Stable phosphorus radicals

Last updated
Spin Density map on phosphinyl radical found by NBO analysis. Spinphosphinyl.png
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]

Contents

Neutral phosphorus radicals

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

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

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. Lappert 1976.png
Photolysis of three-coordinate phosphorus chloride for the synthesis of [(Me3Si)2N]2P by Lappert and co-workers.

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 Grutzmacher and co-workers via reduction of phosphonium salt. Grutzmacher 2001.png
Synthesis of the first stable diphosphanyl radical [Mes*MeP-PMes*] by Grützmacher and co-workers via reduction of phosphonium salt.

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(u3-NtBu)3{u3-Li(thf)}3X] (X = Br, I) by Armstrong and co-workers via oxidation. Armstrong2004-corrected.png
Synthesis of [Me3SiNP(µ3-NtBu)33-Li(thf)}3X] (X = Br, I) by Armstrong and co-workers via oxidation.

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. Ito2006.png
Synthesis of air tolerant and air stable 1,3-diphosphayclobutenyl radical by Ito and co-workers via reduction.

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

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. Cummins2007.png
Resonance structures of [P{NV[N(Np)Ar]3}2] showing delocalisation of radical across vanadium and phosphorus nuclei.

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

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 . Illustrating the "Jack-in-the-box" model. Power2001.png
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 . 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

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

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(u-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. Villinger2011.png
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.
Synthesis of (iPr)CP]2 radical via reduction by Rottschafer and co-workers with resonance structures. Rottschafer 2017.png
Synthesis of (iPr)CP]2 radical via reduction by Rottschafer and co-workers with resonance structures.

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

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(u-NR)]2 (R=Hyp, Ter) radical. Villinger reactivity.png
Reactivity of [P(µ-NR)]2 (R=Hyp, Ter) radical.

Characterisation

Solvation of lithium ions in [Me3SiNP(u3-NtBu)3{u3-Li(thf)}3I] in very dilute THF solutions. Armstrong2004 2.png
Solvation of lithium ions in [Me3SiNP(µ3-NtBu)33-Li(thf)}3I] in very dilute THF solutions.

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

Synthesis

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. Bertand2010.png
Synthesis of [(cAAC)2P2]•+ and [(NHC)2P2]•+ via oxidation with Ph3C B(C6F5)4 by Bertrand and co-workers.
Synthesis of [(TMP)P(cAAC)] via oxidation with Ph3C (C6F5)4B by Bertrand and co-workers. Bertand20100804.png
Synthesis of [(TMP)P(cAAC)] via oxidation with Ph3C (C6F5)4B by Bertrand and co-workers.
Synthesis of [bis(carbene)-PN] visa oxidation with h3C (C6F5)4B by Bertrand and co-workers. Bertrand20100809.png
Synthesis of [bis(carbene)-PN] visa oxidation with h3C (C6F5)4B by Bertrand and co-workers.

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. Geoffroy2003.png
Synthesis of Mes*P -(C(NMe2)2) via a one electron oxidation of a phosphaalkenes with [Cp2Fe]PF6 by Geoffroy and co-workers.

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. Schulz2015.png
General scheme for preparation of cyclic radical cations via oxidation.

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. Ghadwal2019.png
Synthesis of divinyldiphosphene radical cations via oxidation with GaCl3 by Ghadwal and co-workers.

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

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

Synthesis

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. Wang2014.png
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. Wang2016.png
Synthesis of diphosphorus-centred radical anion and the di-radical di-anion via reduction with KC8 by Wang and co-workers.

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 Co and Fe complexes by Tan and co-workers. Tan2017.png
Synthesis of phosphorus radical anion coordinated with Co and Fe complexes by Tan and co-workers.

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. Yamashitaz2020.png
Synthesis of phosphorus radical anion with boryl substituents by Yamashita and co-workers.

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

Reviews

Reactivity

Potential applications

Related Research Articles

In chemistry, a nucleophile is a chemical species that forms bonds by donating an electron pair. All molecules and ions with a free pair of electrons or at least one pi bond can act as nucleophiles. Because nucleophiles donate electrons, they are Lewis bases.

A non-Kekulé molecule is a conjugated hydrocarbon that cannot be assigned a classical Kekulé structure.

5-Dehydro-<i>m</i>-xylylene Chemical compound

5-Dehydro-m-xylylene (DMX) is an aromatic organic triradical and the first known organic molecule to violate Hund's Rule.

[n]Radialenes are alicyclic organic compounds containing n cross-conjugated exocyclic double bonds. The double bonds are commonly alkene groups but those with a carbonyl (C=O) group are also called radialenes. For some members the unsubstituted parent radialenes are elusive but many substituted derivatives are known.

In chemistry, a Zintl phase is a product of a reaction between a group 1 or group 2 and main group metal or metalloid. It is characterized by intermediate metallic/ionic bonding. Zintl phases are a subgroup of brittle, high-melting intermetallic compounds that are diamagnetic or exhibit temperature-independent paramagnetism and are poor conductors or semiconductors.

<span class="mw-page-title-main">Phosphinidene</span> Type of compound

Phosphinidenes are low-valent phosphorus compounds analogous to carbenes and nitrenes, having the general structure RP. The "free" form of these compounds is conventionally described as having a singly-coordinated phosphorus atom containing only 6 electrons in its valence level. Most phosphinidenes are highly reactive and short-lived, thereby complicating empirical studies on their chemical properties. In the last few decades, several strategies have been employed to stabilize phosphinidenes, and researchers have developed a number of reagents and systems that can generate and transfer phosphinidenes as reactive intermediates in the synthesis of various organophosphorus compounds.

<span class="mw-page-title-main">Paramagnetic nuclear magnetic resonance spectroscopy</span> Spectroscopy of paramagnetic compounds via NMR

Paramagnetic nuclear magnetic resonance spectroscopy refers to nuclear magnetic resonance (NMR) spectroscopy of paramagnetic compounds. Although most NMR measurements are conducted on diamagnetic compounds, paramagnetic samples are also amenable to analysis and give rise to special effects indicated by a wide chemical shift range and broadened signals. Paramagnetism diminishes the resolution of an NMR spectrum to the extent that coupling is rarely resolved. Nonetheless spectra of paramagnetic compounds provide insight into the bonding and structure of the sample. For example, the broadening of signals is compensated in part by the wide chemical shift range (often 200 ppm in 1H NMR). Since paramagnetism leads to shorter relaxation times (T1), the rate of spectral acquisition can be high.

<span class="mw-page-title-main">Photoredox catalysis</span>

Photoredox catalysis is a branch of photochemistry that uses single-electron transfer. Photoredox catalysts are generally drawn from three classes of materials: transition-metal complexes, organic dyes, and semiconductors. While organic photoredox catalysts were dominant throughout the 1990s and early 2000s, soluble transition-metal complexes are more commonly used today.

The phosphaethynolate anion, also referred to as PCO, is the phosphorus-containing analogue of the cyanate anion with the chemical formula [PCO] or [OCP]. The anion has a linear geometry and is commonly isolated as a salt. When used as a ligand, the phosphaethynolate anion is ambidentate in nature meaning it forms complexes by coordinating via either the phosphorus or oxygen atoms. This versatile character of the anion has allowed it to be incorporated into many transition metal and actinide complexes but now the focus of the research around phosphaethynolate has turned to utilising the anion as a synthetic building block to organophosphanes.

<span class="mw-page-title-main">Triboracyclopropenyl</span>

The triboracyclopropenyl fragment is a cyclic structural motif in boron chemistry, named for its geometric similarity to cyclopropene. In contrast to nonplanar borane clusters that exhibit higher coordination numbers at boron (e.g., through 3-center 2-electron bonds to bridging hydrides or cations), triboracyclopropenyl-type structures are rings of three boron atoms where substituents at each boron are also coplanar to the ring. Triboracyclopropenyl-containing compounds are extreme cases of inorganic aromaticity. They are the lightest and smallest cyclic structures known to display the bonding and magnetic properties that originate from fully delocalized electrons in orbitals of σ and π symmetry. Although three-membered rings of boron are frequently so highly strained as to be experimentally inaccessible, academic interest in their distinctive aromaticity and possible role as intermediates of borane pyrolysis motivated extensive computational studies by theoretical chemists. Beginning in the late 1980s with mass spectrometry work by Anderson et al. on all-boron clusters, experimental studies of triboracyclopropenyls were for decades exclusively limited to gas-phase investigations of the simplest rings (ions of B3). However, more recent work has stabilized the triboracyclopropenyl moiety via coordination to donor ligands or transition metals, dramatically expanding the scope of its chemistry.

<span class="mw-page-title-main">Trivalent group 14 radicals</span>

A trivalent group 14 radical (also known as a trivalent tetrel radical) is a molecule that contains a group 14 element (E = C, Si, Ge, Sn, Pb) with three bonds and a free radical, having the general formula of R3E•. Such compounds can be categorized into three different types, depending on the structure (or equivalently the orbital in which the unpaired electron resides) and the energetic barrier to inversion. A molecule that remains rigidly in a pyramidal structure has an electron in a sp3 orbital is denoted as Type A. A structure that is pyramidal, but flexible, is denoted as Type B. And a planar structure with an electron that typically would reside in a pure p orbital is denoted as Type C. The structure of such molecules has been determined by probing the nature of the orbital that the unpaired electron resides in using spectroscopy, as well as directly with X-ray methods. Trivalent tetrel radicals tend to be synthesized from their tetravalent counterparts (i.e. R3EY where Y is a species that will dissociate).

<span class="mw-page-title-main">Nontrigonal pnictogen compounds</span>

Nontrigonal pnictogen compounds refer to tricoordinate trivalent pnictogen compounds that are not of typical trigonal pyramidal molecular geometry. By virtue of their geometric constraint, these compounds exhibit distinct electronic structures and reactivities, which bestow on them potential to provide unique nonmetal platforms for bond cleavage reactions.

<span class="mw-page-title-main">Hexaphosphabenzene</span> Chemical compound

Hexaphosphabenzene is a valence isoelectronic analogue of benzene and is expected to have a similar planar structure due to resonance stabilization. Although several other allotropes of phosphorus are stable, no evidence for the existence of P6 has been reported. Preliminary ab initio calculations on the trimerisation of P2 leading to the formation of the cyclic P6 were performed, and it was predicted that hexaphosphabenzene would decompose to free P2 with an energy barrier of 13−15.4 kcal mol−1, and would therefore not be observed in the uncomplexed state under normal experimental conditions. The presence of an added solvent, such as ethanol, might lead to the formation of intermolecular hydrogen bonds which may block the destabilizing interaction between phosphorus lone pairs and consequently stabilize P6. The moderate barrier suggests that hexaphosphabenzene could be synthesized from a [2+2+2] cycloaddition of three P2 molecules. Currently, this is a synthetic endeavour which remains to be conquered.

Among pnictogen group Lewis acidic compounds, unusual lewis acidity of Lewis acidic antimony compounds have long been exploited as both stable conjugate acids of non-coordinating anions, and strong Lewis acid counterparts of well-known superacids. Also, Lewis-acidic antimony compounds have recently been investigated to extend the chemistry of boron because of the isolobal analogy between the vacant p orbital of borane and σ*(Sb–X) orbitals of stiborane, and the similar electronegativities of antimony (2.05) and boron (2.04).

<span class="mw-page-title-main">Allotropes of arsenic</span>

Arsenic in the solid state can be found as gray, black, or yellow allotropes. These various forms feature diverse structural motifs, with yellow arsenic enabling the widest range of reactivity. In particular, reaction of yellow arsenic with main group and transition metal elements results in compounds with wide-ranging structural motifs, with butterfly, sandwich and realgar-type moieties featuring most prominently.

Aluminium(I) nucleophiles are a group of inorganic and organometallic nucleophilic compounds containing at least one aluminium metal center in the +1 oxidation state with a lone pair of electrons strongly localized on the aluminium(I) center.

Intrinsic bond orbitals (IBO) are localized molecular orbitals giving exact and non-empirical representations of wave functions. They are obtained by unitary transformation and form an orthogonal set of orbitals localized on a minimal number of atoms. IBOs present an intuitive and unbiased interpretation of chemical bonding with naturally arising Lewis structures. For this reason IBOs have been successfully employed for the elucidation of molecular structures and electron flow along the intrinsic reaction coordinate (IRC). IBOs have also found application as Wannier functions in the study of solids.

<span class="mw-page-title-main">Polyfluoroalkoxyaluminates</span> Class of chemical compounds

Polyfluoroalkoxyaluminates (PFAA) are weakly coordinating anions many of which are of the form [Al(ORF)4]. Most PFAA's possesses an Al(III) center coordinated by four ORF (RF = -CPh(CF3)2 (hfpp), -CH(CF3)2 (hfip), -C(CH3)(CF3)2 (hftb), -C(CF3)3 (pftb)) ligands, giving the anion an overall -1 charge. The most weakly coordinating PFAA is an aluminate dimer, [F{Al(Opftb)3}2], which possess a bridging fluoride between two Al(III) centers. The first PFAA, [Al(Ohfpp)4], was synthesized in 1996 by Steven Strauss, and several other analogs have since been synthesized, including [Al(Ohfip)4], [Al(Ohftb)4], and [Al(Opftb)4] by Ingo Krossing in 2001. These chemically inert and very weakly coordinating ions have been used to stabilize unusual cations, isolate reactive species, and synthesize strong Brønsted acids.

<span class="mw-page-title-main">Tetraethylammonium trichloride</span> Chemical compound

Tetraethylammonium trichloride (also known as Mioskowski reagent) is a chemical compound with the formula [NEt4][Cl3] consisting of a tetraethylammonium cation and a trichloride as anion. The trichloride is also known as trichlorine monoanion representing one of the simplest polychlorine anions. Tetraethylammonium trichloride is used as reagent for chlorinations and oxidation reactions.

<span class="mw-page-title-main">Carbones</span> Class of molecules

Carbones are a class of molecules containing a carbon atom in the 1D excited state with a formal oxidation state of zero where all four valence electrons exist as unbonded lone pairs. These carbon-based compounds are of the formula CL2 where L is a strongly σ-donating ligand, typically a phosphine (carbodiphosphoranes) or a N-heterocyclic carbene/NHC (carbodicarbenes), that stabilises the central carbon atom through donor-acceptor bonds. Carbones possess high-energy orbitals with both σ- and π-symmetry, making them strong Lewis bases and strong π-backdonor substituents. Carbones possess high proton affinities and are strong nucleophiles which allows them to function as ligands in a variety of main group and transition metal complexes. Carbone-coordinated elements also exhibit a variety of different reactivities and catalyse various organic and main group reactions.  

References

  1. 1 2 3 4 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. 1 2 3 4 5 6 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. 1 2 3 4 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. 1 2 3 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. 1 2 3 4 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. 1 2 3 4 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. 1 2 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 . ISSN   0044-8249.
  13. 1 2 3 4 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. 1 2 3 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. 1 2 3 4 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. 1 2 3 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. 1 2 3 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. 1 2 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. 1 2 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. 1 2 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. 1 2 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. 1 2 3 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. 1 2 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. 1 2 3 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. 1 2 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.