Phosphorus mononitride

Last updated
Phosphorus nitride
Phosphorus-nitride-2D-model.svg
Phosphorus-nitride-3D-balls.png
Phosphorus-nitride-3D-vdW.png
Names
IUPAC name
Azanylidynephosphane
Other names
Phosphorus nitride
Identifiers
3D model (JSmol)
PubChem CID
  • InChI=1S/NP/c1-2
    Key: AOPJVJYWEDDOBI-UHFFFAOYSA-N
  • N#P
Properties
PN
Molar mass 44.981 g·mol−1
Except where otherwise noted, data are given for materials in their standard state (at 25 °C [77 °F], 100 kPa).

Phosphorus mononitride is an inorganic compound with the chemical formula P N. Containing only phosphorus and nitrogen, this material is classified as a binary nitride. From the Lewis structure perspective, it can be represented with a P-N triple bond with a lone pair on each atom. It is isoelectronic with N2, CO, P2, CS and SiO.

Contents

The compound is highly unstable in standard conditions, tending to rapidly self polymerize. It can be isolated within argon and krypton matrices at 10 K (−263.1 °C). [1] Due to its instability, documentation of reactions with other molecules is limited. [2] Most of its reactivity has thus far been probed and studied at transition metal centers. [3] [4]

Phosphorus mononitride was the first identified phosphorus compound in the interstellar medium [5] and is even thought to be an important molecule in the atmospheres of Jupiter and Saturn. [6] [7]

Discovery and interstellar occurrence

The existence of free, gas-phase phosphorus mononitride was confirmed spectroscopically in 1934 by Nobel laureate, Gerhard Herzberg, and coworkers. [8] J. Curry, L. Herzberg, and G. Herzberg made the accidental discovery after observing new bands in the UV region from 2375-2992 Å [9] following an electric discharge within an air-filled tube that had been earlier exposed to phosphorus.

In 1987, phosphorus mononitride was detected in the Orion KL Nebula, the W51M nebula in Aquila, and Saggitarius B2 simultaneously by Turner, Bally, and Ziurys. Data from radio telescopes allowed for observation of rotational lines associated with the J = 2-1, 3-2, 5-4, and 6-5 transitions. [5] [10]

In the following decades, a rapid expansion of interstellar PN observations ensued, detected frequently alongside PO. Examples include within shocked regions of L1157, [11] [12] within the galactic center, [13] in carbon-rich envelopes in CRL 2688 (alongside HCP) and oxygen-rich envelopes toward VY Canis Majoris, [14] TX Camelopardalis, R. Cassiopeiae, and NML Cygni. [15]

ALMA data alongside spectroscopic measurements from the Rosetta probe have shown PN being carried from the comet 67P/Churyumov–Gerasimenko alongside the far more abundant PO. [16] These observations may offer insight to how pre-biotic matter could be transported to planets. In cases where PN and PO are observed in the same region, the latter is more abundant. [17] The consistency of the molecular ratio between these two interstellar molecules across many different interstellar clouds is thought to be a sign of a shared formation pathway between the two molecules. [18] PN is mostly detected in hot, turbulent regions, where the shock induced sputtering of dust grain is thought to contribute to its formation. However, it has also been confirmed in massive dense cores which are by comparison "cold and quiescent". [19] [20]

In 2022, researchers used data from the ALMA Comprehensive High-resolution Extragalactic Molecular Inventory (ALCHEMI) project and reported evidence of phosphorus mononitride in giant molecular clouds within the galaxy, NGC 253. This finding marks phosphorus mononitride as the first extragalactic phosphorus containing molecule detected as well. [17] In 2023, Ziurys and coworkers showed the existence of PN and PO in WB89-621 (22.6 kpc from the galactic center) using rotational spectroscopy. Prior, phosphorus was only observed in the inner Milky Way (12kpc). Since supernovae do not occur in outer regions of the galaxy, the detection of these phosphorus-bearing molecules in WB89-621 provides evidence of additional alternative sources of phosphorus formation, such as non-explosive, lower mass asymptotic giant branch stars. The levels were detected at comparable values to that in the Solar system. [21]

Electronic structure, spectral and bonding properties

PN 2D electron density Laplacian contour plot. Calculated analogously to Kupka et al. at the CCSD/aug-cc-pCVQZ level. PN 2D Electron Density Laplacian Contour.png
PN 2D electron density Laplacian contour plot. Calculated analogously to Kupka et al. at the CCSD/aug-cc-pCVQZ level.
N2 2D electron density Laplacian contour plot. Calculated analogously to Kupka et al. at the CCSD/aug-cc-pCVQZ level. N2 2D Electron Density Laplacian Contour.png
N2 2D electron density Laplacian contour plot. Calculated analogously to Kupka et al. at the CCSD/aug-cc-pCVQZ level.

PN formation from gaseous phosphorus and nitrogen is endothermic.

½ P2 + ½ N2 = PN (ER = 117 ± 10kJ/mol) [25]

Early mass spectrometry studies by Gingerich yielded a PN dissociation energy D0 of 146.6 ± 5.0 kcal/mol (613 ± 21 kJ/mol; 6.36 ± 0.22 eV). [26]

It is predicted to have a high proton affinity (PA = 191 kcal/mol (800 kJ/mol)). [27]

Early rotational analysis of 24 of the bands from Herzberg's original study suggested a PN internuclear distance of 1.49 Å, intermediate between N2 (1.094 Å) and P2 (1.856 Å). The associated electronic transition, 1Π → 1Σ, was noted to be similar to that of the isoelectronic CS and SiO molecules. [28] Later rotational spectra studies aligned well with these findings, for example analysis of millimeter wave rotational PN spectra from a microwave spectrometer yielded a bond distance of 1.49085 (2) Å. [29] [30]

Infrared studies of gaseous PN at high temperatures assign its vibrational frequency (ωe) to 1337.24 cm-1 and interatomic separation of 1.4869 Å. [8] [31]

Simple comparisons to tabulated experimental and calculated bond lengths match well with a PN triple bond according to Pyykkö's Triple-Bond Covalent Radii. [32]

NBO analyses support a single neutral resonance structure with a PN triple bond and one lone pair on each atom. However, natural population analysis shows nitrogen as significantly negatively charged (-0.82603) and phosphorus as significantly positively charged (0.82603). This is in line with the large dipole moment and partial ionic character reflecting the electron density contour plots. [33] [34] [35] [36] [37] [24]

Monomeric PN in a krypton matrix at 10 K (−263.1 °C) gives rise to a single IR band at 1323 cm-1. [1]

Auer and Neese have produced calculated gas phase 31P and 15N NMR chemical shifts of 51.61 and -344.71 respectively at the CCSD(T)/p4 level of theory. [38] However, different functionals and basis sets yield dramatically different predictions for chemical shielding and so far experimental NMR shifts for phosphorus mononitride remain elusive. [22]

Molecular beam electric resonance spectroscopy has been used to determine the radio frequency spectrum of phosphorus mononitride generated from P3N5 thermolysis; the experimental results showed an experimental PN dipole moment (μ) of 2.7465 +- 0.0001 D, 2.7380 +-0.001 D, and 2.7293 +-0.0001 D for the first three vibrational levels respectively. [39]

Occupied intrinsic bond orbitals (IBOs) of PN. Example calculated at the PBE0-D3BJ level of theory using the def2-TZVP basis set with ORCA. (Left): Two lone pairs + one P-N sigma bond. (Middle + Right): Degenerate perpendicular pi bond pair. PN IBOs Occupied.png
Occupied intrinsic bond orbitals (IBOs) of PN. Example calculated at the PBE0-D3BJ level of theory using the def2-TZVP basis set with ORCA. (Left): Two lone pairs + one P-N sigma bond. (Middle + Right): Degenerate perpendicular pi bond pair.
Valence virtual IBOs of PN. Example calculated at the PBE0-D3BJ level of theory using the def2-TZVP basis set with ORCA. (Top): Degenerate pi antibonding orbitals relevant to pi backbonding (in relation to LUMO). PN IBOs Unoccopied Virtual.png
Valence virtual IBOs of PN. Example calculated at the PBE0-D3BJ level of theory using the def2-TZVP basis set with ORCA. (Top): Degenerate pi antibonding orbitals relevant to pi backbonding (in relation to LUMO).

Its dipole moment is larger than PO (1.88 D), despite the greater electronegativity difference between the constituent P and O atoms and similar bond length (1.476 Å). [41] This a result of the significant differences in bonds and charge distribution within the PN and PO molecules. The large PN dipole moment makes it very favorable with respect to radio-astronomical studies in comparison to N2 - which lacks this property. [41]

In consideration to molecular orbitals of PN, direct analogies can be drawn to the bonding in the N2 molecule. It consists of an P-N σ bonding orbital (HOMO), with two perpendicular degenerate P-N pi bonding orbitals. Likewise, the LUMOs of PN, which consist of a degenerate PN pi-antibonding set, allow it to backbond with orbitals of appropriate symmetry.

However, in comparison to N2, the HOMO of PN is higher in energy (est. -9.2 eV vs -12.2 eV), and, the LUMOs are lower in energy (-2.3 eV vs -0.6 eV), thus making it both a better σ-donor and pi-acceptor as a ligand. [42]

Evidently, the smaller HOMO-LUMO gap of PN, combined with its polar nature and low dissociation energy contribute to its much greater reactivity than dinitrogen (including at the interstellar level). [43]

Preparation and formation

Interstellar formation

The pathways to the formation of PN are still not fully understood, but likely involve competing gaseous phase reactions with other interstellar molecules. Important schemes are shown below along with competing exothermic reactions: [44]

PO + N → PN + O

PO + N → P + NO (Competing)

Another important, very exothermic formation reaction:

PH + N → PN + H

From carbon containing environments: [45]

P + CN → PN + C

N + CP → PN + C

An important destruction pathway:

PN + N → N2 + P

The abundance of interstellar PN is additionally perturbed by cosmic-ray ionization, visual extinction, and adsorption/desorption from dust grains. [45] [46]

Electric discharge

Moldenhauer and Dörsam first generated transient PN in 1924 using an electric discharge through N2 and phosphorus vapors, where the characterized product was a notably robust powder containing equal parts phosphorus and nitrogen. [47] This same method led to the actual first observation of PN by Gerhard and coworkers.

PN has also been produced at room temperature using microwave discharges on mixtures of gaseous PCl3 and N2 under moderate vacuum. This preparation was employed to achieve high resolution FTIR spectra of PN. [48]

Flash pyrolysis

Atkins and Timms later generated PN via flash pyrolysis of P3N5 under high vacuum, allowing the recording of the PN infrared spectrum within a cryogenic krypton matrix. Solid triphosphorus pentanitride generates gaseous, free PN when heated to 800–900 °C (1,070–1,170 K) under high vacuum. [1]

P3N5 flash pyrolysis methodology from Atkins and Timms. Depicts intermediate(s) in matrix isolation, and polymerization. P3N5 pyrolysis scheme.png
P3N5 flash pyrolysis methodology from Atkins and Timms. Depicts intermediate(s) in matrix isolation, and polymerization.

Monomeric PN can only be isolated in krypton or argon matrices at 10 K (−263.1 °C). Upon warming up past 30 K (−243.2 °C), cyclotriphosphazene, which has D3h symmetry, is formed (up to 50 K (−223.2 °C) before krypton matrix melts). The (PN)3 trimer and is planar and aromatic, with 15N-labelling experiments revealing a planar E' mode band at 1141 cm-1. [25] No dimers or other oligomers are even transiently observed.

Without a cryoscopic matrix, these reactions result in the immediate formation of (PN)n polymers. [1]

Thermolysis experiments of dimethyl phosphoramidate have shown PN to form as a major decomposition product along with many other minor components including the ·P=O radical and HOP=O. This is contrasting to dimethyl methylphosphonate in which said minor components become the major decomposition products, highlighting significantly diverging pathways. [49]

Phosphinoazide pyrolysis to PN and byproducts from Qian, Wende, Schreiner, and Mardyukov. At lower pyrolysis temperatures, a different major product forms. Phosphino Azide Pyrolysis to PN.png
Phosphinoazide pyrolysis to PN and byproducts from Qian, Wende, Schreiner, and Mardyukov. At lower pyrolysis temperatures, a different major product forms.

In 2023, Qian et al. proposed PN to be generated as a major product along with CO and cyclopentadienone byproducts when (o-phenyldioxyl)phosphinoazide is heated to 850 °C (following the loss of N2). However, efforts to observe free PN in argon matrixes using this method were unsuccessful due to band overlaps. [2]

Dehalogenation of hexachlorophosphazene

Dehalogenation reaction of hexachlorophosphazene using molten silver from Schnockel et al. Hexachlorophosphazene Dehalogenation.png
Dehalogenation reaction of hexachlorophosphazene using molten silver from Schnöckel et al.

Schnöckel and coworkers later showed an alternative synthesis involving the dehalogenation of hexachlorophosphazene with molten silver, with concomitant loss of AgCl. In both this route and the P3N5 thermolysis route, only trace P2 and P4 formation is detected even at 1,200 K (930 °C), showing the reaction temperatures occur far from thermodynamic equilibrium. [25]

Anthracene release from dibenzo-7λ3 -phosphanorbornadiene derivatives

Room temperature PN release reagent, N3PA, from Cummins and coworkers. N3PA Dissociation.png
Room temperature PN release reagent, N3PA, from Cummins and coworkers.

The aforementioned methods require very high temperatures which are incompatible with standard, homogeneous solution state chemistry.

In 2022, Cummins and coworkers prepared and isolated a molecular PN precursor, N3PA which rapidly decomposes to N2, anthracene, and PN in solution at room temperature (t½ = 30 minutes). With the combination of vacuum and heating to 42 °C, this dissociation is explosive. [42]

Reactivity

Reactions of phosphorus mononitride with other molecules are rare and rather difficult to carry out. The formation of the intermediate (PN)3 trimer (which itself is only isolated in matrices) is highly favorable:

3PN ⇌ (PN)3 (-334 +/- 60 kJ/mol) [25]

PN generated in both the gaseous phase or in solution that is not subjected to trapping via noble gas matrices or particular metal complexes results in rapid self polymerization even in cases where trapping agents such as dienes or alkynes are present (differentiating its reactivity profile from related molecules such as P2). [50] [51]

Light driven reversible PN oxidation to a phosphinonitrene at 10 K from Qian et al. Phosphino nitrene PN reversible oxidation.png
Light driven reversible PN oxidation to a phosphinonitrene at 10 K from Qian et al.

Phosphorus mononitride's tendency to rapidly polymerize with itself has dominated its reactivity, greatly hindering both the study and diversity of products in its reactions with organic molecules.

In 2023, a rare case of documented reactivity with an organic molecule was reported by Qian and coworkers who demonstrated reversible photoisomerization between o-benzoquinone supported phosphinonitrene and o-benzoquinone stabilized phosphorus mononitride at 10 K, which can be isolated in an argon matrix. [2]

Ligation, stabilization, and reactivity at transition metals

The majority of documented well-defined PN reactivity has been carried out at transition metal centers. The electronic and molecular orbital similarities it shares with N2 make it a viable ligating species. While free PN is unstable, phosphorus mononitride has been prepared at metal coordination sites where it can exist as an isolable terminal ligand within a complex. [3] [42] In alternative cases, PN ligands can also exist as only as transient, highly reactive intermediates featuring rich chemistry. [4] As a terminal ligand, cases of both preferential P and N bonding modes have been discovered.

[MoPN] and [MoNP] preparation and photoisomerization from Smith and coworkers. MoPN and MoNP.png
[MoPN] and [MoNP] preparation and photoisomerization from Smith and coworkers.

Smith and co-workers isolated the first stable M-PN (and M-NP) complexes, using methodology to generate the PN moiety at metal sites. They reacted a tris(amido) Mo(VI) terminal phosphide complex with a tris(carbene)borate Fe(IV) terminal nitride, which undergo reductive coupling to form the corresponding neutral bridging PhB(iPr2Im)3Fe-NP-Mo(N3N) complex. Notably, the Mo-N-P bond angle in the bridging compound is nearly perfectly linear with an N-P bond length of 1.509(6) Å (only slightly elongated from free PN indicating significant multiple bond character). [3] Addition of 3 equivalents of strongly lewis basic tert-butyl isocyanide results in the release of the iron adduct as a [PhB(iPr2Im)3Fe-(CNtBu)3]+ cation in the second coordination sphere. The corresponding terminal linear Mo-PN anion can be isolated and converted to its linear Mo-NP isomer by exposure to white light in the solid state. The M-NP isomer of the ligand was determined to be more pi-acidic (N-P = 1.5913(1) Å and P-N = 1.5363(1) Å) and more thermodynamically stable than its isomer. [3]

Diamagnetic iron-NP complex from N3PA precursor. PN Transition Metal Capture.png
Diamagnetic iron-NP complex from N3PA precursor.

Cummins and co-workers exploited their N3PA free PN releasing reagent to "trap" and isolate a stable terminal (dppe)(Cp*)Fe-NP complex as a BArF24 salt. The NP bond length in this case was very short at 1.493(2) Å, almost unperturbed from gaseous PN, which is consistent with minimal pi-backbonding from the iron center. Studies confirmed the NP binding mode (as opposed to PN) to be energetically preferred by 36.6 kcal/mol (153 kJ/mol) in this iron complex, creating a significant barrier to isomerization (thought to arise from Pauli repulsion effects). [42] Studies of phosphorus mononitride chemistry at tris(amido) vanadium complexes undertaken by Cummins and coworkers provides the bulk of PN reactivity examples at transition metals to date. In this system, PN is synthetically generated at a vanadium center from respective dibenzo-7λ3 -phosphanorbornadiene derivative precursors. However, it is not stable as a terminal ligand, and instead immediately undergoes trimerization. Notably, a thermodynamic equilibrium exists between this trimer species, along with a dimer and non-observed monomeric intermediate fragment. [4] [50]

Synthesis of tris(amido) vanadium PN trimer, dimer, and corresponding monomeric V-NP reactive intermediate from Cummins and coworkers. Vanadium PN Compounds Preparation.png
Synthesis of tris(amido) vanadium PN trimer, dimer, and corresponding monomeric V-NP reactive intermediate from Cummins and coworkers.



The V-NP fragment undergoes singlet phosphinidene reactivity ([2+1] additions) with alkene and alkyne trapping agents, generating phosphiranes and phospherenes respectively. The products generated from such additions exist in equilibrium (in the case with cis-4-octene and bis-trimethylsilylacetylene), where retention of the cis-4-octene conformer is observed. Upon heating, they reversibly add to generate the V-NP dimer. Such reactivity demonstrates stark contrasts from P2 as a ligand which instead undergoes formal cycloaddition chemistry. [4]

Reactivity of PN generated at vanadium tris(amido) complexes. Reversible singlet phosphinidene reactivity and equilibrium. Vanadium NP Reactivity and Phosphinidene Transfer.png
Reactivity of PN generated at vanadium tris(amido) complexes. Reversible singlet phosphinidene reactivity and equilibrium.



Applications

The robust nature of PN reaction products such as (PN)n, could find use in heat resistant ceramics or as fire suppressing materials. [52]

There has long been interest in studying PN and its reaction products like (PN)n polymers, noting their relevance to precursors/intermediates in the production of fertilizers. [51] [53]

See also

Related Research Articles

<span class="mw-page-title-main">Nitrogen</span> Chemical element, symbol N and atomic number 7

Nitrogen is a chemical element; it has symbol N and atomic number 7. Nitrogen is a nonmetal and the lightest member of group 15 of the periodic table, often called the pnictogens. It is a common element in the universe, estimated at seventh in total abundance in the Milky Way and the Solar System. At standard temperature and pressure, two atoms of the element bond to form N2, a colorless and odorless diatomic gas. N2 forms about 78% of Earth's atmosphere, making it the most abundant uncombined element in air. Because of the volatility of nitrogen compounds, nitrogen is relatively rare in the solid parts of the Earth.

<span class="mw-page-title-main">Hydroxyl radical</span> Neutral form of the hydroxide ion (OH−)

The hydroxyl radical, HO, is the neutral form of the hydroxide ion (HO). Hydroxyl radicals are highly reactive and consequently short-lived; however, they form an important part of radical chemistry. Most notably hydroxyl radicals are produced from the decomposition of hydroperoxides (ROOH) or, in atmospheric chemistry, by the reaction of excited atomic oxygen with water. It is also an important radical formed in radiation chemistry, since it leads to the formation of hydrogen peroxide and oxygen, which can enhance corrosion and SCC in coolant systems subjected to radioactive environments. Hydroxyl radicals are also produced during UV-light dissociation of H2O2 (suggested in 1879) and likely in Fenton chemistry, where trace amounts of reduced transition metals catalyze peroxide-mediated oxidations of organic compounds.

In chemistry, a hypervalent molecule is a molecule that contains one or more main group elements apparently bearing more than eight electrons in their valence shells. Phosphorus pentachloride, sulfur hexafluoride, chlorine trifluoride, the chlorite ion, and the triiodide ion are examples of hypervalent molecules.

In chemistry, a nitride is an inorganic compound of nitrogen. The "nitride" anion, N3- ion, is very elusive but compounds of nitride are numerous, although rarely naturally occurring. Some nitrides have a found applications, such as wear-resistant coatings (e.g., titanium nitride, TiN), hard ceramic materials (e.g., silicon nitride, Si3N4), and semiconductors (e.g., gallium nitride, GaN). The development of GaN-based light emitting diodes was recognized by the 2014 Nobel Prize in Physics. Metal nitrido complexes are also common.

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

In chemistry, a phosphaalkyne is an organophosphorus compound containing a triple bond between phosphorus and carbon with the general formula R-C≡P. Phosphaalkynes are the heavier congeners of nitriles, though, due to the similar electronegativities of phosphorus and carbon, possess reactivity patterns reminiscent of alkynes. Due to their high reactivity, phosphaalkynes are not found naturally on earth, but the simplest phosphaalkyne, phosphaethyne (H-C≡P) has been observed in the interstellar medium.

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

Magnesium nitride, which possesses the chemical formula Mg3N2, is an inorganic compound of magnesium and nitrogen. At room temperature and pressure it is a greenish yellow powder.

Diphosphorus is an inorganic chemical with the chemical formula P
2
. Unlike nitrogen, its lighter pnictogen neighbor which forms a stable N2 molecule with a nitrogen to nitrogen triple bond, phosphorus prefers a tetrahedral form P4 because P-P pi-bonds are high in energy. Diphosphorus is, therefore, very reactive with a bond-dissociation energy (117 kcal/mol or 490 kJ/mol) half that of dinitrogen. The bond distance has been measured at 1.8934 Å.

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

Uranium nitrides is any of a family of several ceramic materials: uranium mononitride (UN), uranium sesquinitride (U2N3) and uranium dinitride (UN2). The word nitride refers to the −3 oxidation state of the nitrogen bound to the uranium.

<span class="mw-page-title-main">Transition metal dinitrogen complex</span> Coordination compounds with N2

Transition metal dinitrogen complexes are coordination compounds that contain transition metals as ion centers the dinitrogen molecules (N2) as ligands.

<span class="mw-page-title-main">Allotropes of phosphorus</span> Solid forms of the element phosphorus

Elemental phosphorus can exist in several allotropes, the most common of which are white and red solids. Solid violet and black allotropes are also known. Gaseous phosphorus exists as diphosphorus and atomic phosphorus.

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

Diazenylium is the chemical N2H+, an inorganic cation that was one of the first ions to be observed in interstellar clouds. Since then, it has been observed for in several different types of interstellar environments, observations that have several different scientific uses. It gives astronomers information about the fractional ionization of gas clouds, the chemistry that happens within those clouds, and it is often used as a tracer for molecules that are not as easily detected (such as N2). Its 1–0 rotational transition occurs at 93.174 GHz, a region of the spectrum where Earth's atmosphere is transparent and it has a significant optical depth in both cold and warm clouds so it is relatively easy to observe with ground-based observatories. The results of N2H+ observations can be used not only for determining the chemistry of interstellar clouds, but also for mapping the density and velocity profiles of these clouds.

Sulfur mononitride is an inorganic compound with the molecular formula SN. It is the sulfur analogue of and isoelectronic to the radical nitric oxide, NO. It was initially detected in 1975, in outer space in giant molecular clouds and later the coma of comets. This spurred further laboratory studies of the compound. Synthetically, it is produced by electric discharge in mixtures of nitrogen and sulfur compounds, or combustion in the gas phase and by photolysis in solution.

<span class="mw-page-title-main">Imidogen</span> Inorganic radical with the chemical formula NH

Imidogen is an inorganic compound with the chemical formula NH. Like other simple radicals, it is highly reactive and consequently short-lived except as a dilute gas. Its behavior depends on its spin multiplicity.

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

Triphosphorus pentanitride is an inorganic compound with the chemical formula P3N5. Containing only phosphorus and nitrogen, this material is classified as a binary nitride. While it has been investigated for various applications this has not led to any significant industrial uses. It is a white solid, although samples often appear colored owing to impurities.

<span class="mw-page-title-main">Solid nitrogen</span> Solid form of the 7th element

Solid nitrogen is a number of solid forms of the element nitrogen, first observed in 1884. Solid nitrogen is mainly the subject of academic research, but low-temperature, low-pressure solid nitrogen is a substantial component of bodies in the outer Solar System and high-temperature, high-pressure solid nitrogen is a powerful explosive, with higher energy density than any other non-nuclear material.

Argon compounds, the chemical compounds that contain the element argon, are rarely encountered due to the inertness of the argon atom. However, compounds of argon have been detected in inert gas matrix isolation, cold gases, and plasmas, and molecular ions containing argon have been made and also detected in space. One solid interstitial compound of argon, Ar1C60 is stable at room temperature. Ar1C60 was discovered by the CSIRO.

<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">Phosphorus monoxide</span> Chemical compound

Phosphorus monoxide is an unstable radical inorganic compound with molecular formula PO.

The inorganic imides are compounds containing an ion composed of nitrogen bonded to hydrogen with formula HN2−. Organic imides have the NH group, and two single or one double covalent bond to other atoms. The imides are related to the inorganic amides (H2N), the nitrides (N3−) and the nitridohydrides (N3−•H).

References

  1. 1 2 3 4 5 Atkins, Robert M.; Timms, Peter L. (1977). "The matrix infrared spectrum of PN and SiS". Spectrochimica Acta Part A: Molecular Spectroscopy. 33 (9): 853–857. Bibcode:1977AcSpA..33..853A. doi:10.1016/0584-8539(77)80083-4. ISSN   0584-8539.
  2. 1 2 3 4 5 Qian, Weiyu; Wende, Raffael C.; Schreiner, Peter R.; Mardyukov, Artur (2023-04-18). "Selective Preparation of Phosphorus Mononitride (P≡N) from Phosphinoazide and Reversible Oxidation to Phosphinonitrene". Angewandte Chemie International Edition. 62 (23): e202300761. doi: 10.1002/anie.202300761 . ISSN   1433-7851. PMID   36877095.
  3. 1 2 3 4 5 Martinez, Jorge L.; Lutz, Sean A.; Beagan, Daniel M.; Gao, Xinfeng; Pink, Maren; Chen, Chun-Hsing; Carta, Veronica; Moënne-Loccoz, Pierre; Smith, Jeremy M. (2020-09-01). "Stabilization of the Dinitrogen Analogue, Phosphorus Nitride". ACS Central Science. 6 (9): 1572–1577. doi: 10.1021/acscentsci.0c00944 . ISSN   2374-7943. PMC   7517109 . PMID   32999932.
  4. 1 2 3 4 5 6 Courtemanche, Marc-André; Transue, Wesley J.; Cummins, Christopher C. (2016-12-21). "Phosphinidene Reactivity of a Transient Vanadium P≡N Complex". Journal of the American Chemical Society. 138 (50): 16220–16223. doi: 10.1021/jacs.6b10545 . ISSN   0002-7863. PMID   27958729.
  5. 1 2 Turner, B. E.; Bally, John (1987-10-01). "Detection of Interstellar PN: The First Identified Phosphorus Compound in the Interstellar Medium". The Astrophysical Journal. 321: L75. Bibcode:1987ApJ...321L..75T. doi:10.1086/185009. ISSN   0004-637X.
  6. Viana, Rommel B.; Pereira, Priscila S. S.; Macedo, Luiz G. M.; Pimentel, André S. (2009-09-18). "A quantum chemical study on the formation of phosphorus mononitride". Chemical Physics. 363 (1): 49–58. Bibcode:2009CP....363...49V. doi:10.1016/j.chemphys.2009.07.008. ISSN   0301-0104.
  7. Yung, Yuk L.; DeMore, William B. (1999-03-11), "Origins", Photochemistry of Planetary Atmospheres, Oxford University Press, doi:10.1093/oso/9780195105018.003.0007, ISBN   978-0-19-510501-8 , retrieved 2023-11-28
  8. 1 2 Curry, J.; Herzberg, L.; Herzberg, G. (1933-10-01). "Spectroscopic Evidence for the Molecule PN". The Journal of Chemical Physics. 1 (10): 749. Bibcode:1933JChPh...1..749C. doi:10.1063/1.1749238. ISSN   0021-9606.
  9. Herzberg, G. (1972-07-14). "Spectroscopic Studies of Molecular Structure". Science. 177 (4044): 123–138. Bibcode:1972Sci...177..123H. doi:10.1126/science.177.4044.123. ISSN   0036-8075. PMID   17779905.
  10. Ziurys, L. M. (1987). "Detection of interstellar PN - The first phosphorus-bearing species observed in molecular clouds". The Astrophysical Journal. 321 (1 Pt 2): L81-5. Bibcode:1987ApJ...321L..81Z. doi:10.1086/185010. ISSN   0004-637X. PMID   11542218.
  11. Yamaguchi, Takahiro; Takano, Shuro; Sakai, Nami; Sakai, Takeshi; Liu, Sheng-Yuan; Su, Yu-Nung; Hirano, Naomi; Takakuwa, Shigehisa; Aikawa, Yuri; Nomura, Hideko; Yamamoto, Satoshi (2011-10-25). "Detection of Phosphorus Nitride in the Lynds 1157 B1 Shocked Region". Publications of the Astronomical Society of Japan. 63 (5): L37–L41. doi: 10.1093/pasj/63.5.l37 . ISSN   0004-6264.
  12. Lefloch, Bertrand; Vastel, C.; Viti, S.; Jimenez-Serra, I.; Codella, C.; Podio, L.; Ceccarelli, C.; Mendoza, E.; Lepine, J. R. D.; Bachiller, R. (2016-08-02). "Phosphorus-bearing molecules in solar-type star-forming regions: first PO detection". Monthly Notices of the Royal Astronomical Society. 462 (4): 3937–3944. arXiv: 1608.00048 . doi: 10.1093/mnras/stw1918 . ISSN   0035-8711.
  13. Rivilla, V M; Jiménez-Serra, I; Zeng, S; Martín, S; Martín-Pintado, J; Armijos-Abendaño, J; Viti, S; Aladro, R; Riquelme, D; Requena-Torres, M; Quénard, D; Fontani, F; Beltrán, M T (2018-01-05). "Phosphorus-bearing molecules in the Galactic Center". Monthly Notices of the Royal Astronomical Society: Letters. 475 (1): L30–L34. arXiv: 1712.07006 . doi:10.1093/mnrasl/slx208. ISSN   1745-3925.
  14. Milam, S. N.; Halfen, D. T.; Tenenbaum, E. D.; Apponi, A. J.; Woolf, N. J.; Ziurys, L. M. (2008). "Constraining Phosphorus Chemistry in Carbon- and Oxygen-Rich Circumstellar Envelopes: Observations of PN, HCP, and CP". The Astrophysical Journal. 684 (1): 618–625. Bibcode:2008ApJ...684..618M. doi: 10.1086/589135 . ISSN   0004-637X.
  15. Ziurys, L. M.; Schmidt, D. R.; Bernal, J. J. (2018-04-05). "New Circumstellar Sources of PO and PN: The Increasing Role of Phosphorus Chemistry in Oxygen-rich Stars". The Astrophysical Journal. 856 (2): 169. Bibcode:2018ApJ...856..169Z. doi: 10.3847/1538-4357/aaafc6 . hdl: 10150/627635 . ISSN   1538-4357.
  16. Rivilla, V M; Drozdovskaya, M N; Altwegg, K; Caselli, P; Beltrán, M T; Fontani, F; van der Tak, F F S; Cesaroni, R; Vasyunin, A; Rubin, M; Lique, F; Marinakis, S; Testi, L; Balsiger, H; Berthelier, J J (2020-01-15). "ALMA and ROSINA detections of phosphorus-bearing molecules: the interstellar thread between star-forming regions and comets". Monthly Notices of the Royal Astronomical Society. 492 (1): 1180–1198. arXiv: 1911.11647 . doi: 10.1093/mnras/stz3336 . ISSN   0035-8711.
  17. 1 2 Haasler, D.; Rivilla, V. M.; Martín, S.; Holdship, J.; Viti, S.; Harada, N.; Mangum, J.; Sakamoto, K.; Muller, S.; Tanaka, K.; Yoshimura, Y.; Nakanishi, K.; Colzi, L.; Hunt, L.; Emig, K. L. (2022). "First extragalactic detection of a phosphorus-bearing molecule with ALCHEMI: Phosphorus nitride (PN)". Astronomy & Astrophysics. 659: A158. arXiv: 2112.04849 . Bibcode:2022A&A...659A.158H. doi:10.1051/0004-6361/202142032. ISSN   0004-6361.
  18. Bernal, J. J.; Koelemay, L. A.; Ziurys, L. M. (2021-01-01). "Detection of PO in Orion-KL: Phosphorus Chemistry in the Plateau Outflow". The Astrophysical Journal. 906 (1): 55. Bibcode:2021ApJ...906...55B. doi: 10.3847/1538-4357/abc87b . ISSN   0004-637X.
  19. Fontani, F.; Rivilla, V. M.; Caselli, P.; Vasyunin, A.; Palau, A. (2016-05-06). "Phosphorus-Bearing Molecules in Massive Dense Cores". The Astrophysical Journal. 822 (2): L30. arXiv: 1604.02565 . Bibcode:2016ApJ...822L..30F. doi: 10.3847/2041-8205/822/2/L30 . ISSN   2041-8213.
  20. Mininni, C; Fontani, F; Rivilla, V M; Beltrán, M T; Caselli, P; Vasyunin, A (2018-02-21). "On the origin of phosphorus nitride in star-forming regions". Monthly Notices of the Royal Astronomical Society: Letters. 476 (1): L39–L44. arXiv: 1802.00623 . doi:10.1093/mnrasl/sly026. ISSN   1745-3925.
  21. Koelemay, L. A.; Gold, K. R.; Ziurys, L. M. (2023-11-08). "Phosphorus-bearing molecules PO and PN at the edge of the Galaxy". Nature. 623 (7986): 292–295. Bibcode:2023Natur.623..292K. doi: 10.1038/s41586-023-06616-1 . ISSN   0028-0836. PMC   10632128 . PMID   37938703.
  22. 1 2 3 Kupka, Teobald; Leszczyńska, Małgorzata; Ejsmont, Krzysztof; Mnich, Adrianna; Broda, Małgorzata; Thangavel, Karthick; Kaminský, Jakub (2019-08-23). "Phosphorus mononitride: A difficult case for theory". International Journal of Quantum Chemistry. 119 (24). doi:10.1002/qua.26032. ISSN   0020-7608. S2CID   202072021.
  23. 1 2 Lu, Tian; Chen, Feiwu (2011-12-08). "Multiwfn: A multifunctional wavefunction analyzer". Journal of Computational Chemistry. 33 (5): 580–592. doi:10.1002/jcc.22885. ISSN   0192-8651. PMID   22162017.
  24. 1 2 3 Bader, Richard F W (1990-12-13). Atoms in Molecules. Oxford University PressOxford. doi:10.1093/oso/9780198551683.001.0001. ISBN   978-0-19-855168-3.
  25. 1 2 3 4 5 Ahlrichs, Reinhart; Bär, Michael; Plitt, Harald S.; Schnöckel, Hansgeorg (1989). "The stability of PN and (PN)3. Ab initio calculations and matrix infrared investigations". Chemical Physics Letters. 161 (2): 179–184. doi:10.1016/0009-2614(89)85053-5. ISSN   0009-2614.
  26. Gingerich, Karl A. (1969). "Gaseous phosphorus compounds. III. Mass spectrometric study of the reaction between diatomic nitrogen and phosphorus vapor and dissociation energy of phosphorus mononitride and diatomic phosphorus". The Journal of Physical Chemistry. 73 (8): 2734–2741. doi:10.1021/j100842a047. ISSN   0022-3654.
  27. Adams, N. G.; McIntosh, B. J.; Smith, D. (1990-06-01). "Production of phosphorus-containing molecules in interstellar clouds". Astronomy and Astrophysics. 232: 443. Bibcode:1990A&A...232..443A. ISSN   0004-6361.
  28. JEVONS, W. (1934). "Band Spectrum of PN and its Significance". Nature. 133 (3364): 619–620. Bibcode:1934Natur.133..619J. doi: 10.1038/133619b0 . ISSN   0028-0836.
  29. Hoeft, J.; Tiemann, E.; Törring, T. (1972-04-01). "Rotationsspektrum des PN". Zeitschrift für Naturforschung A. 27 (4): 703–704. Bibcode:1972ZNatA..27..703H. doi: 10.1515/zna-1972-0424 . ISSN   1865-7109.
  30. Wyse, F. C.; Manson, E. L.; Gordy, W. (1972-08-01). "Millimeter Wave Rotational Spectrum and Molecular Constants of 31P14N". The Journal of Chemical Physics. 57 (3): 1106–1108. doi:10.1063/1.1678365. ISSN   0021-9606.
  31. Maki, Arthur G.; Lovas, Frank J. (1981). "The infrared spectrum of 31P14N near 1300 cm−1". Journal of Molecular Spectroscopy. 85 (2): 368–374. doi:10.1016/0022-2852(81)90209-5. ISSN   0022-2852.
  32. Pyykkö, Pekka; Atsumi, Michiko (2009-11-23). "Molecular Double-Bond Covalent Radii for Elements Li–E112". Chemistry – A European Journal. 15 (46): 12770–12779. doi:10.1002/chem.200901472. ISSN   0947-6539. PMID   19856342.
  33. Weinhold, F.; Landis, C.R.; Glendening, E.D. (2016-06-23). "What is NBO analysis and how is it useful?". International Reviews in Physical Chemistry. 35 (3): 399–440. doi:10.1080/0144235x.2016.1192262. ISSN   0144-235X. S2CID   100034050.
  34. 1 2 3 Grimme, Stefan; Ehrlich, Stephan; Goerigk, Lars (2011). "Effect of the damping function in dispersion corrected density functional theory". Journal of Computational Chemistry. 32 (7): 1456–1465. doi:10.1002/jcc.21759. ISSN   0192-8651. PMID   21370243.
  35. 1 2 3 Grimme, Stefan; Antony, Jens; Ehrlich, Stephan; Krieg, Helge (2010-04-16). "A consistent and accurate ab initio parametrization of density functional dispersion correction (DFT-D) for the 94 elements H-Pu". The Journal of Chemical Physics. 132 (15). Bibcode:2010JChPh.132o4104G. doi:10.1063/1.3382344. ISSN   0021-9606. PMID   20423165.
  36. Weigend, Florian; Ahlrichs, Reinhart (2005). "Balanced basis sets of split valence, triple zeta valence and quadruple zeta valence quality for H to Rn: Design and assessment of accuracy". Physical Chemistry Chemical Physics. 7 (18): 3297. Bibcode:2005PCCP....7.3297W. doi:10.1039/b508541a. ISSN   1463-9076. PMID   16240044.
  37. Weigend, Florian (2006). "Accurate Coulomb-fitting basis sets for H to Rn". Physical Chemistry Chemical Physics. 8 (9): 1057. Bibcode:2006PCCP....8.1057W. doi:10.1039/b515623h. ISSN   1463-9076.
  38. Stoychev, Georgi L.; Auer, Alexander A.; Izsák, Róbert; Neese, Frank (2018-02-13). "Self-Consistent Field Calculation of Nuclear Magnetic Resonance Chemical Shielding Constants Using Gauge-Including Atomic Orbitals and Approximate Two-Electron Integrals". Journal of Chemical Theory and Computation. 14 (2): 619–637. doi:10.1021/acs.jctc.7b01006. ISSN   1549-9618. PMID   29301077.
  39. Raymonda, John; Klemperer, William (1971-07-01). "Molecular Beam Electric Resonance Spectrum of 31P14N". The Journal of Chemical Physics. 55 (1): 232–233. doi:10.1063/1.1675513. ISSN   0021-9606.
  40. 1 2 Knizia, Gerald (2013-10-17). "Intrinsic Atomic Orbitals: An Unbiased Bridge between Quantum Theory and Chemical Concepts". Journal of Chemical Theory and Computation. 9 (11): 4834–4843. arXiv: 1306.6884 . doi:10.1021/ct400687b. ISSN   1549-9618. PMID   26583402. S2CID   17717923.
  41. 1 2 Müller, Holger S. P.; Woon, David E. (2013-11-05). "Calculated Dipole Moments for Silicon and Phosphorus Compounds of Astrophysical Interest". The Journal of Physical Chemistry A. 117 (50): 13868–13877. Bibcode:2013JPCA..11713868M. doi:10.1021/jp4083807. ISSN   1089-5639. PMID   24138156.
  42. 1 2 3 4 5 6 7 Eckhardt, André K.; Riu, Martin-Louis Y.; Ye, Mengshan; Müller, Peter; Bistoni, Giovanni; Cummins, Christopher C. (2022). "Taming phosphorus mononitride". Nature Chemistry. 14 (8): 928–934. Bibcode:2022NatCh..14..928E. doi:10.1038/s41557-022-00958-5. ISSN   1755-4349. PMID   35697930. S2CID   249627769.
  43. Bhasi, Priya; Nhlabatsi, Zanele P.; Sitha, Sanyasi (2017). "Reactivity of phosphorus mononitride and interstellar formation of molecules containing phospazo linkage: A computational study on the reaction between HSi (X2Π) and PN (X1Σ+)". Journal of Theoretical and Computational Chemistry. 16 (8): 1750075. doi:10.1142/s0219633617500754. ISSN   0219-6336.
  44. Millar, T. J.; Bennett, A.; Herbst, E. (1987-11-01). "An efficient gas phase synthesis for interstellar PN". Monthly Notices of the Royal Astronomical Society. 229 (1): 41P–44P. doi: 10.1093/mnras/229.1.41p . ISSN   0035-8711.
  45. 1 2 Chantzos, J.; Rivilla, V. M.; Vasyunin, A.; Redaelli, E.; Bizzocchi, L.; Fontani, F.; Caselli, P. (2020). "The first steps of interstellar phosphorus chemistry". Astronomy & Astrophysics. 633: A54. arXiv: 1910.13449 . Bibcode:2020A&A...633A..54C. doi: 10.1051/0004-6361/201936531 . hdl: 10995/90217 . ISSN   0004-6361.
  46. Indriolo, Nick; McCall, Benjamin J. (2012-01-03). "INVESTIGATING THE COSMIC-RAY IONIZATION RATE IN THE GALACTIC DIFFUSE INTERSTELLAR MEDIUM THROUGH OBSERVATIONS OF H+3". The Astrophysical Journal. 745 (1): 91. arXiv: 1111.6936 . Bibcode:2012ApJ...745...91I. doi: 10.1088/0004-637x/745/1/91 . ISSN   0004-637X.
  47. Moldenhauer, Wilhelm; Dörsam, H. (1926-05-05). "Über die Vereinigung von Phosphor und Stickstoff unter dem Einflusse elektrischer Entladungen". Berichte der Deutschen Chemischen Gesellschaft (A and B Series). 59 (5): 926–931. doi:10.1002/cber.19260590514. ISSN   0365-9488.
  48. Ahmad, I.K.; Hamilton, P.A. (1995). "The Fourier Transform Infrared Spectrum of PN". Journal of Molecular Spectroscopy. 169 (1): 286–291. Bibcode:1995JMoSp.169..286A. doi:10.1006/jmsp.1995.1022. ISSN   0022-2852.
  49. Liang, Shuyu; Hemberger, Patrick; Levalois-Grützmacher, Joëlle; Grützmacher, Hansjörg; Gaan, Sabyasachi (2017-04-05). "Probing Phosphorus Nitride (P≡N) and Other Elusive Species Formed upon Pyrolysis of Dimethyl Phosphoramidate". Chemistry – A European Journal. 23 (23): 5595–5601. doi:10.1002/chem.201700402. ISSN   0947-6539. PMID   28378378.
  50. 1 2 3 Eckhardt, André K.; Riu, Martin-Louis Y.; Müller, Peter; Cummins, Christopher C. (2022-01-12). "Staudinger Reactivity and Click Chemistry of Anthracene (A)-Based Azidophosphine N3PA". Inorganic Chemistry. 61 (3): 1270–1274. doi:10.1021/acs.inorgchem.1c03753. ISSN   0020-1669. PMID   35020379. S2CID   245914394.
  51. 1 2 Huffman, E. O.; Tarbutton, Grady; Elmore, Kelly L.; Cate, W. E.; Walters, H. K.; Elmore, G. V. (1954). "Synthesis of Phosphorus Nitrides1". Journal of the American Chemical Society. 76 (24): 6239–6243. doi:10.1021/ja01653a006. ISSN   0002-7863.
  52. Skaggs, S. R., Kaizerman, J., & Tapscott, R. E. (1995). Phosphorus Nitrides As Fire Extinguishing Agents. In Proceedings of Halon Options Technical Working Conference, Albuquerque, NM, USA (pp. 345-355).
  53. Hong, Ki Choong (1962). Synthesis of phosphorus nitride related high analysis fertilizers (Thesis). Iowa State University. doi:10.31274/rtd-180815-795.