Argon compounds

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 CSIRO.

Argon ionises at 15.76 eV, which is higher than hydrogen, but lower than helium, neon or fluorine.[1] Molecules containing argon can be van der Waals molecules held together very weakly by London dispersion forces. Ionic molecules can be bound by charge induced dipole interactions. With gold atoms there can be some covalent interaction.[2] Experimental methods used to study argon compounds have included inert gas matrices, infrared spectroscopy to study stretching and bending movements, microwave spectroscopy and far infrared to study rotation, and also visible and ultraviolet spectroscopy to study different electronic configurations including excimers. Mass spectroscopy is used to study ions.[3] Computation methods have been used to theoretically compute molecule parameters, and predict new stable molecules. Computational ab initio methods used have included CCSD(T), MP2 (Møller–Plesset perturbation theory of the second order), CIS and CISD. For heavy atoms, effective core potentials are used to model the inner electrons, so that their contributions do not have to be individually computed. More powerful computers since the 1990s have made this kind of in silico study much more popular, being much less risky and simpler than an actual experiment.[3] This article is mostly based on experimental or observational results.

The argon fluoride laser is important in photolithography of silicon chips. These lasers make a strong ultraviolet emission at 192 nm.[4]

Argonium

Main article: Argonium

Argonium (ArH+) is an ion combining a proton and an argon atom. It is found in interstellar space in diffuse atomic hydrogen gas where the fraction of molecular hydrogen H2 is in the range of 0.0001 to 0.001.[1]

Argonium is formed when H2+ reacts with Ar atoms:[1]

Ar + H+
2
→ ArH+ + H[1]

and it is also produced from Ar+ ions produced by cosmic rays and X-rays from neutral argon:

Ar+ + H2 → *ArH+ + H[1] 1.49 eV.[5]

When ArH+ encounters an electron, dissociative recombination can occur, but it is extremely slow for lower energy electrons, allowing ArH+ to survive for a much longer time than many other similar protonated cations.

ArH+ + e → ArH* → Ar + H[1]

Artificial ArH+ made from earthly argon contains mostly the isotope 40Ar rather than the cosmically abundant 36Ar. Artificially it is made by an electric discharge through an argon-hydrogen mixture.[6]

Natural occurrence

In the Crab Nebula, ArH+ occurs in several spots revealed by emission lines. The strongest place is in the Southern Filament. This is also the place with the strongest concentration of Ar+ and Ar2+ ions.[5] The column density of ArH+ in the Crab Nebula is between 1012 and 1013 atoms per square centimeter.[5] Possibly the energy required to excite the ions so that then can emit, comes from collisions with electrons or hydrogen molecules.[5] Towards the Milky Way centre the column density of ArH+ is around 2×1013 cm−2.[1]

Cluster argon cations

The diargon cation, Ar+
2
has a binding energy of 1.29 eV.[7]

The triargon cation Ar+
3
is linear, but has one Ar−Ar bond shorter than the other. Bond lengths are 2.47 and 2.73 ångströms. The dissociation energy to Ar and Ar2+ is 0.2 eV. In line with the molecule's asymmetry, the charge is calculated as +0.10, +0.58 and +0.32 on each argon atom, so that it greatly resembles Ar+
2
bound to a neutral Ar atom.[8]

Larger charged argon clusters are also detectable in mass spectroscopy. The tetraargon cation is also linear. Ar+
13
icosahedral clusters have an Ar+
3
core, whereas Ar+
19
is dioctahedral with an Ar+
4
core. The linear Ar+
4
core has +0.1 charge on the outer atoms, and +0.4 charge on each or the inner atoms. For larger charged argon clusters, the charge is not distributed on more than four atoms. Instead the neutral outer atoms are attracted by induced electric polarization.[9] The charged argon clusters absorb radiation, from the near infrared, through visible to ultraviolet. The charge core, Ar+
2
, Ar+
3
or Ar+
4
is called a chromophore. Its spectrum is modified by the first shell of neutral atoms attached. Larger clusters have the same spectrum as the smaller ones. When photons are absorbed in the chromophore, it is initially electronically excited, but then energy is transferred to the whole cluster in the form of vibration. Excess energy is removed by outer atoms evaporating from the cluster one at a time. The process of destroying a cluster by light is called photofragmentation.[9]

Negatively-charged argon clusters are thermodynamically unstable, and therefore cannot exist. Argon has a negative electron affinity.[9]

Argon monohydride

Neutral argon hydride, also known as argon monohydride (ArH), was the first discovered noble gas hydride. J. W. C. Johns discovered an emission line of ArH at 767 nm and announced the find in 1970. The molecule was synthesized using X-ray irradiation of mixtures of argon with hydrogen-rich molecules such as H2, H2O, CH4 and CH3OH.[10] The X-ray excited argon atoms are in the 4p state.[11]

Argon monohydride is unstable in its ground state, 4s, as a neutral inert gas atom and a hydrogen atom repel each other at normal intermolecular distances. When a higher-energy-level ArH* emits a photon and reaches the ground state, the atoms are too close to each other, and they repel and break up. However a van der Waals molecule can exist with a long bond.[12] However, excited ArH* can form stable Rydberg molecules, also known as excimers. These Rydberg molecules can be considered as a protonated argon core, surrounded by an electron in one of many possible higher energy states.[13]

Formation: Ar + ν → Ar*;  Ar* + H2 → ArH* + H[10]

Note that some molecules bind hydrogen too strongly for the reaction to proceed. For example, acetylene will not form ArH this way.[10]

In the van der Waals molecule of ArH, the bond length is calculated to be about 3.6 Å and the dissociation energy calculated to be 0.404 kJ/mol (33.8 cm−1).[14] The bond length in ArH* is calculated as 1.302 Å.[15]

The spectrum of argon monohydride, both ArH* and ArD*, has been studied. The lowest bound state is termed A2Σ+ or 5s. Another low lying state is known as 4p, made up of C2Σ+ and B2π states. Each transition to or from higher level states corresponds to a band. Known bands are 3p  5s, 4p  5s, 5p  5s (band origin 17486.527 cm−1[16]), 6p  5s (band origin 21676.90 cm−1[16]) 3dσ  4p, 3dπ  4p (6900 cm−1), 3dδ  4p (8200–8800 cm−1), 4dσ  4p (15075 cm−1), 6s  4p (7400–7950 cm−1), 7s  4p (predicted at 13970 cm−1, but obscured), 8s  4p (16750 cm−1), 5dπ  4p (16460 cm−1), 5p  6s (band origin 3681.171 cm−1),[17] 4f  5s (20682.17 and 20640.90 cm−1 band origin for ArD and ArH), 4f  3dπ (7548.76 and 7626.58 ccm−1), 4f  3dδ (6038.47 and 6026.57 cm−1), 4f  3dσ (4351.44 cm−1 for ArD).[12] The transitions going to 5s, 3dπ  5s and 5dπ  5s, are strongly predissociated, blurring out the lines.[17] In the UV spectrum a continuous band exists from 200 to 400 nm. This band is due to two different higher states: B2Π  A2Σ+ radiates over 210–450 nm, and E2Π  A2Σ+ is between 180 and 320 nm.[18] A band in the near infrared from 760 to 780 nm.[19]

Other ways to make ArH include a Penning-type discharge tube, or other electric discharges. Yet another way is to create a beam of ArH+ (argonium) ions and then neutralize them in laser-energized caesium vapour. By using a beam, the lifetimes of the different energy states can be observed, by measuring the profile of electromagnetic energy emitted at different wavelengths.[20] The E2π state of ArH has a radiative lifetime of 40 ns. For ArD the lifetime is 61 ns. The B2Π state has a lifetime of 16.6 ns in ArH and 17 ns in ArD.[18]

Argon polyhydrides

The argon dihydrogen cation ArH+
2
has been predicted to exist and to be detectable in the interstellar medium. However it has not been detected as of 2016.[21] ArH+
2
is predicted to be linear in the form Ar−H−H. The H−H distance is 0.94 Å. The dissociation barrier is only 2 kcal/mol (8 kJ/mol), and ArH+
2
readily loses a hydrogen atom to yield ArH+.[22] The force constant of the ArH bond in this is 1.895 mdyne2 (1.895×1012 Pa).[23]

The argon trihydrogen cation ArH+
3
has been observed in the laboratory.[21][24] ArH2D+, ArHD+
2
and ArD+
3
have also been observed.[25] The argon trihydrogen cation is planar in shape, with an argon atom off the vertex of a triangle of hydrogen atoms.[26]

Argoxonium

The argoxonium ion ArOH+ is predicted to be bent molecular geometry in the 11A′ state. 3Σ is a triplet state 0.12 eV higher in energy, and 3A″ is a triplet state 0.18 eV higher. The Ar−O bond is predicted to be 1.684 Å long[21] and to have a force constant of 2.988 mdyne/Å2 (2.988×1012 Pa).[23]

ArNH+

ArNH+ is a possible ionic molecule to detect in the lab, and in space, as the atoms that compose it are common. ArNH+ is predicted to be more weakly bound than ArOH+, with a force constant in the Ar−N bond of 1.866 mdyne/Å2 (1.866×1012 Pa). The angle at the nitrogen atom is predicted to be 97.116°. The Ar−N lengths should be 1.836 Å and the N−H bond length would be 1.046 Å[23][27]

Argon dinitrogen cation

The argon dinitrogen linear cationic complex has also been detected in the lab:

Ar + N+
2
ArN+
2
photodissociation Ar+ + N2.[21]

The dissociation yields Ar+, as this is a higher-energy state.[7] The binding energy is 1.19 eV.[7] The molecule is linear. The distance between two nitrogen atoms is 1.1 Å. This distance is similar to that of neutral N2 rather than that of N+
2
ion. The distance between one nitrogen and the argon atom is 2.2 Å.[7] The vibrational band origin for the nitrogen bond in ArN+
2
(V = 0  1) is at 2272.2564 cm−1 compared with N2+ at 2175 and N2 at 2330 cm−1.[7]

In the process of photon decay, the dissociation of the molecule when hit by photons, it is three times more likely to yield Ar+ + N2 compared to Ar + N+
2
.[28]

ArHN+
2

ArHN+
2
has been produced in a supersonic jet expansion of gas and detected by Fourier transform microwave spectroscopy.[24] The molecule is linear, with the atoms in the order Ar−H−N−N. The Ar−H distance is 1.864 Å. There is a stronger bond between hydrogen and argon than in ArHCO+.[29]

The molecule is made by the following reaction:

ArH+ + N2ArHN+
2
.[29]

Bis(dinitrogen) argon cation

The argon ion can bond two molecules of dinitrogen (N2) to yield an ionic complex with a linear shape and structure N=N−+Ar−N=N. The N=N bond length is 1.1014 Å, and the nitrogen to argon bond length is 2.3602 Å. 1.7 eV of energy is required to break this apart to N2 and ArN+
2
. The band origin of an infrared band due to antisymmetric vibration of the N=N bonds is at 2288.7272 cm−1. Compared to N2 it is redshifted 41.99 cm−1. The ground state rotational constant of the molecule is 0.034296 cm−1.[28]

Ar(N
2
)+
2
is produced by a supersonic expansion of a 10:1 mixture of argon with nitrogen through a nozzle, which is impacted by an electron beam.[28]

ArN2O+

ArN2O+ absorbs photons in four violet–ultraviolet wavelength bands leading to breakup of the molecule. The bands are 445–420, 415–390, 390–370, and 342 nm.[30][31]

ArHCO+

ArHCO+ has been produced in a supersonic-jet expansion of gas and detected by Fabry–Perot-type Fourier transform microwave spectroscopy.[24][32]

The molecule is made by this reaction

ArH+ + CO → ArHCO+.[29]

Carbon dioxide–argon ion

ArCO+
2
can be excited to form ArCO+
2
* where the positive charge is moved from the carbon dioxide part to the argon. This molecule may occur in the upper atmosphere. Experimentally the molecule is made from a low-pressure argon gas with 0.1% carbon dioxide, irradiated by a 150 V electron beam. Argon is ionized, and can transfer the charge to a carbon dioxide molecule.[33] The dissociation energy of ArCO+
2
is 0.26 eV.[33]

ArCO+
2
+ CO2 → Ar + CO
2
·CO+
2
(yields 0.435 eV.)[33]

van der Waals molecules

Neutral argon atoms bind very weakly to other neutral atoms or molecules to form van der Waals molecules. These can be made by expanding argon under high pressure mixed with the atoms of another element. These atoms can be produced by evaporation with a laser. Alternatively by an electric discharge. The known molecules include AgAr, Ag2Ar, NaAr, KAr, MgAr, CaAr, SrAr, ZnAr, CdAr, HgAr, SiAr,[34] InAr, CAr,[35] GeAr,[36] SnAr,[37] and BAr.[38] SiAr was made from silicon atoms derived from Si(CH3)4.[39]

In addition to the very weakly bound van der Waals molecules, electronically-excited molecules with the same formula exist. As a formula these can be written ArX*, with the "*" indicating an excited state. The atoms are much more strongly bound with a covalent bond. They can be modeled as an ArX+ surrounded by a higher energy shell with one electron. This outer electron can change energy by exchanging photons and so can fluoresce. The widely used argon fluoride laser makes use of the ArF* excimer to produce strong ultraviolet radiation at 192 nm. The Argon chloride laser using ArCl* produces even shorter ultraviolet at 175 nm, but is too feeble for application.[40] The argon chloride in this laser comes from argon and chlorine molecules.[41]

Argon clusters

Cooled argon gas can form clusters of atoms. Diargon has a binding energy of 0.012 eV, but the Ar13 and Ar19 clusters have a sublimation energy (per atom) of 0.06 eV. For liquid argon Ar the energy increases to 0.08 eV. Clusters of up to several hundred argon atoms have been detected. These argon clusters are icosahedral in shape, consisting of shells of atoms arranged around a central atom. The structure changes for clusters with more than 800 atoms to resemble a tiny crystal with a face centered cubic (fcc) structure, as in solid argon. It is the surface energy that maintains an icosahedral shape, but for larger clusters internal pressure will attract the atoms into an fcc arrangement.[9] Neutral argon clusters are transparent to visible light.[9]

Diatomic van der Waals molecules

Molecule Binding energy
ground Σ state
(cm−1)
Binding energy
excited Π state
(cm−1)
Ground state
bond length
(Å)
Excited state
bond length
(Å)
CAS number[42]
ArH 30736-04-0
ArHe 12254-69-2
LiAr 42.5 925 4.89 2.48[43]
BAr 149358-32-7
ArNe 12301-65-4
NaAr 40 560 56633-38-6
MgAr 44 246 72052-59-6
AlAr 143752-09-4
SiAr[44]
ArCl 54635-29-9
Ar2 12595-59-4
KAr 42 373 12446-47-8
CaAr 62 134 72052-60-9
SrAr 68 136
NiAr 401838-48-0
ZnAr 96 706 72052-61-0
GaAr 149690-22-2
GeAr[45]
KrAr 51184-77-1
AgAr 90 1200
CdAr 106 544 72052-62-1
InAr[46] 146021-90-1
SnAr[47]
ArXe 58206-67-0
AuAr 195245-92-2
HgAr 131 446 87193-95-1

ArO* is also formed when dioxygen trapped in an argon matrix is subjected to vacuum ultraviolet. It can be detected by its luminescence:

O2 + hvO+
2
+ e;  O+
2
+ e → 2O*;  O* + Ar → ArO*.[48]

Light emitted by ArO* has two main bands, one at 2.215 eV, and a weaker one at 2.195 eV.[49]

Argon sulfide, ArS* luminesces in the near infrared at 1.62 eV. ArS is made from from UV irradiated OCS in an argon matrix. The excited states lasts for 7.4 and 3.5 μs for spectrum peak and band respectively.[50]

Triatomic van der Waals molecules

Cluster molecules containing dichlorine and more than one argon atom can be made by forcing a 95:5 mixture of helium and argon and a trace of chlorine though a nozzle. ArCl2 exists in a T shape. Ar2Cl2 has a distorted tetrahedron shape, with the two argon atoms 4.1 Å from each other, and their axis 3.9 Å from the Cl2. The van der Waals bond energy is 447 cm−1. Ar3Cl2 also exists with a van der Waals bond energy of 776 cm−1.[51]

The linear Ar·Br2 molecule has a continuous spectrum for bromine molecule X  B transitions. The spectrum of bromine is blue-shifted and spread out when it binds an argon atom.[52]

ArI2 shows a spectrum that adds satellite bands to the higher vibrational bands of I2.[53] The ArI2 molecule has two different isomers, one shape is linear, and the other is T shaped. The dynamics of ArI2 is complex. Breakup occurs through different routes in the two isomers. The T shape undergoes intramolecular vibrational relaxation, whereas the linear one directly breaks apart.[54] Diiodine clusters, I2Arn have been made.[55]

The ArClF cluster has a linear shape.[56] The argon atom is closest to the chlorine atom.[52]

Linear ArBrCl can also rearrange to ArClBr, or a T-shaped isomer.[57]

Multiple argon atoms can "solvate" a water molecule forming a monolayer around the H2O. Ar12·H2O is particularly stable, having an icosahedral shape. Molecules from Ar·H2O to Ar14·H2O have been studied.[58]

Polyatomic van der Waals molecules

Other polyatomic van der Waals compounds of argon, include those of fluorobenzene,[59] formyl radical (ArHCO),[60] 7-azaindole,[61] glyoxal,[62] sodium chloride (ArNaCl),[63] ArHCl,[64] and cyclopentanone.[65]

Molecule Name Ground state
binding energy
(cm−1)
Closest atom
to argon
Ground state
bond length
(Å)
Bond angle
from atom
(degrees)
CAS number references
(CH3)2F2Si·Ar Difluorodimethylsilane – argon
CH2F2·Ar Difluoromethane – argon F 3.485 58.6 [66]
CF3CN trifluoromethylcyanide argon C1 3.73 77 947504-98-5 [67]
CF2HCH3·Ar 1,1-difluoroethane argon F [68]
CH2FCH2F·Ar 1,2-difluoroethane argon 181 F 3.576 61 264131-14-8 [69]
CH3CHO·Ar Acetaldehyde argon 161 C-1 3.567 76.34 158885-13-3 [70]
C2H4O·Ar oxirane argon 200 O 3.606 (CM) 72.34 [71]

Aqueous argon

Argon dissolved in water causes the pH to rise to 8.0,[72] apparently by reducing the number of oxygen atoms available to bind protons.[73]

With ice, argon forms a clathrate hydrate. Up to 0.6 GPa, the clathrate has a cubic structure. Between 0.7 and 1.1 GPa the clathrate has a tetragonal structure. Between 1.1 and 6.0 GPa the structure is body centered orthorhombic. Over 6.1 GPa, the clathrate converts into solid argon and ice VII.[74] At atmospheric pressure the clathrate is stable below 147 K.[75] At 295 K the argon pressure from the clathrate is 108 MPa.[76]

Argon fluorohydride

Argon fluorohydride was an important discovery in the rejuvenation of the study of noble gas chemistry. HArF is stable in solid form at temperatures below 17 K.[77] It is prepared by photolosis of hydrogen fluoride in a solid argon matrix.[78] HArArF would have such a low barrier to decomposition that it will likely never be observed.[79] However HBeArF is predicted to be more stable than HArF.[80]

Actinide compounds

CUO in a solid argon matrix can bind one, or a few argon atoms to yield CUO·Ar, CUO·Ar3 or CUO·Ar4. CUO itself is made by evaporating uranium atoms into carbon monoxide. Uranium acts as a strong Lewis acid in CUO[78][81] and forms bonds with energies of about 3.2 kcal/mol (13.4 kJ/mol) with argon. The argon acts as a Lewis base. Its electron density is inserted into an empty 6d orbital on the uranium atom. The spectrum of CUO is changed by argon so that the U−O stretch frequency changes from 872.2 to 804.3 cm−1 and the U−C stretch frequency from 1047.3 to 852.5 cm−1.[82] The significant change in the spectrum occurs because the CUO is changed from a singlet state (in gas phase or solid neon) to a triplet state, with argon or noble gas complexing.[83] The argon–uranium bond length is 3.16 Å.[82] This is shorter than the sum of atomic radii of U and Ar of 3.25 Å, but considerably longer than a normal covalent bond to uranium. For example, U−Cl in UCl6 is 2.49 Å.[83] When xenon is included in the solid argon matrix up to a few percent, additional van der Waals molecules are formed: CUO·Ar3Xe, CUO·Ar2Xe2, CUO·ArXe3 and CUO·Xe4.[81] Similarly krypton can substitute for argon in CUO·Ar3Kr, CUO·Ar2Kr2, CUO·ArKr3 and CUO·Kr4.[83] The shape of these molecules is roughly octahedral, with a uranium centre and with the noble gas atoms around the equator.[83]

UO+
2
can bind up to five noble gas atoms in a ring around a linear O=+U=O core.[84] These molecules are produced when uranium metal is laser ablated into dioxygen. This produces UO, UO2, UO3, U+, and importantly UO+
2
. UO+
2
is then condensed into a noble gas matrix, either a pure element or a mixture. Heavier noble gas atoms will tend to displace the lighter atoms. Ionic molecules produced this way include UO
2
Ne
4
Ar+
, UO
2
Ne
3
Ar+
2
, UO
2
Ne
2
Ar+
3
, UO
2
NeAr+
4
, UO
2
Ar+
5
, UO
2
Ar
4
Kr+
, UO
2
Ar
3
Kr+
2
, UO
2
Ar
2
Kr+
3
, UO
2
ArKr+
4
, UO
2
Ar
4
Xe+
, UO
2
Ar
3
Xe+
2
, UO
2
Ar
2
Xe+
3
, and UO
2
ArXe+
4
, which are identified by a shift in the U=O antisymmetric stretching frequency.[84]

Neutral UO2 condensed in solid argon is converted from one electronic state to another by the argon atom ligands. In argon the electron configuration is 5f2(δφ) whereas in neon it is 5f17s1 (the state 3H4g compared to 3Φ2u). This is because the argon atoms have a larger antibonding interaction with the 7s1 electron, forcing it into a different subshell. The argonated compound has a stretching frequency of 776 cm−1 compared to 914.8 cm−1 in neon.[85] The argon uranium dioxide molecule is likely UO2Ar5.[86]

Beryllium oxide

When beryllium atoms react with oxygen in a solid argon matrix (or beryllia is evaporated into the matrix) ArBeO will be formed, and is observable by its infrared spectrum. The beryllia molecule is strongly polarised, and the argon atom is attracted to the beryllium atom.[83][87] The bond strength of Ar−Be is calculated to be 6.7 kcal/mol (28 kJ/mol). The Ar−Be bond length is predicted to be 2.042 Å.[88]

Analogously, beryllium reacting with hydrogen sulfide and trapped in an argon matrix at 4 K forms ArBeS. It has a binding energy calculated to be 12.8 kcal/mol (54 kJ/mol).[89]

ArBeO2CO (beryllium carbonate) has been prepared (along with Ne, Kr and Xe adducts).[90]

Carbonyl compounds

Group 6 elements can form reactive pentacarbonyls that can react with argon. These were actually argon compounds discovered in 1975, and were known before the discovery of HArF, but are usually overlooked.[91] Tungsten normally forms a hexacarbonyl, but when subject to ultraviolet radiation it breaks into a reactive pentacarbonyl. When this is condensed into a noble gas matrix the infrared and UV spectrum varies considerably depending on the noble gas used. This is because the noble gas present binds to the vacant position on the tungsten atom. Similar results also occur with molybdenum and chromium.[92] Argon is only very weakly bound to tungsten in ArW(CO)5.[83][93] The Ar−W bondlength is predicted to be 2.852 Å.[92] The same substance is produced for a brief time in supercritical argon at 21 °C.[94] For ArCr(CO)5 the band maximum is at 533 nm (compared to 624 nm in neon, and 518 nm in krypton). Forming 18-electron complexes, the shift in spectrum due to different matrices was much smaller, only around 5 nm. This clearly indicates the formation of a molecule using atoms from the matrix.[3]

Other carbonyls and compexed carbonyls also have reports of bonding to argon. These include Ru(CO)2(PMe3)2Ar, Ru(CO)2(dmpe)2Ar, η6-C6H6Cr(CO)2Ar.[95] Evidence also exists for ArHMn(CO)4, ArCH3Mn(CO)4, and fac-(η2-dfepe)Cr(CO)3Ar.[3]

Other noble gas complexes have been studied by photolysis of carbonyls dissolved in liquid rare gas, possibly under pressure. These Kr or Xe complexes decay on the time scale of seconds, but argon does not seem to have been studied this way. The advantage of liquid noble gases is that the medium is completely transparent to infrared radiation, which is needed to study the bond vibration in the solute.[3]

Attempts have been made to study cabonyl–argon adducts in the gas phase, but the interaction appears to be too weak to observe a spectrum. In the gas form, the absorption lines are broadened into bands because of rotation that happens freely in a gas.[3] The argon adducts in liquids or gases are unstable as the molecules easily react with the other photolysis products, or dimerize, eliminating argon.[3]

Coinage metal monohalides

The argon coinage metal monohalides were the first noble gas metal halides discovered, when the metal monohalide molecules were put through an argon jet. There were first found in Vancouver in 2000.[96] ArMX with M = Cu, Ag or Au and X = F, Cl or Br have been prepared. The molecules are linear. In ArAuCl the Ar−Au bond is 2.47 Å, the stretching frequency is 198 cm−1 and the dissociation energy is 47 kJ/mol.[97] ArAgBr also has been made.[97] ArAgF has a dissociation energy of 21 kJ/mol.[97] The Ar−Ag bond-length in these molecules is 2.6 Å.[97] ArAgCl is isoelectronic with AgCl
2
which is better known.[97] The Ar−Cu bond length in these molecules is 2.25 Å.[97]

Transition metal oxides

In a solid argon matrix VO2 forms VO2Ar2, and VO4 forms VO4·Ar with binding energy calculated to be 12.8 and 5.0 kcal/mol (53 and 21 kJ/mol).[98] Scandium in the form of ScO+ coordinates five argon atoms to yield ScOAr+
5
.[99] these argon atoms can be substituted by numbers of krypton or xenon atoms to yield even more mixed noble gas molecules. With yttrium, YO+ bonds six argon atoms, and these too can be substituted by varying numbers of krypton or xenon atoms.[100]

In the case of transition metal monoxides, ScO, TiO and VO do not form a molecule with one argon atom. However CrO, MnO, FeO, CoO and NiO can each coordinate one argon atom in a solid argon matrix.[101] The metal monoxide molecules can be produced by laser ablation of the metal trioxide, followed by condensation on solid argon. ArCrO absorbs at 846.3 cm−1, ArMnO at 833.1, ArFeO at 872.8, ArCoO at 846.2, Ar58NiO at 825.7 and Ar60NiO at 822.8 cm−1. All these molecules are linear.[101]

There are also claims of argon forming coordination molecules in NbO2Ar2, NbO4Ar, TaO4Ar,[102] VO2Ar2, VO4Ar,[98] Rh(η2-O2)Ar2, Rh(η2-O2)2Ar2, Rh(η2-O2)2(η1-OO)Ar.[103][104][105]

Other transition metal compounds

ArNiN2 binds argon with 11.52 kcal/mol. The bending frequency of ArN2 is changed from 310.7 to 358.7 cm−1 when argon attaches to the nickel atom.[106]

Other ions

Some other binary ions observed that contain argon include BaAr2+ and BaAr2+
2
,[107] VAr+, CrAr+, FeAr+, CoAr+, and NiAr+.[3]

Gold and silver cluster ions can bind argon. Known ions are Au
3
Ar+
, Au
3
Ar+
2
, Au
3
Ar+
3
, Au
2
AgAr+
3
and AuAg
2
Ar+
3
. These have a triangular shaped metallic core with argon bound at the vertexes.[2]

ArF+ is also known[3] to be formed in the reaction

F+
2
+ Ar → ArF+ + F

and also

Ar+ + F2 → ArF+ + F.

and also

SF2+
4
+ Ar → ArF+ + SF+
3
.[108]

The ions can be produced by ultraviolet light at 79.1 nm or less.[109] The ionisation energy of fluorine is higher than that of argon, so breakup occurs thus:

ArF+ → Ar+ + F.[110]

The millimeter wave spectrum of ArF+ between 119.0232 and 505.3155 GHz has been measured to calculate molecular constants B0 = 14.8788204 GHz, D0 = 28.718 kHz.[111] There is a possibility that a solid salt of ArF+ could be prepared with SbF
6
or AuF
6
anions.[110][112]

Excited or ionized argon atoms can react with molecular iodine gas to yield ArI+[113] Argon plasma is used as an ionisation source and carrier gas in inductively coupled plasma mass spectrometry. This plasma reacts with samples to produce monatomic ions, but also forms argon oxide (ArO+), and argon nitride (ArN+) cations, which can cause isobaric interference with detection and measurement of iron-56 (56Fe) and iron-54 (54Fe), respectively, in mass spectrometry.[114] Argon chloride cations can interfere with the detection of arsenic as Ar35Cl+ has a mass-to-charge ratio almost identical to that of arsenic's one stable isotope, 75As.[115] In these circumstances ArO+ may be removed by reaction with NH3.[116] Alternatively electrothermal vaporization or using helium gas can avoid these interference problems.[114] Argon can also form an anion with chlorine, ArCl,[117] though this is not a problem for mass spectrometry applications as only cations are detected.

The argon borynium ion, BAr+ is produced when BBr+ at energies between 9 and 11 eV reacts with argon atoms. 90% of the positive charge is on the argon atom.[118]

ArC+ ions can be formed when argon ions impact carbon monoxide with energies between 21 and 60 eV. However more C+ ions are formed, and when the energy is on the high side, O+ is higher.[119]

ArN+ can form when argon ions impact dinitrogen with energies between 8.2 and 41.2 eV and peaking around 35 eV. However far more N+
2
and N+ are produced.[120]

ArXe+ is held together with a strength of 1445 cm−1 when it is in the X electronic state, but 1013 cm−1 when it is in the B excited state.[31]

Metal–argon cations are called "argides". The argide ions produced during mass spectroscopy have higher intensity when the binding energy of the ion is higher. Transition elements have higher binding and ion flux intensity compared to main group elements. Argides can be formed in the plasma by excited argon atoms reacting with another element atom, or by an argon atom binding with another ion:

Ar+ + M → ArM+ + e; M+ + Ar → ArM+.[121]

Doubly charged cations, called superelectrophiles, are capable of reacting with argon. Ions produced include ArCF2+
2
ArCH+
2
, ArBF+
2
and ArBF2+
containing bonds between argon and carbon or boron.[122]

Doubly ionised acetylene HCCH2+ reacts inefficiently with argon to yield HCCAr2+. This product competes with the formation of Ar+ and argonium.[123]

The SiF2+
3
ion reacts with argon to yield ArSiF2+
2
.[124]

Ion Bond length
(Å)
Dissociation energy
(kJ/mol)[3]
ArH+ 3.4 eV
LiAr+[121]
BeAr+[121]
BAr+[118] 2.590 210
ArC+[125]
ArN+[121] 3.5 2.16 eV[126]
ArO+[121]
ArF+[110] 1.637 194
NaAr+[121] 19.3
MgAr+[121]
AlAr+[121]
SiAr+[121]
ArP+[121]
ArS+[121]
ArCl+[121]
Ar+
2
[121]
ScAr+[121]
VAr+ 37
CrAr+ 28
FeAr+
CoAr+ 49
MnAr+[121]
NiAr+ 53
CuAr+[121]
GaAr+[121]
AsAr+[121]
RbAr+[127]
NbAr+ 37
AgAr+[121]
ArI+[113]

Solid compounds

Solid buckminsterfullerene has small spaces between the C60 balls. Under 200 MPa pressure and 200 °C heat for 12 hours, argon can be intercalated into the solid to form crystalline Ar1C60. Once this cools down it is stable at standard conditions for months. Argon atoms occupy octahedral interstitial sites. The crystalline lattice size is almost unchanged at room temperature, but is slightly larger than pure C60 below 265 K. However argon does stop the buckyballs spinning below 250 K, a lower temperature than in pure C60.[128]

Solid C70 fullerene will also absorb argon under pressure of 200 MPa and at a temperature of 200 °C. C70·Ar has argon in octahdedral sites and has the rock salt structure, with cubic crystals in which the lattice parameter is 15.001 Å. This compares to the pure C70 lattice parameter of 16.964 Å, so the argon forces the crystals to expand slightly. The C70 ellipsoidal balls rotate freely in the solid, they are not locked into position by extra argon atoms filling the holes. Argon gradually escapes over a couple of days when the solid is stored at standard conditions, so that C70·Ar is less stable than C60·Ar. This is likely to be due to the shape and internal rotation allowing channels through which Ar atoms can move.[129]

When fullerenes are dissolved and crystallized from toluene, solids may form with toluene included as part of the crystal. However, if this crystallization is performed under a high pressure argon atmosphere, toluene is not included, being replaced by argon. The argon is then removed from the resultant crystal by heating.[130]

The argon clathrate is described in the Aqueous argon section.

Argon difluoride, ArF2, is predicted to be stable at pressures over 57 GPa. It should be an electrical insulator.[131]

At around 4 K there are two phases where neon and argon are mixed as a solid: Ne2Ar and Ar2Ne.[132] With Kr, solid argon forms a disorganized mixture.[133]

Under high pressure stoichiometric solids are formed with hydrogen and oxygen: Ar(H2)2 and Ar(O2)3.[134]

Ar(H2)2 crystallises in the hexagonal C14 MgZn2 Laves phase. It is stable to at least 200 GPa, but is predicted to change at 250 GPa to an AlB2 structure. At even higher pressures the hydrogen molecules should break up followed by metallization.[134]

Oxygen and argon under pressure at room temperature form several different alloys with different crystal structures. Argon atoms and oxygen molecules are similar in size, so that a greater range of miscibility occurs compared to other gas mixtures. Solid argon can dissolve up to 5% oxygen without changing structure. Below 50% oxygen a hexagonal close packed phase exists. This is stable from about 3GPa to 8.5 GPa. Typical formula is ArO. With more oxygen between 5.5 and 7 GPa, a cubic Pm3n structure exists, but under higher pressure it changes to a I-42d space group form. With more than 8.5 GPa these alloys separate to solid argon and ε-oxygen. The cubic structure has a unit cell edge of 5.7828 Å at 6.9 GPa. The representative formula is Ar(O2)3.[135]

Using density-functional theory ArHe2 is predicted to exist with the MgCu2 Laves phase structure at high pressures below 13.8 GPa. Above 13.8 GPa it transforms to AlB2 structure.[136]

References

  1. 1 2 3 4 5 6 7 Schilke, P.; Neufeld, D. A.; Müller, H. S. P.; Comito, C.; Bergin, E. A.; Lis, D. C.; Gerin, M.; Black, J. H.; Wolfire, M.; Indriolo, N.; Pearson, J. C.; Menten, K. M.; Winkel, B.; Sánchez-Monge, Á.; Möller, T.; Godard, B.; Falgarone, E. (4 June 2014). "Ubiquitous argonium (ArH+) in the diffuse interstellar medium: A molecular tracer of almost purely atomic gas". Astronomy & Astrophysics. 566: A29. Bibcode:2014A&A...566A..29S. doi:10.1051/0004-6361/201423727.
  2. 1 2 Shayeghi, Armin; Johnston, Roy L.; Rayner, David M.; Schäfer, Rolf; Fielicke, André (1 September 2015). "The Nature of Bonding between Argon and Mixed Gold–Silver Trimers". Angewandte Chemie International Edition. 54 (36): 10675–10680. doi:10.1002/anie.201503845.
  3. 1 2 3 4 5 6 7 8 9 10 Grills, D. C.; George, M. W. (2001). "Transition metal–noble gas complexes". Advances in Inorganic Chemistry. Advances in Inorganic Chemistry. 52: 113–150. doi:10.1016/S0898-8838(05)52002-6. ISBN 9780120236527.
  4. La Fontaine, B. (October 2010). "Lasers and Moore's Law". SPIE Professional: 20.
  5. 1 2 3 4 Barlow, M. J.; Swinyard, B. M.; Owen, P. J.; Cernicharo, J.; Gomez, H. L.; Ivison, R. J.; Krause, O.; Lim, T. L.; Matsuura, M.; Miller, S.; Olofsson, G.; Polehampton, E. T. (12 December 2013). "Detection of a Noble Gas Molecular Ion, 36ArH+, in the Crab Nebula". Science. 342 (6164): 1343–1345. Bibcode:2013Sci...342.1343B. doi:10.1126/science.1243582. PMID 24337290.
  6. Brown, John M.; Jennings, D. A.; Vanek, M.; Zink, L. R.; Evenson, K. M. (April 1988). "The pure rotational spectrum of ArH+". Journal of Molecular Spectroscopy. 128 (2): 587–589. Bibcode:1988JMoSp.128..587B. doi:10.1016/0022-2852(88)90173-7.
  7. 1 2 3 4 5 Linnartz, H. (16 August 2002). "Rotationally Resolved Infrared Spectrum of the Charge Transfer Complex [Ar–N2]+". Science. 297 (5584): 1166–1167. Bibcode:2002Sci...297.1166L. doi:10.1126/science.1074201. PMID 12183626.
  8. Bowers, Michael T.; Palke, William E.; Robins, Kathleen; Roehl, Coleen; Walsh, Sherrie (May 1991). "On the structure and photodissociation dynamics of Ar+
    3
    ". Chemical Physics Letters. 180 (3): 235–240. Bibcode:1991CPL...180..235B. doi:10.1016/0009-2614(91)87146-3.
  9. 1 2 3 4 5 Johnston, Roy L. (2002). Atomic and Molecular Clusters. London: Taylor & Francis. pp. 50–62. ISBN 0-7484-0930-0.
  10. 1 2 3 Berman, Michael; Kaldor, Uzi; Shmulovich, J.; Yatsiv, S. (December 1981). "Rydberg states and the observed spectrum of ArH". Chemical Physics. 63 (1–2): 165–173. Bibcode:1981CP.....63..165B. doi:10.1016/0301-0104(81)80318-7.
  11. Shmulovich, J.; Yatsiv, S. (15 October 1980). "Excitation Kinetics of ArH* Luminescence in X-Ray Irradiated Argon-Hydrogen Mixtures". Chemical Physics Letters. 75 (2): 319–323.
  12. 1 2 Dąbrowski, I.; Tokaryk, D. W.; Lipson, R. H.; Watson, J. K. G. (May 1998). "New Rydberg–Rydberg Transitions of the ArH and ArD Molecules. III. emission from the 4f complexes". Journal of Molecular Spectroscopy. 189 (1): 110–123. Bibcode:1998JMoSp.189..110D. doi:10.1006/jmsp.1997.7524. PMID 9571129.
  13. Jungen, C.; Roche, A. L.; Arif, M. (15 August 1997). "The Rydberg spectrum of ArH and KrH: calculation by R-matrix and generalized quantum defect theory". Philosophical Transactions of the Royal Society A: Mathematical, Physical and Engineering Sciences. 355 (1729): 1481–1506. Bibcode:1997RSPTA.355.1481J. doi:10.1098/rsta.1997.0072.
  14. Partridge, Harry; Bauschlicher, Charles W. (February 1999). "The dissociation energies of He2, HeH, and ArH: A bond function study". Molecular Physics. 96 (4): 705–710. Bibcode:1999MolPh..96..705P. doi:10.1080/00268979909483006.
  15. Parker, Eric P.; Ortiz, J.V. (November 1989). "Electron propagator calculations on the discrete spectra of ArH and NeH". Chemical Physics Letters. 163 (4–5): 366–370. Bibcode:1989CPL...163..366P. doi:10.1016/0009-2614(89)85151-6.
  16. 1 2 Dąbrowski, I.; Tokaryk, D. W.; Vervloet, M.; Watson, J. K. G. (1996). "New Rydberg–Rydberg transitions of the ArH and ArD molecules. I. Emission from np states of ArD". The Journal of Chemical Physics. 104 (21): 8245. Bibcode:1996JChPh.104.8245D. doi:10.1063/1.471578.
  17. 1 2 Dąbrowski, I.; Tokaryk, D. W.; Watson, J. K. G. (26 November 1997). "New Rydberg–Rydberg Transitions of the ArH and ArD Molecules. II. Emission from nd and ns States to the 4p State". Journal of Molecular Spectroscopy. 189 (1): 95–109. Bibcode:1998JMoSp.189...95D. doi:10.1006/jmsp.1997.7523. PMID 9571128.
  18. 1 2 Wunderlich, C.; Betz, V.; Bruckmeier, R.; Figger, H. (1993). "Lifetime measurements on ArH and ArD". The Journal of Chemical Physics. 98 (12): 9362. Bibcode:1993JChPh..98.9362W. doi:10.1063/1.464416.
  19. Van Hemert, Marc C.; Dohmann, Helmut; Peyerimhoff, Sigrid D. (December 1986). "Theoretical study of radiative and predissociative processes in ArH and ArD". Chemical Physics. 110 (1): 55–66. Bibcode:1986CP....110...55V. doi:10.1016/0301-0104(86)85144-8.
  20. Wunderlich, C.; Witzel, B.; Bruckmeier, R.; Figger, H. (November 1994). "UV spectra of ArH and ArD as emitted by a neutralized ion beam". Canadian Journal of Physics. 72 (11–12): 1236–1240. Bibcode:1994CaJPh..72.1236W. doi:10.1139/p94-159.
  21. 1 2 3 4 Theis, Riley A.; Fortenberry, Ryan C. (March 2016). "Potential interstellar noble gas molecules: ArOH+ and NeOH+ rovibrational analysis from quantum chemical quartic force fields". Molecular Astrophysics. 2: 18–24. Bibcode:2016MolAs...2...18T. doi:10.1016/j.molap.2015.12.001.
  22. Kuntz, P. J.; Roach, A. C. (1972). "Ion-Molecule Reactions of the Rare Gases with Hydrogen. Part 1. Diatomics in Molecules. Potential Energy Surface for ArH2+". J. Chem. Soc., Faraday Trans. 2. 68: 259. doi:10.1039/F29726800259.
  23. 1 2 3 Fortenberry, Ryan C. (June 2016). "Quantum astrochemical spectroscopy". International Journal of Quantum Chemistry. doi:10.1002/qua.25180.
  24. 1 2 3 Bailleux, S.; Bogey, M.; Bolvin, H.; Civiš, S.; Cordonnier, M.; Krupnov, A.F.; Tretyakov, M. Y.; Walters, A.; Coudert, L. H. (July 1998). "Sub-Millimeter-Wave Spectroscopy of the Ar·H+
    3
    and Ar·D+
    3
    Ionic Complexes". Journal of Molecular Spectroscopy. 190 (1): 130–139. Bibcode:1998JMoSp.190..130B. doi:10.1006/jmsp.1998.7564. PMID 9645933.
  25. Burenin, A. V. (6 May 2010). "Description of the torsional motion in the isotopically asymmetric ionic complexes ArH2D+ and ArD2H+". Optics and Spectroscopy. 108 (4): 506–511. Bibcode:2010OptSp.108..506B. doi:10.1134/S0030400X1004003X.
  26. Burenin, A. V. (4 September 2009). "Description of torsional motion in ionic complexes ArH+
    3
    and ArD+
    3
    ". Optics and Spectroscopy. 107 (2): 228–234. Bibcode:2009OptSp.107..228B. doi:10.1134/S0030400X09080116.
  27. Novak, Carlie M.; Fortenberry, Ryan C. (April 2016). "Theoretical rovibrational analysis of the covalent noble gas compound ArNH+". Journal of Molecular Spectroscopy. 322: 29–32. Bibcode:2016JMoSp.322...29N. doi:10.1016/j.jms.2016.03.003.
  28. 1 2 3 Linnartz, H.; Verdes, D.; Knowles, P. J.; Lakin, N. M.; Rosmus, P.; Maier, J. P. (2000). "Linear and centrosymmetric N[sub 2]⋯Ar[sup +]⋯N[sub 2]". The Journal of Chemical Physics. 113 (3): 895. Bibcode:2000JChPh.113..895L. doi:10.1063/1.481868.
  29. 1 2 3 Seki, Kouji; Sumiyoshi, Yoshihiro; Endo, Yasuki (2002). "Pure rotational spectra of the Ar–HN+
    2
    and the Kr–HN+
    2
    ionic complexes". The Journal of Chemical Physics. 117 (21): 9750. Bibcode:2002JChPh.117.9750S. doi:10.1063/1.1518025.
  30. Nizkorodov, S. A.; Maier, J. P.; Bieske, E. J. (1995). "The infrared spectrum of the N2H+–He ion-neutral complex" (PDF). The Journal of Chemical Physics. 102 (13): 5570. Bibcode:1995JChPh.102.5570N. doi:10.1063/1.469286.
  31. 1 2 Bieske, E. J.; Soliva, A. M.; Friedmann, A.; Maier, J. P. (1992). "Photoinitiated charge transfer in N2O+–Ar". The Journal of Chemical Physics. 96 (10): 7535. Bibcode:1992JChPh..96.7535B. doi:10.1063/1.462405.
  32. Ohshima, Yasuhiro; Sumiyoshi, Yoshihiro; Endo, Yasuki (1997). "Rotational spectrum of the Ar–HCO+ ionic complex". The Journal of Chemical Physics. 106 (7): 2977. Bibcode:1997JChPh.106.2977O. doi:10.1063/1.473416.
  33. 1 2 3 Illies, Andreas J.; Jarrold, M. F.; Wagner-Redeker, W.; Bowers, Michael T. (May 1985). "Photoinduced intramolecular charge transfer: photodissociation of carbon dioxide ion (1+)-argon (CO2+·Ar) cluster ions". Journal of the American Chemical Society. 107 (10): 2842–2849. doi:10.1021/ja00296a003.
  34. Jouvet, C.; Lardeux-Dedonder, C.; Martrenchard, S.; Solgadi, D. (1991). "Fluorescence excitation spectrum of silver–argon van der Waals complex". The Journal of Chemical Physics. 94 (3): 1759. Bibcode:1991JChPh..94.1759J. doi:10.1063/1.459949.
  35. Lei, Jie; Dagdigian, Paul J. (2000). "Laser fluorescence excitation spectroscopy of the CAr van der Waals complex". The Journal of Chemical Physics. 113 (2): 602. Bibcode:2000JChPh.113..602L. doi:10.1063/1.481835.
  36. Tao, Chong; Dagdigian, Paul J. (2003). "Laser fluorescence excitation spectroscopy of the GeAr van der Waals complex". The Journal of Chemical Physics. 118 (3): 1242. Bibcode:2003JChPh.118.1242T. doi:10.1063/1.1529662.
  37. Tao, Chong; Dagdigian, Paul J. (2004). "Spectroscopic investigation of nonbonding interactions of group-14 atoms with rare gases: The SnAr van der Waals complex". The Journal of Chemical Physics. 120 (16): 7512. Bibcode:2004JChPh.120.7512T. doi:10.1063/1.1665957. PMID 15267664.
  38. Yang, Xin; Dagdigian, Paul J. (1997). "Fluorescence excitation and depletion spectroscopy of the BAr complex: Electronic states correlating with the excited valence B(2s2p22D) asymptote". The Journal of Chemical Physics. 106 (16): 6596. Bibcode:1997JChPh.106.6596Y. doi:10.1063/1.473649.
  39. Dedonder-Lardeux, C.; Jouvet, C.; Richard-Viard, M.; Solgadi, D. (1990). "Fluorescence excitation spectrum of the Si–Ar van der Waals complex". The Journal of Chemical Physics. 92 (5): 2828. Bibcode:1990JChPh..92.2828D. doi:10.1063/1.457929.
  40. Setser, edited by D. W., ed. (1979). Reactive Intermediates in the Gas Phase: Generation and Monitoring. New York: Academic Press. ISBN 0-12-637450-3.
  41. Shuaibov, A. K.; Shimon, L. L.; Shevera, I. V.; Volovich, P. N. (5 February 2011). "A high-frequency UV-VUV radiation source operating on a mixture of argon with chlorine molecules". Technical Physics Letters. 34 (1): 14–16. doi:10.1134/S1063785008010057. ISSN 1063-7850.
  42. Hüttner, W. "Molecular Constants". doi:10.1007/978-3-540-69954-5.
  43. Brühl, Rüdiger; Zimmermann, Dieter (February 1995). "High-resolution laser spectroscopy of the A ← X transition of LiAr". Chemical Physics Letters. 233 (4): 455–460. Bibcode:1995CPL...233..455B. doi:10.1016/0009-2614(94)01476-C.
  44. Tao, Chong; Teslja, Alexey; Dagdigian, Paul J.; Atahan, Sule; Alexander, Millard H. (2002). "Laser spectroscopic study of the SiAr van der Waals complex". The Journal of Chemical Physics. 116 (21): 9239. doi:10.1063/1.1473814.
  45. Tao, Chong; Dagdigian, Paul J. (2003). "Laser fluorescence excitation spectroscopy of the GeAr van der Waals complex". The Journal of Chemical Physics. 118 (3): 1242. doi:10.1063/1.1529662.
  46. Fawzy, Wafaa M.; Le Roy, Robert J.; Simard, Benoit; Niki, Hideaki; Hackett, Peter A. (1993). "Determining repulsive potentials of InAr from oscillatory bound→continuum emission". The Journal of Chemical Physics. 98 (1): 140. doi:10.1063/1.464663.
  47. Tao, Chong; Dagdigian, Paul J. (2004). "Spectroscopic investigation of nonbonding interactions of group-14 atoms with rare gases: The SnAr van der Waals complex". The Journal of Chemical Physics. 120 (16): 7512. doi:10.1063/1.1665957.
  48. Laarmann, T.; Wabnitz, H.; von Haeften, K.; Möller, T. (2008). "Photochemical processes in doped argon-neon core-shell clusters: The effect of cage size on the dissociation of molecular oxygen". The Journal of Chemical Physics. 128 (1): 014502. Bibcode:2008JChPh.128a4502L. doi:10.1063/1.2815798. PMID 18190199.
  49. Taylor, R. V.; Walker, W. C. (1979). "Photoluminescence of ArO and KrO in doped rare-gas matrices". The Journal of Chemical Physics. 70 (1): 284. Bibcode:1979JChPh..70..284T. doi:10.1063/1.437161.
  50. Taylor, R. V.; Walker, W. C. (1979). "Photoluminescence of ArS, KrS, and XeS in rare-gas matrices". Applied Physics Letters. 35 (5): 359. Bibcode:1979ApPhL..35..359T. doi:10.1063/1.91149.
  51. Bieler, Craig R.; Evard, Dwight D.; Janda, Kenneth C. (September 1990). "Argon chloride (Ar2Cl2 and Ar3Cl2): structure, bond energy, and dissociation dynamics". The Journal of Physical Chemistry. 94 (19): 7452–7457. doi:10.1021/j100382a027.
  52. 1 2 Pio, Jordan M.; van der Veer, Wytze E.; Bieler, Craig R.; Janda, Kenneth C. (2008). "Product state resolved excitation spectroscopy of He–, Ne–, and Ar–Br2 linear isomers: Experiment and theory". The Journal of Chemical Physics. 128 (13): 134311. doi:10.1063/1.2885047.
  53. Kubiak, Glenn; Fitch, Pamela S. H.; Wharton, Lennard; Levy, Donald H. (1978). "The fluorescence excitation spectrum of the ArI2 van der Waals complex". The Journal of Chemical Physics. 68 (10): 4477. Bibcode:1978JChPh..68.4477K. doi:10.1063/1.435530.
  54. Rohrbacher, Andreas; Halberstadt, Nadine; Janda, Kenneth C. (October 2000). "The Dynamics of Noble Gas – Halogen Molecules and Clusters". Annual Review of Physical Chemistry. 51 (1): 405–433. doi:10.1146/annurev.physchem.51.1.405.
  55. Fei, Suli; Zheng, Xiaonan; Heaven, Michael C.; Tellinghuisen, Joel (1992). "Spectroscopy and relaxation dynamics of I2Arn clusters. Geminate recombination and cluster fragmentation". The Journal of Chemical Physics. 97 (9): 6057. doi:10.1063/1.463716.
  56. Harris, Stephen J. (1974). "Intermolecular potential between an atom and a diatomic molecule: The structure of ArCIF". The Journal of Chemical Physics. 61 (1): 193. doi:10.1063/1.1681622.
  57. Zheng, Rui; Li, Song; Chen, Shan-Jun; Chen, Yan; Zheng, Li-Min (September 2015). "Theoretical investigation of potential energy surface and bound states for the van der Waals complex Ar–BrCl dimer". Chemical Physics. 458: 77–85. doi:10.1016/j.chemphys.2015.07.013.
  58. Liu, Suyan; Bačić, Zlatko; Moskowitz, Jules W.; Schmidt, Kevin E. (1994). "ArnH2O (n = 1–14) van der Waals clusters: Size evolution of equilibrium structures". The Journal of Chemical Physics. 101 (10): 8310. doi:10.1063/1.468097.
  59. Gascooke, Jason R.; Alexander, Ula N.; Lawrance, Warren D. (2012). "Two-dimensional laser induced fluorescence spectroscopy of van der Waals complexes: Fluorobenzene-Arn (n = 1,2)". The Journal of Chemical Physics. 136 (13): 134309. Bibcode:2012JChPh.136m4309G. doi:10.1063/1.3697474. PMID 22482554.
  60. Wright, Scott A.; Dagdigian, Paul J. (1997). "Fluorescence excitation spectroscopy of the Ar–HCO(X̃2Aʹ,B̃2Aʹ) van der Waals complex". The Journal of Chemical Physics. 107 (3): 680. Bibcode:1997JChPh.107..680W. doi:10.1063/1.474469.
  61. Kang, Cheolhwa; Yi, John T.; Pratt, David W. (2005). "High resolution electronic spectra of 7-azaindole and its Ar atom van der Waals complex". The Journal of Chemical Physics. 123 (9): 094306. Bibcode:2005JChPh.123i4306K. doi:10.1063/1.1990119.
  62. Lapierre, Luc; Frye, Donald; Dai, Hai-Lung (1992). "Isomeric structures and van der Waals vibrational frequencies of the glyoxal·Ar complexes. I. Fluorescence excitation spectroscopy". The Journal of Chemical Physics. 96 (4): 2703. Bibcode:1992JChPh..96.2703L. doi:10.1063/1.462018.
  63. Mizoguchi, Asao; Endo, Yasuki; Ohshima, Yasuhiro (1998). "Rotational spectrum of a salt-containing van der Waals complex: Ar–NaCl". The Journal of Chemical Physics. 109 (24): 10539. Bibcode:1998JChPh.10910539M. doi:10.1063/1.477754.
  64. Girardet, C.; Lakhlifi, A.; Laroui, B. (1992). "Infrared profile of van der Waals dimers HCl–RG* (RG*=Ar, Kr, Xe) trapped in rare gas matrices". The Journal of Chemical Physics. 97 (11): 7955. Bibcode:1992JChPh..97.7955G. doi:10.1063/1.463470.
  65. Lin, Wei; Brooks, Andrew H.; Minei, Andrea J.; Novick, Stewart E.; Pringle, Wallace C. (6 February 2014). "Microwave Spectra and Structure of the Argon–Cyclopentanone and Neon–Cyclopentanone van der Waals Complexes". The Journal of Physical Chemistry A. 118 (5): 856–861. doi:10.1021/jp410381r.
  66. Demaison, J. (2010). "172 C2H4ArF2 1,1-Difluoroethane - argon (1/1)". 29D1: 351–351. doi:10.1007/978-3-642-10371-1_174. ISSN 1616-9530.
  67. Melandri, Sonia; Velino, Biagio; Favero, Paolo G; Dell'Erba, Adele; Caminati, Walther (April 2000). "Investigation of a van der Waals complex with C1 symmetry: the free-jet rotational spectrum of 1,2-difluoroethane–Ar". Chemical Physics Letters. 321 (1–2): 31–36. doi:10.1016/S0009-2614(00)00326-2.
  68. Melandri, Sonia; Dell'Erba, Adele; Favero, Paolo G.; Caminati, Walther (November 2003). "Millimeter-wave investigation, simplified interpretation of the fourfold rotational spectrum, and dynamics of the internal motions of acetaldehyde–argon". Journal of Molecular Spectroscopy. 222 (1): 121–128. doi:10.1016/S0022-2852(03)00197-8.
  69. Moreschini, Paolo; Caminati, Walther; Favero, Paolo G.; Legon, Anthony C. (December 2001). "Pathways for inversion in the oxirane–argon complex". Journal of Molecular Structure. 599 (1–3): 81–87. doi:10.1016/S0022-2860(01)00846-8.
  70. Holleran, Eugene M.; Hennessy, John T.; LaPietra, Frank R. (August 1967). "Anomalous pH changes in solutions of inert gases possibly indicating their basicity". The Journal of Physical Chemistry. 71 (9): 3081–3084. doi:10.1021/j100868a060.
  71. Tadros, T. F.; Sadek, H. (May 1969). "Medium effects in aqueous non-electrolyte solutions". Electrochimica Acta. 14 (5): 377–388. doi:10.1016/0013-4686(69)80014-9.
  72. Hirai, Hisako; Uchihara, Yukako; Nishimura, Yukiko; Kawamura, Taro; Yamamoto, Yoshitaka; Yagi, Takehiko (October 2002). "Structural Changes of Argon Hydrate under High Pressure". The Journal of Physical Chemistry B. 106 (43): 11089–11092. doi:10.1021/jp021458l.
  73. Yang, L.; Tulk, C. A.; Klug, D. D.; Chakoumakos, B. C.; Ehm, L.; Molaison, J. J.; Parise, J. B.; Simonson, J. M. (January 2010). "Guest disorder and high pressure behavior of argon hydrates". Chemical Physics Letters. 485 (1–3): 104–109. Bibcode:2010CPL...485..104Y. doi:10.1016/j.cplett.2009.12.024.
  74. Sugahara, Keisuke; Kaneko, Ryuji; Sasatani, Arata; Sugahara, Takeshi; Ohgaki, Kazunari (30 October 2008). "Thermodynamic and Raman Spectroscopic Studies of Ar Hydrate System". The Open Thermodynamics Journal. 2 (1): 95–99. doi:10.2174/1874396X00802010095.
  75. Räsänen, Markku; Khriachtchev, Leonid; Pettersson, Mika; Runeberg, Nino; Lundell, Jan (24 August 2000). "A stable argon compound". Nature. 406 (6798): 874–876. doi:10.1038/35022551. PMID 10972285.
  76. 1 2 McDowell, Sean (1 May 2006). "Studies of Neutral Rare-Gas Compounds and their Non-Covalent Interactions with Other Molecules". Current Organic Chemistry. 10 (7): 791–803. doi:10.2174/138527206776818964.
  77. Jiménez Halla, C. Óscar C.; Fernández, Israel; Frenking, Gernot (2 January 2009). "Is it Possible To Synthesize a Neutral Noble Gas Compound Containing a Ng–Ng Bond? A Theoretical Study of H–Ng–Ng–F (Ng=Ar, Kr, Xe)". Angewandte Chemie International Edition. 48 (2): 366–369. doi:10.1002/anie.200803252.
  78. Jayasekharan, T.; Ghanty, Tapan K. (2007). "Significant increase in the stability of rare gas hydrides on insertion of beryllium atom". The Journal of Chemical Physics. 127 (11): 114314. Bibcode:2007JChPh.127k4314J. doi:10.1063/1.2768936. PMID 17887844.
  79. 1 2 Liang, Binyong; Andrews, Lester; Li, Jun; Bursten, Bruce E. (August 2002). "Noble Gas−Actinide Compounds: Evidence for the Formation of Distinct CUO(Ar)4−n(Xe)n and CUO(Ar)4−n(Kr)n (n = 1, 2, 3, 4) Complexes". Journal of the American Chemical Society. 124 (31): 9016–9017. doi:10.1021/ja026432m. PMID 12148982.
  80. 1 2 Li, J. (28 February 2002). "Noble Gas–Actinide Compounds: Complexation of the CUO Molecule by Ar, Kr, and Xe Atoms in Noble Gas Matrices". Science. 295 (5563): 2242–2245. Bibcode:2002Sci...295.2242L. doi:10.1126/science.1069342. PMID 11872801.
  81. 1 2 3 4 5 6 Andrews, Lester; Liang, Binyong; Li, Jun; Bursten, Bruce E. (March 2003). "Noble Gas−Actinide Complexes of the CUO Molecule with Multiple Ar, Kr, and Xe Atoms in Noble-Gas Matrices". Journal of the American Chemical Society. 125 (10): 3126–3139. doi:10.1021/ja027819s. PMID 12617681.
  82. 1 2 Wang, Xuefeng; Andrews, Lester; Li, Jun; Bursten, Bruce E. (3 May 2004). "Significant Interactions between Uranium and Noble-Gas Atoms: Coordination of the UO+
    2
    Cation by Ne, Ar, Kr, and Xe Atoms". Angewandte Chemie International Edition. 43 (19): 2554–2557. doi:10.1002/anie.200453790.
  83. Denning, Robert G. (May 2007). "Electronic Structure and Bonding in Actinyl Ions and their Analogs". The Journal of Physical Chemistry A. 111 (20): 4125–4143. Bibcode:2007JPCA..111.4125D. doi:10.1021/jp071061n. PMID 17461564.
  84. Li, Jun; Bursten, Bruce E.; Andrews, Lester; Marsden, Colin J. (March 2004). "On the Electronic Structure of Molecular UO2 in the Presence of Ar Atoms: Evidence for Direct U−Ar Bonding". Journal of the American Chemical Society. 126 (11): 3424–3425. doi:10.1021/ja039933w. PMID 15025460.
  85. Thompson, Craig A.; Andrews, Lester (January 1994). "Noble Gas Complexes with BeO: Infrared Spectra of NG-BeO (NG = Ar, Kr, Xe)". Journal of the American Chemical Society. 116 (1): 423–424. doi:10.1021/ja00080a069.
  86. Veldkamp, A.; Frenking, G. (August 1994). "Structures and bond energies of the noble gas complexes NgBeO (Ng=Ar, Kr, Xe)". Chemical Physics Letters. 226 (1–2): 11–16. Bibcode:1994CPL...226...11V. doi:10.1016/0009-2614(94)00697-0.
  87. Wang, Qiang; Wang, Xuefeng (21 February 2013). "Infrared Spectra of NgBeS (Ng = Ne, Ar, Kr, Xe) and BeS2 in Noble-Gas Matrices". The Journal of Physical Chemistry A. 117 (7): 1508–1513. Bibcode:2013JPCA..117.1508W. doi:10.1021/jp311901a. PMID 23327099.
  88. Zhang, Qingnan; Chen, Mohua; Zhou, Mingfei; Andrada, Diego M.; Frenking, Gernot (19 March 2015). "Experimental and Theoretical Studies of the Infrared Spectra and Bonding Properties of NgBeCO3 and a Comparison with NgBeO (Ng = He, Ne, Ar, Kr, Xe)". The Journal of Physical Chemistry A. 119 (11): 2543–2552. Bibcode:2015JPCA..119.2543Z. doi:10.1021/jp509006u. PMID 25321412.
  89. Young, Nigel A. (March 2013). "Main group coordination chemistry at low temperatures: A review of matrix isolated Group 12 to Group 18 complexes". Coordination Chemistry Reviews. 257 (5–6): 956–1010. doi:10.1016/j.ccr.2012.10.013.
  90. 1 2 Ehlers, Andreas W.; Frenking, Gernot; Baerends, Evert Jan (October 1997). "Structure and Bonding of the Noble Gas−Metal Carbonyl Complexes M(CO)5−Ng (M = Cr, Mo, W and Ng = Ar, Kr, Xe)". Organometallics. 16 (22): 4896–4902. doi:10.1021/om9704962.
  91. Wells, J. R.; Weitz, Eric (April 1992). "Rare gas-metal carbonyl complexes: bonding of rare gas atoms to the Group VIB pentacarbonyls". Journal of the American Chemical Society. 114 (8): 2783–2787. doi:10.1021/ja00034a003.
  92. Sun, Xue-Zhong; George, Michael W.; Kazarian, Sergei G.; Nikiforov, Sergei M.; Poliakoff, Martyn (January 1996). "Can Organometallic Noble Gas Compounds Be Observed in Solution at Room Temperature? A Time-Resolved Infrared (TRIR) and UV Spectroscopic Study of the Photochemistry of M(CO)5 (M = Cr, Mo, and W) in Supercritical Noble Gas and CO2 Solution". Journal of the American Chemical Society. 118 (43): 10525–10532. doi:10.1021/ja960485k.
  93. Hope, Eric G. (March 2013). "Coordination chemistry of the noble gases and noble gas fluorides". Coordination Chemistry Reviews. 257 (5–6): 902–909. doi:10.1016/j.ccr.2012.07.017.
  94. Grochala, Wojciech (2007). "Atypical compounds of gases, which have been called 'noble'". Chemical Society Reviews. 36 (10): 1632–1655. doi:10.1039/B702109G. PMID 17721587.
  95. 1 2 3 4 5 6 Evans, Corey J.; Lesarri, Alberto; Gerry, Michael C. L. (June 2000). "Noble Gas−Metal Chemical Bonds. Microwave Spectra, Geometries, and Nuclear Quadrupole Coupling Constants of Ar−AuCl and Kr−AuCl". Journal of the American Chemical Society. 122 (25): 6100–6105. doi:10.1021/ja000874l.
  96. 1 2 Zhao, Yanying; Gong, Yu; Chen, Mohua; Zhou, Mingfei (February 2006). "Noble Gas−Transition-Metal Complexes: Coordination of VO2 and VO4 by Ar and Xe Atoms in Solid Noble Gas Matrixes". The Journal of Physical Chemistry A. 110 (5): 1845–1849. Bibcode:2006JPCA..110.1845Z. doi:10.1021/jp056476s. PMID 16451016.
  97. Zhao, Yanying; Wang, Guanjun; Chen, Mohua; Zhou, Mingfei (August 2005). "Noble Gas−Transition Metal Complexes: Coordination of ScO+ by Multiple Ar, Kr, and Xe Atoms in Noble Gas Matrixes". The Journal of Physical Chemistry A. 109 (30): 6621–6623. Bibcode:2005JPCA..109.6621Z. doi:10.1021/jp053148j. PMID 16834012.
  98. Zhao, Yanying; Gong, Yu; Chen, Mohua; Ding, Chuanfan; Zhou, Mingfei (December 2005). "Coordination of ScO+ and YO+ by Multiple Ar, Kr, and Xe Atoms in Noble Gas Matrixes: A Matrix Isolation Infrared Spectroscopic and Theoretical Study". The Journal of Physical Chemistry A. 109 (51): 11765–11770. Bibcode:2005JPCA..10911765Z. doi:10.1021/jp054517e. PMID 16366626.
  99. 1 2 Zhao, Yanying; Gong, Yu; Zhou, Mingfei (September 2006). "Matrix Isolation Infrared Spectroscopic and Theoretical Study of NgMO (Ng = Ar, Kr, Xe; M = Cr, Mn, Fe, Co, Ni) Complexes". The Journal of Physical Chemistry A. 110 (37): 10777–10782. Bibcode:2006JPCA..11010777Z. doi:10.1021/jp064100o. PMID 16970371.
  100. Zhao, Yanying; Zheng, Xuming; Zhou, Mingfei (July 2008). "Coordination of niobium and tantalum oxides by Ar, Xe and O2: Matrix isolation infrared spectroscopic and theoretical study of NbO2(Ng)2 (Ng=Ar, Xe) and MO4(X) (M=Nb, Ta; X=Ar, Xe, O2) in solid argon". Chemical Physics. 351 (1–3): 13–18. Bibcode:2008CP....351...13Z. doi:10.1016/j.chemphys.2008.03.026.
  101. Yang, Rui; Gong, Yu; Zhou, Han; Zhou, Mingfei (January 2007). "Matrix Isolation Infrared Spectroscopic and Theoretical Study of Noble Gas Coordinated Rhodium−Dioxygen Complexes". The Journal of Physical Chemistry A. 111 (1): 64–70. Bibcode:2007JPCA..111...64Y. doi:10.1021/jp0662000. PMID 17201389.
  102. Brock, D. S.; Schrobilgen, G. J. (9 October 2014). "Noble-Gas Chemistry". Comprehensive Inorganic Chemistry II. 1: Main-Group Elements, Including Noble Gases (2): 755–822. doi:10.1016/B978-0-08-097774-4.00128-5.
  103. Zhao, YanYing; Zhou, MingFei (21 February 2010). "Are matrix isolated species really "isolated"? Infrared spectroscopic and theoretical studies of noble gas–transition metal oxide complexes". Science China Chemistry. 53 (2): 327–336. doi:10.1007/s11426-010-0044-9.
  104. Ono, Yuriko; Taketsugu, Tetsuya (2004). "Theoretical study of Ng–NiN2 (Ng=Ar,Ne,He)". The Journal of Chemical Physics. 120 (13): 6035. Bibcode:2004JChPh.120.6035O. doi:10.1063/1.1650310. PMID 15267486.
  105. Koyanagi, Gregory K.; Bohme, Diethard K. (7 January 2010). "Strong Closed-Shell Interactions: Observed Formation of BaRg2+ Molecules in the Gas Phase at Room Temperature". The Journal of Physical Chemistry Letters. 1 (1): 41–44. doi:10.1021/jz900009q.
  106. Lockyear, Jessica F.; Parkes, Michael A.; Price, Stephen D. (7 February 2011). "Fast and Efficient Fluorination of Small Molecules by SF42+". Angewandte Chemie. 123 (6): 1358–1360. doi:10.1002/ange.201006486.
  107. Berkowitz, J.; Chupka, W. A. (November 1970). "Diatomic ions of noble gas fluorides". Chemical Physics Letters. 7 (4): 447–450. doi:10.1016/0009-2614(70)80331-1.
  108. 1 2 3 Frenking, Gernot; Koch, Wolfram; Deakyne, Carol A.; Liebman, Joel F.; Bartlett, Neil (January 1989). "The ArF+ cation. Is it stable enough to be isolated in a salt?". Journal of the American Chemical Society. 111 (1): 31–33. doi:10.1021/ja00183a005.
  109. Bogey, M.; Cordonnier, M.; Demuynck, C.; Destombes, J. L. (1992). "The millimeter- and submillimeter-wave spectrum of ArF+". Journal of Molecular Spectroscopy. 155 (1): 217–219.
  110. Selig, Henry; Holloway, John H. (27 May 2005). "Cationic and anionic complexes of the noble gases". Topics in Current Chemistry. 124: 33–90. doi:10.1007/3-540-13534-0_2.
  111. 1 2 Henglein, A.; Muccini, G. A. (October 1960). "Neue Ion–Molekül-Reaktionen und ihre Bedeutung für die Strahlenchemie" [New ion–molecule reactions and their importance for radiation chemistry]. Berichte der Bunsengesellschaft für physikalische Chemie (in German). 64 (8–9): 1015. doi:10.1002/bbpc.19600640807. ISSN 0005-9021.
  112. 1 2 Alcock, Nancy (1 June 1993). "Flame, Flameless, and Plasma Spectroscopy". Analytical Chemistry. 65 (12): 463–469. doi:10.1021/ac00060a618. ISSN 0003-2700.
  113. Sheppard, Brenda S.; Shen, Wei-Lung; Caruso, Joseph A.; Heitkemper, Douglas T.; Fricke, Fred L. (1990). "Elimination of the argon chloride interference on arsenic speciation in inductively coupled plasma mass spectrometry using ion chromatography". Journal of Analytical Atomic Spectrometry. 5 (6): 431. doi:10.1039/JA9900500431.
  114. MacKay, M.; Rusho, W.; Jackson, D.; McMillin, G.; Winther, B. (2 December 2009). "Physical and Chemical Stability of Iron Sucrose in Parenteral Nutrition". Nutrition in Clinical Practice. 24 (6): 733–737. doi:10.1177/0884533609351528. PMID 19955552.
  115. Lenzer, Thomas; Yourshaw, Ivan; Furlanetto, Michael R.; Reiser, Georg; Neumark, Daniel M. (1999). "Zero electron kinetic energy spectroscopy of the ArCl anion". The Journal of Chemical Physics. 110 (19): 9578. doi:10.1063/1.478923.
  116. 1 2 Koskinen, Jere T. (December 1999). "Novel Rare Gas Ions BXe+, BKr+, and BAr+ Formed in a Halogen/Rare Gas Exchange Reaction". The Journal of Physical Chemistry A. 103 (48): 9565–9568. doi:10.1021/jp993091z.
  117. Flesch, G. D.; Ng, C. Y. (1988). "Observation of the formation of C+, O+, and ArC+ in the collisions of Ar+(2P3/2,1/2) with CO". The Journal of Chemical Physics. 89 (5): 3381. doi:10.1063/1.455708.
  118. Flesch, G. D.; Ng, C. Y. (1990). "Observation of the formation of N+ and ArN+ in the collisions of Ar+(2P3/2,1/2) with N2". The Journal of Chemical Physics. 92 (5): 2876. doi:10.1063/1.457934.
  119. 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 Becker, J. Sabine; Seifert, Gotthard; Saprykin, Anatoli I.; Dietze, Hans-Joachim (1996). "Mass spectrometric and theoretical investigations into the formation of argon molecular ions in plasma mass spectrometry". Journal of Analytical Atomic Spectrometry. 11 (9): 643. doi:10.1039/JA9961100643.
  120. Levee, Lauren; Calogero, Catherine; Barbieri, Edward; Byrne, Steven; Donahue, Courtney; Eisenberg, Michael; Hattenbach, Sean; Le, Julie; Capitani, Joseph F.; Roithová, Jana; Schröder, Detlef (June 2012). "Formation of argon–boron bonds in the reactions of BFn+/2+ cations with neutral argon". International Journal of Mass Spectrometry. 323-324: 2–7. doi:10.1016/j.ijms.2012.05.006.
  121. Ascenzi, Daniela; Tosi, Paolo; Roithová, Jana; Ricketts, Claire L.; Schröder, Detlef; Lockyear, Jessica F.; Parkes, Michael A.; Price, Stephen D. (2008). "Generation of the organo-rare gas dications HCCRg2+ (Rg = Ar and Kr) in the reaction of acetylene dications with rare gases". Physical Chemistry Chemical Physics. 10 (47): 7121. doi:10.1039/B810398D.
  122. Roithová, Jana; Schröder, Detlef (2 November 2009). "Siliciumverbindungen von Neon und Argon". Angewandte Chemie (in German). 121 (46): 8946–8948. doi:10.1002/ange.200903706.
  123. Hillier, I. H.; Guest, M. F.; Ding, A.; Karlau, J.; Weise, J. (1979). "The potential energy curves of ArC+". The Journal of Chemical Physics. 70 (2): 864. Bibcode:1979JChPh..70..864H. doi:10.1063/1.437518.
  124. Broström, L.; Larsson, M.; Mannervik, S.; Sonnek, D. (1991). "The visible photoabsorption spectrum and potential curves of ArN+". The Journal of Chemical Physics. 94 (4): 2734. Bibcode:1991JChPh..94.2734B. doi:10.1063/1.459850.
  125. Freeman, J. H.; McIlroy, R. W. (4 January 1964). "Preparation of Ion-molecules of the Inert Gases in an Electromagnetic Isotope Separator". Nature. 201 (4914): 69–70. Bibcode:1964Natur.201...69F. doi:10.1038/201069a0.
  126. Gadd, G. E.; Moricca, S.; Kennedy, S. J.; Elcombe, M. M.; Evans, P. J.; Blackford, M.; Cassidy, D.; Howard, C. J.; Prasad, P.; Hanna, J. V.; Burchwood, A.; Levy, D. (November 1997). "Novel rare gas interstitial fullerenes of C60 with Ar, Kr and Xe". Journal of Physics and Chemistry of Solids. 58 (11): 1823–1832. Bibcode:1997JPCS...58.1823G. doi:10.1016/S0022-3697(97)00096-6.
  127. Gadd, G. E.; Elcombe, M. M.; Dennis, J.; Moricca, S.; Webb, N.; Cassidy, D.; Evans, P. J. (June 1998). "Novel rare gas interstitial fullerenes of C70". Journal of Physics and Chemistry of Solids. 59 (6–7): 937–944. Bibcode:1998JPCS...59..937G. doi:10.1016/S0022-3697(98)00017-1.
  128. Ujihara, Yuki; Takahashi, Yutaka (2011). "Binary Solid Solution of C60–C70 Prepared by Liquid–Liquid Interfacial Precipitation Method". Journal of the Japan Institute of Metals. 75 (12): 671–677. doi:10.2320/jinstmet.75.671.
  129. Kurzydłowski, Dominik; Zaleski-Ejgierd, Patryk (2016). "High-pressure stabilization of argon fluorides". Phys. Chem. Chem. Phys. 18 (4): 2309–2313. Bibcode:2016PCCP...18.2309K. doi:10.1039/C5CP05725F. PMID 26742478.
  130. Layer, M.; Heitz, M.; Meier, J.; Hunklinger, S. (January 2005). "On solidified mixtures of neon and argon—Evidence for the formation of Ne2Ar- and Ar2Ne-phases". Europhysics Letters (EPL). 69 (1): 95–101. Bibcode:2005EL.....69...95L. doi:10.1209/epl/i2004-10298-x.
  131. Layer, M.; Netsch, A.; Heitz, M.; Meier, J.; Hunklinger, S. (17 May 2006). "Mixing behavior and structural formation of quench-condensed binary mixtures of solid noble gases". Physical Review B. 73 (18): 184116. Bibcode:2006PhRvB..73r4116L. doi:10.1103/PhysRevB.73.184116.
  132. 1 2 Hemley, R. J.; Dera, P. (1 January 2000). "Molecular Crystals". Reviews in Mineralogy and Geochemistry. 41 (1): 369. doi:10.2138/rmg.2000.41.12.
  133. Weck, Gunnar; Dewaele, Agnès; Loubeyre, Paul (28 July 2010). "Oxygen/noble gas binary phase diagrams at 296 K and high pressures". Physical Review B. 82 (1): 014112. Bibcode:2010PhRvB..82a4112W. doi:10.1103/PhysRevB.82.014112.
  134. Cazorla, C.; Errandonea, D.; Sola, E. (10 August 2009). "High-pressure phases, vibrational properties, and electronic structure of Ne(He)2 and Ar(He)2: A first-principles study". Physical Review B. 80 (6): 064105–1. Bibcode:2009PhRvB..80f4105C. doi:10.1103/PhysRevB.80.064105.

External links

This article is issued from Wikipedia - version of the 12/1/2016. The text is available under the Creative Commons Attribution/Share Alike but additional terms may apply for the media files.