solution containing tetrakis(dimethylamino)ethylene (TDE) at room temperature, yielding yellow crystals. [6] The monomer is stable below -30 ºC in the solid state for a few days. At room temperature the species decomposes in solution and in the solid state with a half life of 30 minutes at 3 x 10−2 M.

Synthesis of [Me3SiNP(µ3-NtBu)33-Li(thf)}3X] (X = Br, I) by Armstrong and co-workers via oxidation.[7]

The first structurally characterised phosphorus radical [Me3SiNP(µ3-NtBu)33-Li(thf)}3X] (X = Br, I) was synthesised by Armstrong et al. in 2004 by the oxidation of the starting material with halogens bromide or iodine in a mixture of toluene and THF at 297 K. This produces blue crystals that can be characterised by X-ray crystallography.[7] The steric bulk of the alkyl-imido groups was identified as playing a major role in the stabilising of these radicals.

Synthesis of air tolerant and air stable 1,3-diphosphayclobutenyl radical by Ito and co-workers via reduction.[8]

In 2006, Ito et al. prepared an air tolerant and thermally stable 1,3-diphosphayclobutenyl radical.[8] Sterically bulky phospholkyne (Mes*C≡P) is treated with 0.5 equiv of t-BuLi in THF to form a 1,3 diphosphaalkyl anion. This is reduced with iodine solution to form a red product. The species is a planar four-membered diphosphacyclobutane (C2P2) ring with the Mes* having torsional angles with the C2P2 plane.[8]

Metal stabilised radicals[edit]

In 2007, Cummins et al. synthsised a phosphorus radical using nitridovanadium trisanilide metallo-ligands with similar form to Lappert, Power and co-workers' "jack-in-the-box" diphosphines.[9] This is made by the synthesis of the radical precursor ClP[NV{N(Np)Ar}]3]2 followed by its one electron reduction with Ti[N(tBu)Ar]3 or potassium graphite to yield dark brown crystals in 77% yield.[10] EPR data showed delocalisation of electron spin across the two 51V and one 31P nuclei. This was consistent with computation, supporting the reported resonance structures. This delocalisation across the vanadium atoms was identified as the source of stabilisation for this species due to the ease for transition metals to undergo one-electron chemistry. Cummins and co-workers postulated that the p-character of the system could be tuned by changing the metal centres.

Resonance structures of [P{NV[N(Np)Ar]3}2] showing delocalisation of radical across vanadium and phosphorus nuclei.[10]

Other metals stabilised radicals have been reported by Scheer et al, and Schneider et al using ligand containing tungsten and osmium respectively.[11][12]

Structure and properties[edit]

Schematic of DFT calculation results for diphosphine radical 1 in the solid state, the syn,anti-PR2 radical (1A2 and 1A2 ), the H optimised radical (1B1 and 1B2), the syn,anti-PR2 radical fully optimised (1C), and syn,syn-PR2 radical in optimised geometry 2. Energies are in kJ mol−1.[13] Illustrating the "Jack-in-the-box" model.

As previously mentioned, kinetic stabilisation through bulky ligands has been an effective strategy for producing persisting phosphorus radicals. Delocalisation of the electron has also shown a stabilising effect on phosphorus radical species. This conversely results in more delocalised spin densities, and lower coupling constants relative to 31P localised electron spin. For this reason the spin localisation on the phosphorus atom varies widely for different phosphorus radical species.[2]

Cyclic radicals like that by Ito at al have delocalisation across the rings. In this case X-ray, EPR spectroscopy, and ab initio calculations found that 80-90% of the spin was delocalised on the carbons in the C2P2 ring and the rest on the phosphorus atoms. Despite this, the aP2 constant shows similar spectroscopic property to organic radicals that contain conjugated P=C doubles bond, justifying the resonance structure used for this species.[8]

The phosphinyl radicals synthesised by Lappert and co-workers were found to be stable at room temperature for periods of over 15 days with no effect from short-term heating at 360 K.[4] This stability was assigned to the steric bulk of the substituents and the absence of beta-hydrogen atoms. A structural study of this species conducted using X-ray crystallography, gas-phase electron diffraction, and ab initio molecular orbital calculations found that the source of this stability was not the bulkiness of the CH(SiMe3)2 ligands but the release of strain energy during homolytic cleavage at the P-P bond of the dimer that favoured the existence of the radical.[13] The dimer shows a syn,anti conformation, which allows for better packing but has excessive crowding at the trimethylsilyl groups, while the radical monomer displays syn,syn conformation. Theoretical calculations showed that the process of cleaving the P-P bond (endothermic), relaxation to release steric strain, and rotation about the P-C bond to yield syn,syn conformation on the monomer radical (exothermic by 67.5 kJ for each unit) is an overall exothermic process.[13] The stability of this species can therefore be attributed to the energy release of strain energy by the reorganisation of the ligands as the dimer converts to the radical monomer. This effect have been observed in other systems containing the CH(SiMe3)2 ligand and was dubbed the "Jack-in-the-box" model.[14][15][16] Other ligand with similar flexibility, and ability to undergo conformational changes were identified as PnR2 (Pn - P, As, Sb) and ERR'2 (E = Si, Ge, Sn; R' = bulky ligand).[13]

In 2022, Streubel and co-workers investigated the electron density distribution across centres in metal-coordinated phosphanoxyl complexes.[17] This study showed that tungsten-containing radical complexes have small amounts of spin density on the metal nuclei while in the case of manganese and iron, the spins are purely metal-centred.[18]

Biradicals[edit]

Biradicals are molecules bearing two unpaired electrons. These radicals can interact ferromagnetically (triplet), antiferromagnetically (open-shell singlet) or not interact at all (two-doublet). [2] Biradicaloids/diradicaloids are a class of biradicals with significant radical centre interaction.[2]

Synthesis[edit]

The first phosphorus biradical was reported in 2011 by T. Breweies and co-workers. The biradicaloid [P(µ-NR)]2 (R=Hyp, Ter) was synthesised by the reduction of cyclo-1,3-diphospha (III)-2,4-diazanes using [(Cp2TiCl}2] as the reducing agent. [19] The bulky Ter and Hyp substituents provide a large stabilising effect. This effect is more pronounced with Ter where the biradical is stable in inert atmospheres in the solid state for long periods of time at temperatures up to 224 C. Computational studies determined that the [P(µ-NTer)]2 radical shows an openshell singlet ground state biradical character.[19]

Villinger et al later synthesised a stable cyclopentane-1,3-diyl biradical by the insertion of CO into a P–N bond of diphosphadiazanediyl.[20]

Synthesis of [P(µ-NR)]2 (R=Hyp, Ter) via reduction of cyclo-1,3-diphospha(III)-2,4-diazanes and subsequent CO insertion by Villinger and co-workers.[19][20]
Synthesis of (iPr)CP]2 radical via reduction by Rottschafer and co-workers with resonance structures.[21]

In 2017 D. Rottschäfer et al reported a N-heterocyclic vinylindene-stabilised singlet biradicaloid phosphorus compound (iPr)CP]2 (iPr = 1,3-bis(2,6-diisopropylphenyl)imidazol-2-ylidene). Significant π-e density is transferred to C2P2 ring.[21] The species was found to be diamagnetic with temperature-independent NMR resonances, so can be considered a non-Kekulé molecule.[21]

Structure and properties[edit]

The species by Villinger can undergo reaction with phosphaalkyne forming a five-membered P2N2C heterocycle with a P-C bridge. It can also undergo halogenation and reaction with elemental sulfur.[20]

Reactivity of [P(µ-NR)]2 (R=Hyp, Ter) radical.[20]

Characterisation[edit]

Solvation of lithium ions in [Me3SiNP(µ3-NtBu)33-Li(thf)}3I] in very dilute THF solutions.[7]

Phosphorus radicals are commonly characterized by EPR/ESR to elucidate the spin localisation of the radical across the radical species. Higher coupling constants are indicative of higher localisation on phosphorus nuclei. Quantum chemical calculations on these systems are also used to support this experimental data.[1]

Before the characterization by X-ray crystallography by Armstrong et al, the structure of the phosphorus centred radical [(Me3Si)2CH]2P had been determined by electron diffraction.[4] The diphosphanyl radical [Mes*MeP-PMes*] had been stabilised through doping into crystals of Mes*MePPMeMes*.[6] The radical synthesised by Armstrong et al was found to exist as a distorted PN3Li3X cube in the solid state. They found that upon dissolution in THF, this cubic structure is disrupted, leaving the species to form a solvent-separated ion pair.[7]

Phosphorus radical cations[edit]

Synthesis[edit]

Phosphorus radical cations are often obtained from the one-electron oxidation of diphosphinidenes and phosphalkenes.

Synthesis of [(cAAC)2P2]•+ and [(NHC)2P2]•+ via oxidation with Ph3C+B(C6F5)4 by Bertrand and co-workers.[22]
Synthesis of [(TMP)P(cAAC)]•+ via oxidation with Ph3C+ (C6F5)4B by Bertrand and co-workers.[23]
Synthesis of [bis(carbene)-PN]•+ visa oxidation with h3C+ (C6F5)4B by Bertrand and co-workers.[24]

In 2010, the Bertrand group found that carbene-stabilised diphosphinidenes can undergo one-electron oxidation in toluene with Ph3C+B(C6F5)4 at room temperature in inert atmosphere to produce radical cations (Dipp=2,6-Diisopropylphenyl)[22].  The Bertrand group reported the synthesis of [(cAAC)P2]•+ , [(NHC)P2]•+ and [(NHC)P2]++ . The EPR signal for [(cAAC)P2]•+ is a triplet of quintents, resulting form coupling to with 2 P nuclei and a small coupling with 2 N nuclei. NBO analysis showed spin delocalisation across two phosphorus atoms (0.27e each) and nitrogen atoms(0.14e each). Contrastingly, the [(NHC)P2]•+complex showed delocalisation mostly on phosphorus (0.33e and 0.44e) with little contribution of other elements.[22] Other diradicals synthesised by the Bertrand group involved species single phosphorus atoms. These included [(TMP)P(cAAC)]•+ where spin is localised on phosphorus (67%)[23] and [bis(carbene)-PN]•+ with spin density distributed over phosphorus (0.40e), central nitrogen atom (0.18e), and N atom of cAAC (0.19e). Treatment with this later cation with KC8 returns it to its neutral analogue.[24]

Synthesis of Mes*P-(C(NMe2)2)+ via a one electron oxidation of a phosphaalkenes with [Cp2Fe]PF6 by Geoffroy and co-workers.[25]

In 2003, Geoffroy et al. synthesised Mes*P-(C(NMe2)2)+ through a one electron oxidation of a phosphaalkenes with [Cp2Fe]PF6.[25] A solution of Mes*P-(C(NMe2)2)+ is stable in inert atmosphere in the solid state for a few weeks and a few days in solution. Hyperfine couplings on EPR show strong localisation of the spin to the phosphorus nuclei (0.75e in p orbital). In 2015, the Wang group was able to isolate the crystal structure of this species with use of the oxidant of a weakly coordinating anion Ag[Al(ORF)4].[26] The electron spin density, found by EPR, resides principally on phosphorus 3p and 3s orbitals (68.2% and 2.46% respectively). This was supported by DFT calculations where 80.9% of spin density was found to be localised on phosphorus atom.[26]

General scheme for preparation of cyclic radical cations via oxidation.[27]

Weakly coordinating anions were also used to stabilise cyclic biradical cations synthesised by Schulz and colleagues where the spin density was found to reside exclusively on the phosphorus atoms (0.46e each) in the case of [P(μ-NTer)2P]•+.[28] In the case of [P(μ-NTer)2As]•+ the spin was found to mostly reside on the As nuclei (70.6% on As compared to 29.4% on P atom). Many other cyclic radical cations have been reported.[29]

Synthesis of divinyldiphosphene radical cations via oxidation with GaCl3 by Ghadwal and co-workers.[30]

It is difficult to form radical cations with diphosphenes due to low lying HOMO at the phosphorus centre. Ghadwal and co-workers were able to synthesise a diphosphene radical cation [{(NHC)C(Ph)}P]2•+ using an NHC-derived divinyldiphosphene with a high lying HOMO and an small HOMO-LUMO gap. The stability of the species was identified as the delocalisation of the spin density across the CP2C-unit.[30] The spin density was found to be 11-14% on each P nuclei and 17-21% on each C nuclei.[30]

Structure and properties[edit]

A unique source of stability for phosphorus radical cations is the electrostatic repulsion between radical cations that prevents dimerisation.[31]

Weakly coordinating anions have been used to stabilise biradical cations.[2]

Phosphorus radical anions[edit]

Synthesis[edit]

The most common method for accessing radical anions is through the use of reducing agents.

Synthesis of phosphorus-centred radical anion via reduction usgin K or Li by Wang and co-workers.
Synthesis of diphosphorus-centred radical anion and the di-radical di-anion via reduction with KC8 by Wang and co-workers.[32]

In 2014 the Wang group reported the synthesis of a phosphorus-centred radical anion through the reduction of a phosphaalkene using either Li in DME or K in THF yielding purple crystals.[32] EPR data showed localisation of the spin on 3p (51.09%) and 3s (1.62%) orbitals of phosphorus. They later synthesised a diphosphorus-centred radial anion and the first di-radical di-anion from the reduction of the diphosphaalkene with KC8 in THF in the presence of 18-crown-6.[33] In both cases the spin density resides principally on the phosphorus nuclei.

Synthesis of phosphorus radical anion coordinated with CoII and FeII complexes by Tan and co-workers.[34]

Tan and co-workers used a charge transfer approach to synthesis the phosphorus radical anion coordinated CoII and FeII complexes. Here diazafluorenylidene-substituted phosphaalkene is reacted with low valent transition metal complexes to form phosphorus radical anions coordinated with metal complexes.[34] This species displays a quartet ground state showing weak antiferromagnetic interaction of the phosphorus radical with the high-spim TMII ion. The spin density is mostly localised on TM and phosphorus nuclei.[34] The group further synthesised radical anion lanthanide complexes which also showed antiferromagnetic interaction.[35]

Synthesis of phosphorus radical anion with boryl substituents by Yamashita and co-workers.[36]

The π-acid properties of boryl substituents were employed by Yamashita and co-workers to stabilise phosphorus radical anions.[36] Here the diazafluorenylidene-substituted phosphaalkene is reacted with [Cp*2Ln][BPh4] (Ln = Dy, Tb, and Gd) followed by reduction with KC8 in the absence or presence of 2,2,2-cryptand yielding complexes with radical anion phosphaalkene fragments. EPR and DFT calculations indicate spin density mostly localised on the P nuclei (67.4%).

Further reading[edit]

Reviews[edit]

  • Marque, Sylvain; Tordo, Paul (2005). "Reactivity of Phosphorus Centered Radicals". New Aspects in Phosphorus Chemistry V. Topics in Current Chemistry. Vol. 250. pp. 43–76. doi:10.1007/b100981. ISBN 978-3-540-22498-3.
  • Armstrong, A.; Chivers, T.; Boeré, R. T. (2005). "The Diversity of Stable and Persistent Phosphorus-Containing Radicals". Modern Aspects of Main Group Chemistry. ACS Symposium Series. Vol. 917. pp. 66–80. doi:10.1021/bk-2005-0917.ch005. ISBN 9780841239265.
  • Das, Bindusagar; Makol, Abhishek; Kundu, Subrata (2022). "Phosphorus radicals and radical ions". Dalton Transactions. 51 (33): 12404–12426. doi:10.1039/D2DT01499H. PMID 35920252. S2CID 250659955.

Reactivity[edit]

Potential applications[edit]

References[edit]

  1. ^ a b c d Armstrong, A.; Chivers, T.; Boere, R. T. (2006-10-03). "The Diversity of Stable and Persistent Phosphorus-Containing Radicals". ChemInform. 37 (40). doi:10.1002/chin.200640250. ISSN 0931-7597.
  2. ^ a b c d e f Das, Bindusagar; Makol, Abhishek; Kundu, Subrata (2022). "Phosphorus radicals and radical ions". Dalton Transactions. 51 (33): 12404–12426. doi:10.1039/D2DT01499H. ISSN 1477-9226. PMID 35920252. S2CID 250659955.
  3. ^ Schmidt, Ulrich; Kabitzke, Karlheinz; Markau, Klaus; Müller, Asmus (1966). "Zur Kenntnis zweibindiger Phosphor‐Radikale". Chemische Berichte. 99 (5): 1497–1501. doi:10.1002/cber.19660990512. ISSN 0009-2940.
  4. ^ a b c d Gynane, Michael J. S.; Hudson, Andrew; Lappert, Michael F.; Power, Philip P. (1976). "Synthesis and electron spin resonance study of stable dialkyls and diamides of phosphorus and arsenic, R 1 2 M· and (R 2 2 N) 2 M·". J. Chem. Soc., Chem. Commun. (16): 623–624. doi:10.1039/C39760000623. ISSN 0022-4936.
  5. ^ Hinchley, Sarah L.; Morrison, Carole A.; Rankin, David W. H.; Macdonald, Charles L. B.; Wiacek, Robert J.; Cowley, Alan H.; Lappert, Michael F.; Gundersen, Grete; Clyburne, Jason A. C.; Power, Philip P. (2000-01-01). "Persistent phosphinyl radicals from a bulky diphosphine: an example of a molecular jack-in-the-box". Chemical Communications (20): 2045–2046. doi:10.1039/B004889P. ISSN 1364-548X.
  6. ^ a b c Loss, Sandra; Magistrato, Alessandra; Cataldo, Laurent; Hoffmann, Stefan; Geoffroy, Michel; Röthlisberger, Ursula; Grützmacher, Hansjörg (2001-02-15). "Isolation of a Highly Persistent Diphosphanyl Radical: The Phosphorus Analogue of a Hydrazyl". Angewandte Chemie International Edition. 40 (4): 723–726. doi:10.1002/1521-3773(20010216)40:4<723::aid-anie7230>3.0.co;2-8. ISSN 1433-7851. PMID 11241603.
  7. ^ a b c d Armstrong, Andrea; Chivers, Tristram; Parvez, Masood; Boeré, Rene T. (2004-01-16). "Stable Cubic Phosphorus-Containing Radicals". Angewandte Chemie International Edition. 43 (4): 502–505. doi:10.1002/anie.200353108. ISSN 1433-7851. PMID 14735546.
  8. ^ a b c d Ito, Shigekazu; Kikuchi, Manabu; Yoshifuji, Masaaki; Arduengo, Anthony J.; Konovalova, Tatyana A.; Kispert, Lowell D. (2006-06-26). "Preparation and Characterization of an Air-Tolerant 1,3-Diphosphacyclobuten-4-yl Radical". Angewandte Chemie International Edition. 45 (26): 4341–4345. doi:10.1002/anie.200600950. ISSN 1433-7851. PMID 16739144.
  9. ^ Hinchley, Sarah L.; Morrison, Carole A.; Rankin, David W. H.; Macdonald, Charles L. B.; Wiacek, Robert J.; Cowley, Alan H.; Lappert, Michael F.; Gundersen, Grete; Clyburne, Jason A. C.; Power, Philip P. (2000-01-01). "Persistent phosphinyl radicals from a bulky diphosphine: an example of a molecular jack-in-the-box". Chemical Communications (20): 2045–2046. doi:10.1039/B004889P. ISSN 1364-548X.
  10. ^ a b Agarwal, Paresh; Piro, Nicholas A.; Meyer, Karsten; Müller, Peter; Cummins, Christopher C. (2007-04-20). "An Isolable and Monomeric Phosphorus Radical That Is Resonance-Stabilized by the Vanadium(IV/V) Redox Couple". Angewandte Chemie International Edition. 46 (17): 3111–3114. doi:10.1002/anie.200700059. PMID 17351998.
  11. ^ Scheer, Manfred; Kuntz, Christian; Stubenhofer, Markus; Linseis, Michael; Winter, Rainer F.; Sierka, Marek (2009-03-23). "The Complexed Triphosphaallyl Radical, Cation, and Anion Family". Angewandte Chemie International Edition. 48 (14): 2600–2604. doi:10.1002/anie.200805892. ISSN 1433-7851. PMID 19248064.
  12. ^ Abbenseth, Josh; Delony, Daniel; Neben, Marc C.; Würtele, Christian; de Bruin, Bas; Schneider, Sven (2019-03-06). "Interconversion of Phosphinyl Radical and Phosphinidene Complexes by Proton Coupled Electron Transfer". Angewandte Chemie. 131 (19): 6404–6407. Bibcode:2019AngCh.131.6404A. doi:10.1002/ange.201901470. hdl:11245.1/ffb134b2-06c7-403d-8cdd-321af678f759. ISSN 0044-8249.
  13. ^ a b c d Hinchley, Sarah L.; Morrison, Carole A.; Rankin, David W. H.; Macdonald, Charles L. B.; Wiacek, Robert J.; Voigt, Andreas; Cowley, Alan H.; Lappert, Michael F.; Gundersen, Grete; Clyburne, Jason A. C.; Power, Philip P. (2001-09-01). "Spontaneous Generation of Stable Pnictinyl Radicals from "Jack-in-the-Box" Dipnictines: A Solid-State, Gas-Phase, and Theoretical Investigation of the Origins of Steric Stabilization 1". Journal of the American Chemical Society. 123 (37): 9045–9053. doi:10.1021/ja010615b. ISSN 0002-7863. PMID 11552812.
  14. ^ Fjeldberg, Torgny; Haaland, Arne; Schilling, Birgitte E. R.; Lappert, Michael F.; Thorne, Andrew J. (1986-01-01). "Subvalent Group 4B metal alkyls and amides. Part 8. Germanium and tin carbene analogues MR2[M = Ge or Sn, R = CH(SiMe3)2]: syntheses and structures in the gas phase (electron diffraction); molecular-orbital calculations for MH2 and GeMe2". Journal of the Chemical Society, Dalton Transactions (8): 1551–1556. doi:10.1039/DT9860001551. ISSN 1364-5447.
  15. ^ Goldberg, David E.; Hitchcock, Peter B.; Lappert, Michael F.; Thomas, K. Mark; Thorne, Andrew J.; Fjeldberg, Torgny; Haaland, Arne; Schilling, Birgitte E. R. (1986-01-01). "Subvalent Group 4B metal alkyls and amides. Part 9. Germanium and tin alkene analogues, the dimetallenes M2R4[M = Ge or Sn, R = CH(SiMe3)2]: X-ray structures, molecular orbital calculations for M2H4, and trends in the series M2R′4[M = C, Si, Ge, or Sn; R′= R, Ph, C6H2Me3-2,4,6, or C6H3Et2-2,6]". Journal of the Chemical Society, Dalton Transactions (11): 2387–2394. doi:10.1039/DT9860002387. ISSN 1364-5447.
  16. ^ Fjeldberg, Torgny; Hope, Håkon; Lappert, Michael F.; Power, Philip P.; Thorne, Andrew J. (1983-01-01). "Molecular structures of the main group 4 metal(II) bis(trimethylsilyl)-amides M[N(SiMe3)2]2 in the crystal (X-ray) and vapour (gas-phase electron diffraction)". Journal of the Chemical Society, Chemical Communications (11): 639–641. doi:10.1039/C39830000639. ISSN 0022-4936.
  17. ^ Brehm, Philipp C.; Frontera, Antonio; Streubel, Rainer (2022). "On metal coordination of neutral open-shell P-ligands focusing on phosphanoxyls, their electron residence and reactivity". Chemical Communications. 58 (43): 6270–6279. doi:10.1039/d2cc01302a. ISSN 1359-7345. PMID 35579028. S2CID 248598597.
  18. ^ Brehm, Philipp C.; Frontera, Antonio; Streubel, Rainer (2022). "On metal coordination of neutral open-shell P-ligands focusing on phosphanoxyls, their electron residence and reactivity". Chemical Communications. 58 (43): 6270–6279. doi:10.1039/D2CC01302A. ISSN 1359-7345. PMID 35579028. S2CID 248598597.
  19. ^ a b c Beweries, Torsten; Kuzora, Rene; Rosenthal, Uwe; Schulz, Axel; Villinger, Alexander (2011-09-12). "[P(μ-NTer)]2: A Biradicaloid That Is Stable at High Temperature". Angewandte Chemie International Edition. 50 (38): 8974–8978. doi:10.1002/anie.201103742. PMID 21858902.
  20. ^ a b c d Hinz, Alexander; Schulz, Axel; Villinger, Alexander (2015-02-23). "Stable Heterocyclopentane-1,3-diyls". Angewandte Chemie International Edition. 54 (9): 2776–2779. doi:10.1002/anie.201410276. PMID 25604347.
  21. ^ a b c Rottschäfer, Dennis; Neumann, Beate; Stammler, Hans-Georg; Ghadwal, Rajendra S. (2017-07-06). "N -Heterocyclic Vinylidene-Stabilized Phosphorus Biradicaloid". Chemistry - A European Journal. 23 (38): 9044–9047. doi:10.1002/chem.201702433. PMID 28556982.
  22. ^ a b c Back, Olivier; Donnadieu, Bruno; Parameswaran, Pattiyil; Frenking, Gernot; Bertrand, Guy (2010). "Isolation of crystalline carbene-stabilized P2-radical cations and P2-dications". Nature Chemistry. 2 (5): 369–373. Bibcode:2010NatCh...2..369B. doi:10.1038/nchem.617. ISSN 1755-4330. PMID 20414236.
  23. ^ a b Back, Olivier; Celik, Mehmet Ali; Frenking, Gernot; Melaimi, Mohand; Donnadieu, Bruno; Bertrand, Guy (2010-08-04). "A Crystalline Phosphinyl Radical Cation". Journal of the American Chemical Society. 132 (30): 10262–10263. doi:10.1021/ja1046846. ISSN 0002-7863. PMID 20662507.
  24. ^ a b Kinjo, Rei; Donnadieu, Bruno; Bertrand, Guy (2010-08-09). "Isolation of a Carbene-Stabilized Phosphorus Mononitride and Its Radical Cation (PN +. )". Angewandte Chemie International Edition. 49 (34): 5930–5933. doi:10.1002/anie.201002889. PMID 20632430.
  25. ^ a b Rosa, Patrick; Gouverd, Cyril; Bernardinelli, Gérald; Berclaz, Théo; Geoffroy, Michel (2003-05-23). "Phosphaalkenes with Inverse Electron Density: Electrochemistry, Electron Paramagnetic Resonance Spectra, and Density Functional Theory Calculations of Aminophosphaalkene Derivatives". The Journal of Physical Chemistry A. 107 (24): 4883–4892. Bibcode:2003JPCA..107.4883R. doi:10.1021/jp030023a. ISSN 1089-5639.
  26. ^ a b Pan, Xiaobo; Wang, Xingyong; Zhang, Zaichao; Wang, Xinping (2015-08-18). "Two phosphaalkene radical cations with inverse spin density distributions". Dalton Transactions. 44 (34): 15099–15102. doi:10.1039/C5DT00656B. ISSN 1477-9234. PMID 25828200.
  27. ^ Brückner, Angelika; Hinz, Alexander; Priebe, Jacqueline B.; Schulz, Axel; Villinger, Alexander (2015-05-08). "Cyclic Group 15 Radical Cations". Angewandte Chemie International Edition. 54 (25): 7426–7430. doi:10.1002/anie.201502054. ISSN 1433-7851. PMID 25960190.
  28. ^ Brückner, Angelika; Hinz, Alexander; Priebe, Jacqueline B.; Schulz, Axel; Villinger, Alexander (2015-05-08). "Cyclic Group 15 Radical Cations". Angewandte Chemie International Edition. 54 (25): 7426–7430. doi:10.1002/anie.201502054. ISSN 1433-7851. PMID 25960190.
  29. ^ Su, Yuanting; Zheng, Xin; Wang, Xingyong; Zhang, Xuan; Sui, Yunxia; Wang, Xinping (2014-04-30). "Two Stable Phosphorus-Containing Four-Membered Ring Radical Cations with Inverse Spin Density Distributions". Journal of the American Chemical Society. 136 (17): 6251–6254. doi:10.1021/ja502675d. ISSN 0002-7863. PMID 24731124.
  30. ^ a b c Sharma, Mahendra K.; Rottschäfer, Dennis; Blomeyer, Sebastian; Neumann, Beate; Stammler, Hans-Georg; van Gastel, Maurice; Hinz, Alexander; Ghadwal, Rajendra S. (2019). "Diphosphene radical cations and dications with a π-conjugated C2P2C2-framework". Chemical Communications. 55 (70): 10408–10411. doi:10.1039/c9cc04701h. ISSN 1359-7345. PMID 31403648. S2CID 199540250.
  31. ^ Back, Olivier; Celik, Mehmet Ali; Frenking, Gernot; Melaimi, Mohand; Donnadieu, Bruno; Bertrand, Guy (2010-08-04). "A Crystalline Phosphinyl Radical Cation". Journal of the American Chemical Society. 132 (30): 10262–10263. doi:10.1021/ja1046846. ISSN 0002-7863. PMID 20662507.
  32. ^ a b Pan, Xiaobo; Wang, Xingyong; Zhao, Yue; Sui, Yunxia; Wang, Xinping (2014-07-16). "A Crystalline Phosphaalkene Radical Anion". Journal of the American Chemical Society. 136 (28): 9834–9837. doi:10.1021/ja504001x. ISSN 0002-7863. PMID 24977300.
  33. ^ Tan, Gengwen; Li, Shuyu; Chen, Sheng; Sui, Yunxia; Zhao, Yue; Wang, Xinping (2016-06-01). "Isolable Diphosphorus-Centered Radical Anion and Diradical Dianion". Journal of the American Chemical Society. 138 (21): 6735–6738. doi:10.1021/jacs.6b04081. ISSN 0002-7863. PMID 27182899.
  34. ^ a b c Tan, Gengwen; Li, Jing; Zhang, Li; Chen, Chao; Zhao, Yue; Wang, Xinping; Song, You; Zhang, Yi‐Quan; Driess, Matthias (2017-10-02). "The Charge Transfer Approach to Heavier Main‐Group Element Radicals in Transition‐Metal Complexes". Angewandte Chemie International Edition. 56 (41): 12741–12745. doi:10.1002/anie.201707501. ISSN 1433-7851. PMID 28801962.
  35. ^ Chen, Chao; Hu, Zhaobo; Li, Jing; Ruan, Huapeng; Zhao, Yue; Tan, Gengwen; Song, You; Wang, Xinping (2020-02-17). "Isolable Lanthanide Metal Complexes of a Phosphorus-Centered Radical". Inorganic Chemistry. 59 (4): 2111–2115. doi:10.1021/acs.inorgchem.9b01950. ISSN 0020-1669. PMID 31397564. S2CID 199505801.
  36. ^ a b Chen, Chao; Hu, Zhaobo; Li, Jing; Ruan, Huapeng; Zhao, Yue; Tan, Gengwen; Song, You; Wang, Xinping (2020-02-17). "Isolable Lanthanide Metal Complexes of a Phosphorus-Centered Radical". Inorganic Chemistry. 59 (4): 2111–2115. doi:10.1021/acs.inorgchem.9b01950. ISSN 0020-1669. PMID 31397564. S2CID 199505801.