Xenon

From Self-sufficiency
Jump to: navigation, search
iodinexenoncaesium
Kr

Xe

Rn
Appearance
Colorless gas, exhibiting a blue glow when placed in a high voltage electric field
250px
250px
Spectral lines of xenon
General properties
Name, symbol, number xenon, Xe, 54
Pronunciation /ˈzɛnɒn/[1] ZEN-on
or /ˈziːnɒn/[2] ZEE-non
Element category noble gases
Group, period, block 185, p
Standard atomic weight 131.293(6)g·mol−1
Electron configuration [Kr] 5s2 4d10 5p6
Electrons per shell 2, 8, 18, 18, 8 (Image)
Physical properties
Phase gas
Density (0 °C, 101.325 kPa)
5.894 g/L
Melting point (101.325 kPa) 161.4 K, −111.7 °C, −169.1 °F
Boiling point (101.325 kPa) 165.03 K, −108.12 °C, −162.62 °F
Triple point 161.405 K (-112°C), 81.6[3] kPa
Critical point 289.77 K, 5.841 MPa
Heat of fusion (101.325 kPa) 2.27 kJ·mol−1
Heat of vaporization (101.325 kPa) 12.64 kJ·mol−1
Specific heat capacity (100 kPa, 25 °C) 20.786 J·mol−1·K−1
Vapor pressure
P (Pa) 1 10 100 1 k 10 k 100 k
at T (K) 83 92 103 117 137 165
Atomic properties
Oxidation states 0, +1, +2, +4, +6, +8
(rarely more than 0)
(weakly acidic oxide)
Electronegativity 2.6 (Pauling scale)
Ionization energies 1st: 1170.4 kJ·mol−1
2nd: 2046.4 kJ·mol−1
3rd: 3099.4 kJ·mol−1
Covalent radius 140±9 pm
Van der Waals radius 216 pm
Miscellanea
Crystal structure face-centered cubic
Magnetic ordering diamagnetic[4]
Thermal conductivity (300 K) 5.65×10-3  W·m−1·K−1
Speed of sound (liquid) 1090 m/s; (gas) 169 m/s
CAS registry number 7440-63-3
Most stable isotopes
Main article: Isotopes of xenon
iso NA half-life DM DE (MeV) DP
124Xe 0.095% 124Xe is stable with 70 neutrons
125Xe syn 16.9 h ε 1.652 125I
126Xe 0.089% 126Xe is stable with 72 neutrons
127Xe syn 36.345 d ε 0.662 127I
128Xe 1.91% 128Xe is stable with 74 neutrons
129Xe 26.4% 129Xe is stable with 75 neutrons
130Xe 4.07% 130Xe is stable with 76 neutrons
131Xe 21.2% 131Xe is stable with 77 neutrons
132Xe 26.9% 132Xe is stable with 78 neutrons
133Xe syn 5.247 d β 0.427 133Cs
134Xe 10.4% 134Xe is stable with 80 neutrons
135Xe syn 9.14 h β 1.16 135Cs
136Xe 8.86% 136Xe is stable with 82 neutrons

Xenon is a chemical element represented by the symbol Xe. The element name is pronounced /ˈzɛnɒn/, ZEN-on or /ˈziːnɒn/, ZEE-non. Its atomic number is 54. A colorless, heavy, odorless noble gas, xenon occurs in the Earth's atmosphere in trace amounts.[5] Although generally unreactive, xenon can undergo a few chemical reactions such as the formation of xenon hexafluoroplatinate, the first noble gas compound to be synthesized.[6][7][8]

Naturally occurring xenon consists of nine stable isotopes. There are also over 40 unstable isotopes that undergo radioactive decay. The isotope ratios of xenon are an important tool for studying the early history of the Solar System.[9] Xenon-135 is produced as a result of nuclear fission and acts as a neutron absorber in nuclear reactors.[10]

Xenon is used in flash lamps[11] and arc lamps,[12] and as a general anesthetic.[13] The first excimer laser design used a xenon dimer molecule (Xe2) as its lasing medium,[14] and the earliest laser designs used xenon flash lamps as pumps.[15] Xenon is also being used to search for hypothetical weakly interacting massive particles[16] and as the propellant for ion thrusters in spacecraft.[17]

History

Xenon was discovered in England by William Ramsay and Morris Travers on July 12, 1898, shortly after their discovery of the elements krypton and neon. They found it in the residue left over from evaporating components of liquid air.[18][19] Ramsay suggested the name xenon for this gas from the Greek word ξένον [xenon], neuter singular form of ξένος [xenos], meaning 'foreign(er)', 'strange(r)', or 'guest'.[20][21] In 1902, Ramsay estimated the proportion of xenon in the Earth's atmosphere as one part in 20 million.[22]

During the 1930s, engineer Harold Edgerton began exploring strobe light technology for high speed photography. This led him to the invention of the xenon flash lamp, in which light is generated by sending a brief electrical current through a tube filled with xenon gas. In 1934, Edgerton was able to generate flashes as brief as one microsecond with this method.[11][23][24]

In 1939, Albert R. Behnke Jr. began exploring the causes of "drunkenness" in deep-sea divers. He tested the effects of varying the breathing mixtures on his subjects, and discovered that this caused the divers to perceive a change in depth. From his results, he deduced that xenon gas could serve as an anesthetic. Although Lazharev, in Russia, apparently studied xenon anesthesia in 1941, the first published report confirming xenon anesthesia was in 1946 by J. H. Lawrence, who experimented on mice. Xenon was first used as a surgical anesthetic in 1951 by Stuart C. Cullen, who successfully operated on two patients.[25]

Xenon and the other noble gases were for a long time considered to be completely chemically inert and not able to form compounds. However, while teaching at the University of British Columbia, Neil Bartlett discovered that the gas platinum hexafluoride (PtF6) was a powerful oxidizing agent that could oxidize oxygen gas (O2) to form dioxygenyl hexafluoroplatinate (O2+[PtF6]).[26] Since O2 and xenon have almost the same first ionization potential, Bartlett realized that platinum hexafluoride might also be able to oxidize xenon. On March 23, 1962, he mixed the two gases and produced the first known compound of a noble gas, xenon hexafluoroplatinate.[27][8] Bartlett thought its composition to be Xe+[PtF6], although later work has revealed that it was probably a mixture of various xenon-containing salts.[28][29][30] Since then, many other xenon compounds have been discovered,[31] along with some compounds of the noble gases argon, krypton, and radon, including argon fluorohydride (HArF),[32] krypton difluoride (KrF2),[33][34] and radon fluoride.[35] By 1971, more than 80 xenon compounds were known.[36][37]

Characteristics

An atom of xenon is defined as having a nucleus with 54 protons. At standard temperature and pressure, pure xenon gas has a density of 5.761 kg/m3, about 4.5 times the surface density of the Earth's atmosphere, 1.217 kg/m3.[38] As a liquid, xenon has a density of up to 3.100 g/mL, with the density maximum occurring at the triple point.[39] Under the same conditions, the density of solid xenon, 3.640 g/cm3, is larger than the average density of granite, 2.75 g/cm3.[39] Using gigapascals of pressure, xenon has been forced into a metallic phase.[40]

Solid xenon changes from face-centered cubic (fcc) to hexagonal close packed (hcp) crystal phase under pressure and begins to turn metallic at about 140 GPa, with no noticeable volume change in the hcp phase. It is completely metallic at 155 GPa. When metalized, xenon looks sky blue because it absorbs red light and transmits other visible frequencies. Such behavior is unusual for a metal and is explained by the relatively small widths of the electron bands in metallic xenon.[41][42]

Xenon is a member of the zero-valence elements that are called noble or inert gases. It is inert to most common chemical reactions (such as combustion, for example) because the outer valence shell contains eight electrons. This produces a stable, minimum energy configuration in which the outer electrons are tightly bound.[43] However, xenon can be oxidized by powerful oxidizing agents, and many xenon compounds have been synthesized.

In a gas-filled tube, xenon emits a blue or lavenderish glow when the gas is excited by electrical discharge. Xenon emits a band of emission lines that span the visual spectrum,[44] but the most intense lines occur in the region of blue light, which produces the coloration.[45]

Occurrence and production

Xenon is a trace gas in Earth's atmosphere, occurring at 0.087±0.001 parts per million (μL/L), or approximately 1 part per 11.5 million,[46] and is also found in gases emitted from some mineral springs.

Xenon is obtained commercially as a byproduct of the separation of air into oxygen and nitrogen. After this separation, generally performed by fractional distillation in a double-column plant, the liquid oxygen produced will contain small quantities of krypton and xenon. By additional fractional distillation steps, the liquid oxygen may be enriched to contain 0.1–0.2% of a krypton/xenon mixture, which is extracted either via adsorption onto silica gel or by distillation. Finally, the krypton/xenon mixture may be separated into krypton and xenon via distillation.[47][48] Extraction of a liter of xenon from the atmosphere requires 220 watt-hours of energy.[49] Worldwide production of xenon in 1998 was estimated at 5,000–7,000 m3.[50] Because of its low abundance, xenon is much more expensive than the lighter noble gases—approximate prices for the purchase of small quantities in Europe in 1999 were 10 /L for xenon, 1 €/L for krypton, and 0.20 €/L for neon.[50]

Within the Solar System, the nucleon fraction of xenon is 1.56 × 10−8, for an abundance of one part in 64 million of the total mass.[51] Xenon is relatively rare in the Sun's atmosphere, on Earth, and in asteroids and comets. The planet Jupiter has an unusually high abundance of xenon in its atmosphere; about 2.6 times as much as the Sun.[52] This high abundance remains unexplained and may have been caused by an early and rapid buildup of planetesimals—small, subplanetary bodies—before the presolar disk began to heat up.[53] (Otherwise, xenon would not have been trapped in the planetesimal ices.) The problem of the low terrestrial xenon may potentially be explained by covalent bonding of xenon to oxygen within quartz, hence reducing the outgassing of xenon into the atmosphere.[54]

Unlike the lower mass noble gases, the normal stellar nucleosynthesis process inside a star does not form xenon. Elements more massive than iron-56 have a net energy cost to produce through fusion, so there is no energy gain for a star when creating xenon.[55] Instead, xenon is formed during supernova explosions,[56] by the slow neutron capture process (s-process) of red giant stars that have exhausted the hydrogen at their cores and entered the asymptotic giant branch,[57] in classical novae explosions[58] and from the radioactive decay of elements such as iodine, uranium and plutonium.[59]

Isotopes and isotopic studies

Naturally occurring xenon is made of nine stable isotopes, the most of any element with the exception of tin, which has ten. Xenon and tin are the only elements to have more than seven stable isotopes.[60] The isotopes 124Xe, 134Xe and 136Xe are predicted to undergo double beta decay, but this has never been observed so they are considered to be stable.[61] Besides these stable forms, there are over 40 unstable isotopes that have been studied. 129Xe is produced by beta decay of 129I, which has a half-life of 16 million years, while 131mXe, 133Xe, 133mXe, and 135Xe are some of the fission products of both 235U and 239Pu,[59] and therefore used as indicators of nuclear explosions.

Nuclei of two of the stable isotopes of xenon, 129Xe and 131Xe, have non-zero intrinsic angular momenta (nuclear spins, suitable for nuclear magnetic resonance). The nuclear spins can be aligned beyond ordinary polarization levels by means of circularly polarized light and rubidium vapor.[62] The resulting spin polarization of xenon nuclei can surpass 50% of its maximum possible value, greatly exceeding the equilibrium value dictated by the Boltzmann distribution (typically 0.001% of the maximum value at room temperature, even in the strongest magnets). Such non-equilibrium alignment of spins is a temporary condition, and is called hyperpolarization. The process of hyperpolarizing the xenon is called optical pumping (although the process is different from pumping a laser).[63]

Because a 129Xe nucleus has a spin of 1/2, and therefore a zero electric quadrupole moment, the 129Xe nucleus does not experience any quadrupolar interactions during collisions with other atoms, and thus its hyperpolarization can be maintained for long periods of time even after the laser beam has been turned off and the alkali vapor removed by condensation on a room-temperature surface. Spin polarization of 129Xe it can persist from several seconds for xenon atoms dissolved in blood[64] to several hours in the gas phase[65] and several days in deeply frozen solid xenon.[66] In contrast, 131Xe has a nuclear spin value of 3/2 and a nonzero quadrupole moment, and has T1 relaxation times in the millisecond and second ranges.[67]

Some radioactive isotopes of xenon, for example, 133Xe and 135Xe, are produced by neutron irradiation of fissionable material within nuclear reactors.[6] 135Xe is of considerable significance in the operation of nuclear fission reactors. 135Xe has a huge cross section for thermal neutrons, 2.6×106 barns,[10] so it acts as a neutron absorber or "poison" that can slow or stop the chain reaction after a period of operation. This was discovered in the earliest nuclear reactors built by the American Manhattan Project for plutonium production. Fortunately the designers had made provisions in the design to increase the reactor's reactivity (the number of neutrons per fission that go on to fission other atoms of nuclear fuel).[68] 135Xe reactor poisoning played a major role in the Chernobyl disaster.[69] A shutdown or decrease of power of a reactor can result in buildup of 135Xe and getting the reactor into the iodine pit.

Under adverse conditions, relatively high concentrations of radioactive xenon isotopes may be found emanating from nuclear reactors due to the release of fission products from cracked fuel rods,[70] or fissioning of uranium in cooling water.[71]

Because xenon is a tracer for two parent isotopes, xenon isotope ratios in meteorites are a powerful tool for studying the formation of the solar system. The iodine-xenon method of dating gives the time elapsed between nucleosynthesis and the condensation of a solid object from the solar nebula. In 1960, physicist John H. Reynolds discovered that certain meteorites contained an isotopic anomaly in the form of an overabundance of xenon-129. He inferred that this was a decay product of radioactive iodine-129. This isotope is produced slowly by cosmic ray spallation and nuclear fission, but is produced in quantity only in supernova explosions. As the half-life of 129I is comparatively short on a cosmological time scale, only 16 million years, this demonstrated that only a short time had passed between the supernova and the time the meteorites had solidified and trapped the 129I. These two events (supernova and solidification of gas cloud) were inferred to have happened during the early history of the Solar System, as the 129I isotope was likely generated before the Solar System was formed, but not long before, and seeded the solar gas cloud with isotopes from a second source. This supernova source may also have caused collapse of the solar gas cloud.[72][73]

In a similar way, xenon isotopic ratios such as 129Xe/130Xe and 136Xe/130Xe are also a powerful tool for understanding planetary differentiation and early outgassing.[9] For example, The atmosphere of Mars shows a xenon abundance similar to that of Earth: 0.08 parts per million,[74] however Mars shows a higher proportion of 129Xe than the Earth or the Sun. As this isotope is generated by radioactive decay, the result may indicate that Mars lost most of its primordial atmosphere, possibly within the first 100 million years after the planet was formed.[75][76] In another example, excess 129Xe found in carbon dioxide well gases from New Mexico was believed to be from the decay of mantle-derived gases soon after Earth's formation.[59][77]

Compounds

See also: Category:Xenon compounds

After Neil Bartlett's discovery in 1962 that xenon can form chemical compounds, a large number of xenon compounds have been discovered and described. Almost all known xenon compounds contain the electronegative atoms fluorine or oxygen.[78]

Halides

File:Xenon tetrafluoride.JPG
XeF4 crystals, 1962

Three fluorides are known: XeF2, XeF4, and XeF6. The fluorides are the starting point for the synthesis of almost all xenon compounds.

The solid, crystalline difluoride XeF2 is formed when a mixture of fluorine and xenon gases is exposed to ultraviolet light.[79] Ordinary daylight is sufficient.[80] Long-term heating of XeF2 at high temperatures under an NiF2 catalyst yields XeF6.[81] Pyrolysis of XeF6 in the presence of NaF yields high-purity XeF4.[82]

The xenon fluorides behave as both fluoride acceptors and fluoride donors, forming salts that contain such cations as XeF+ and Xe2F3+, and anions such as XeF5, XeF7, and XeF82−. The green, paramagnetic Xe2+ is formed by the reduction of XeF2 by xenon gas.[78]

XeF2 is also able to form coordination complexes with transition metal ions. Over 30 such complexes have been synthesized and characterized.[81]

Whereas the xenon fluorides are well-characterized, the other halides are not known, the only exception being the dichloride, XeCl2. Xenon dichloride is reported to be an endothermic, colorless, crystalline compound that decomposes into the elements at 80°C, formed by the high-frequency irradiation of a mixture of xenon, fluorine, and silicon or carbon tetrachloride.[83] However, doubt has been raised as to whether XeCl2 is a real compound and not merely a van der Waals molecule consisting of weakly bound Xe atoms and Cl2 molecules.[84] Theoretical calculations indicate that the linear molecule XeCl2 is less stable than the van der Waals complex.[85]

Oxides and oxohalides

Only two oxides of xenon are known: xenon trioxide (XeO3) and xenon tetroxide (XeO4), both of which are dangerously explosive and powerful oxidizing agents. Xenon dioxide (XeO2) remains elusive — only the XeOO+ cation has been identified by infrared spectroscopy in solid argon.[86]

Xenon does not react with oxygen directly; the trioxide is formed by the hydrolysis of XeF6:[87]

XeF6 + 3 H2OXeO3 + 6 HF

XeO3 is weakly acidic, dissolving in alkali to form unstable xenate salts containing the HXeO4 anion. These unstable salts easily disproportionate into xenon gas and perxenate salts, containing the XeO4−6 anion.[88]

Barium perxenate, when treated with concentrated sulfuric acid, yields gaseous xenon tetroxide:[83]

Ba2XeO6 + 2 H2SO4 → 2 BaSO4 + 2 H2O + XeO4

To prevent decomposition, the xenon tetroxide thus formed is quickly cooled to form a pale-yellow solid. It explodes above −35.9 °C into xenon and oxygen gas.

A number of xenon oxyfluorides are known, including XeOF2, XeOF4, XeO2F2, and XeO3F2. XeOF2 is formed by the reaction of OF2 with xenon gas at low temperatures. It may also be obtained by the partial hydrolysis of XeF4. It disproportionates at −20 °C into XeF2 and XeO2F2.[89] XeOF4 is formed by the partial hydrolysis of XeF6,[90] or the reaction of XeF6 with sodium perxenate, Na4XeO6. The latter reaction also produces a small amount of XeO3F2. XeOF4 reacts with CsF to form the XeOF5 anion,[89][91] while XeOF3 reacts with the alkali metal fluorides KF, RbF and CsF to form the XeOF4 anion.[92]

Other compounds

Recently, there has been an interest in xenon compounds where xenon is directly bonded to a less electronegative element than fluorine or oxygen, particularly carbon.[93] Electron-withdrawing groups, such as groups with fluorine substitution, are necessary to stabilize these compounds.[88] Numerous such compounds have been characterized, including:[89][94]

  • C6F5–Xe+–N≡C–CH3, where C6F5 is the pentafluorophenyl group.
  • [C6F5]2Xe
  • C6F5–Xe–X, where X is CN, F, or Cl.
  • R–C≡C–Xe+, where R is C2F5 or tert-butyl.
  • C6F5–XeF+2
  • (C6F5Xe)2Cl+

Other compounds containing xenon bonded to a less electronegative element include F–Xe–N(SO2F)2 and F–Xe–BF2. The latter is synthesized from dioxygenyl tetrafluoroborate, O2BF4, at −100 °C.[89][95]

An unusual ion containing xenon is the tetraxenonogold(II) cation, AuXe2+4, which contains Xe–Au bonds.[96] This ion occurs in the compound AuXe4(Sb2F11)2, and is remarkable in having direct chemical bonds between two notoriously unreactive atoms, xenon and gold, with xenon acting as a transition metal ligand.

In 1995, M. Räsänen and co-workers, scientists at the University of Helsinki in Finland, announced the preparation of xenon dihydride (HXeH), and later xenon hydride-hydroxide (HXeOH), hydroxenoacetylene (HXeCCH), and other Xe-containing molecules.[97] In 2008, Khriachtchev et al. reported the preparation of HXeOXeH by the photolysis of water within a cryogenic xenon matrix.[98] Deuterated molecules, HXeOD and DXeOH, have also been produced.[99]

Clathrates and excimers

In addition to compounds where xenon forms a chemical bond, xenon can form clathrates—substances where xenon atoms are trapped by the crystalline lattice of another compound. An example is xenon hydrate (Xe·5.75 H2O), where xenon atoms occupy vacancies in a lattice of water molecules.[100] This clathrate has a melting point of 24 °C.[101] The deuterated version of this hydrate has also been produced.[102] Such clathrate hydrates can occur naturally under conditions of high pressure, such as in Lake Vostok underneath the Antarctic ice sheet.[103] Clathrate formation can be used to fractionally distill xenon, argon and krypton.[104]

Xenon can also form endohedral fullerene compounds, where a xenon atom is trapped inside a fullerene molecule. The xenon atom trapped in the fullerene can be monitored via 129Xe nuclear magnetic resonance (NMR) spectroscopy. Using this technique, chemical reactions on the fullerene molecule can be analyzed, due to the sensitivity of the chemical shift of the xenon atom to its environment. However, the xenon atom also has an electronic influence on the reactivity of the fullerene.[105]

While xenon atoms are at their ground energy state, they repel each other and will not form a bond. When xenon atoms becomes energized, however, they can form an excimer (excited dimer) until the electrons return to the ground state. This entity is formed because the xenon atom tends to fill its outermost electronic shell, and can briefly do this by adding an electron from a neighboring xenon atom. The typical lifetime of a xenon excimer is 1–5 ns, and the decay releases photons with wavelengths of about 150 and 173 nm.[106][107] Xenon can also form dimers with other elements, such as the halogens bromine, chlorine and fluorine.[108]

Applications

Although xenon is rare and relatively expensive to extract from the Earth's atmosphere, it still has a number of applications.

Illumination and optics

Gas-discharge lamps

Xenon is used in light-emitting devices called xenon flash lamps, which are used in photographic flashes and stroboscopic lamps;[11] to excite the active medium in lasers which then generate coherent light;[109] and, occasionally, in bactericidal lamps.[110] The first solid-state laser, invented in 1960, was pumped by a xenon flash lamp,[15] and lasers used to power inertial confinement fusion are also pumped by xenon flash lamps.[111]

File:Xenon short arc 1.jpg
Xenon short-arc lamp
File:Xenon discharge tube.jpg
Xenon gas discharge tube

Continuous, short-arc, high pressure xenon arc lamps have a color temperature closely approximating noon sunlight and are used in solar simulators. That is, the chromaticity of these lamps closely approximates a heated black body radiator that has a temperature close to that observed from the Sun. After they were first introduced during the 1940s, these lamps began replacing the shorter-lived carbon arc lamps in movie projectors.[12] They are employed in typical 35mm and IMAX film projection systems, automotive HID headlights and other specialized uses. These arc lamps are an excellent source of short wavelength ultraviolet radiation and they have intense emissions in the near infrared, which is used in some night vision systems.

The individual cells in a plasma display use a mixture of xenon and neon that is converted into a plasma using electrodes. The interaction of this plasma with the electrodes generates ultraviolet photons, which then excite the phosphor coating on the front of the display.[112][113]

Xenon is used as a "starter gas" in high pressure sodium lamps. It has the lowest thermal conductivity and lowest ionization potential of all the non-radioactive noble gases. As a noble gas, it does not interfere with the chemical reactions occurring in the operating lamp. The low thermal conductivity minimizes thermal losses in the lamp while in the operating state, and the low ionization potential causes the breakdown voltage of the gas to be relatively low in the cold state, which allows the lamp to be more easily started.[114]

Lasers

In 1962, a group of researchers at Bell Laboratories discovered laser action in xenon,[115] and later found that the laser gain was improved by adding helium to the lasing medium.[116][117] The first excimer laser used a xenon dimer (Xe2) energized by a beam of electrons to produce stimulated emission at an ultraviolet wavelength of 176 nm.[14] Xenon chloride and xenon fluoride have also been used in excimer (or, more accurately, exciplex) lasers.[118] The xenon chloride excimer laser has been employed, for example, in certain dermatological uses.[119]

Medical

Anesthesia

Xenon has been used as a general anaesthetic. Although it is expensive, anesthesia machines that can deliver xenon are about to appear on the European market, because advances in recovery and recycling of xenon have made it economically viable.[49][120]

Two physiological mechanisms for xenon anesthesia have been proposed. The first one involves the inhibition of the calcium ATPase pump—the mechanism cells use to remove calcium (Ca2+)—in the cell membrane of synapses.[121] This results from a conformational change when xenon binds to nonpolar sites inside the protein. [122] The second mechanism focuses on the non-specific interactions between the anesthetic and the lipid membrane.[123]

Xenon has a minimum alveolar concentration (MAC) of 71%, making it 50% more potent than N2O as an anesthetic. Thus it can be used in concentrations with oxygen that have a lower risk of hypoxia. Unlike nitrous oxide (N2O), xenon is not a greenhouse gas and so it is also viewed as environmentally friendly.[124] Xenon vented into the atmosphere is being returned to its original source, so no environmental impact is likely.

Neuroprotectant

Xenon is finding application in treating brain injuries, since it is an antagonist of N-methyl-d-aspartate receptors (NMDA receptors). These receptors exacerbate the damage from oxygen deprivation and xenon performs better as a neuroprotectant than either ketamine or nitrous oxide, which have undesired side-effects.[125] Xenon gas was added as an ingredient of the ventilation mix for a newborn baby at St. Michael's Hospital, Bristol, England, whose life chances were otherwise very compromised, and was successful, leading to the authorisation of clinical trials for similar cases.[126][127]

Imaging

Gamma emission from the radioisotope 133Xe of xenon can be used to image the heart, lungs, and brain, for example, by means of single photon emission computed tomography. 133Xe has also been used to measure blood flow.[128][129][130]

Xenon, particularly hyperpolarized 129Xe, is a useful contrast agent for magnetic resonance imaging (MRI). In the gas phase, it can be used to image empty space such as cavities in a porous sample or alveoli in lungs. Hyperpolarization renders 129Xe much more detectable via magnetic resonance imaging and has been used for studies of the lungs and other tissues. It can be used, for example, to trace the flow of gases within the lungs.[131][132] Because xenon is soluble in water and also in hydrophobic solvents, it can be used to image various soft living tissues.[133][134][135]

NMR spectroscopy

Because of the atom's large, flexible outer electron shell, the NMR spectrum changes in response to surrounding conditions, and can therefore be used as a probe to measure the chemical circumstances around the xenon atom. For instance xenon dissolved in water, xenon dissolved in hydrophobic solvent, and xenon associated with certain proteins can be distinguished by NMR.[136][137]

Hyperpolarized xenon can be used by surface-chemists. Normally, it is difficult to characterize surfaces using NMR, because signals from the surface of a sample will be overwhelmed by signals from the far-more-numerous atomic nuclei in the bulk. However, nuclear spins on solid surfaces can be selectively polarized, by transferrering spin polarization to them from hyperpolarized xenon gas. This makes the surface signals strong enough to measure, and distinguishes them from bulk signals.[138][139]

Other

In nuclear energy applications, xenon is used in bubble chambers,[140] probes, and in other areas where a high molecular weight and inert nature is desirable. A by-product of nuclear weapon testing is the release of radioactive xenon-133 and xenon-135. The detection of these isotopes is used to monitor compliance with test ban treaties,[141] as well as to confirm nuclear test explosions by states such as North Korea.[142]

File:Xenon ion engine prototype.png
A prototype of a xenon ion engine being tested at NASA's Jet Propulsion Laboratory.

Liquid xenon is being used as a medium for detecting hypothetical weakly interacting massive particles, or WIMPs. When a WIMP collides with a xenon nucleus, it should, theoretically, strip an electron and create a primary scintillation. By using xenon, this burst of energy could then be readily distinguished from similar events caused by particles such as cosmic rays.[16] However, the XENON experiment at the Gran Sasso National Laboratory in Italy and the ZEPLIN-II and ZEPLIN-III experiments at the Boulby Underground Laboratory in the UK have thus far failed to find any confirmed WIMPs. Even if no WIMPs are detected, the experiments will serve to constrain the properties of dark matter and some physics models.[143][144] The current detector at the Gran Sasso facility has demonstrated sensitivity comparable to that of the best cryogenic detectors, and the sensitivity was expected to be increased by an order of magnitude in 2009.[145]

Xenon is the preferred fuel for ion propulsion of spacecraft because of its low ionization potential per atomic weight, and its ability to be stored as a liquid at near room temperature (under high pressure) yet be easily converted back into a gas to fuel the engine. The inert nature of xenon makes it environmentally friendly and less corrosive to an ion engine than other fuels such as mercury or caesium. Xenon was first used for satellite ion engines during the 1970s.[146] It was later employed as a propellant for Europe's SMART-1 spacecraft[17] and for the three ion propulsion engines on NASA's Dawn Spacecraft.[147]

Chemically, the perxenate compounds are used as oxidizing agents in analytical chemistry. Xenon difluoride is used as an etchant for silicon, particularly in the production of microelectromechanical systems (MEMS).[148] The anticancer drug 5-fluorouracil can be produced by reacting xenon difluoride with uracil.[149] Xenon is also used in protein crystallography. Applied at pressures from 0.5 to 5 MPa (5 to 50 atm) to a protein crystal, xenon atoms bind in predominantly hydrophobic cavities, often creating a high quality, isomorphous, heavy-atom derivative, which can be used for solving the phase problem.[150][151]

Precautions

Many oxygen-containing xenon compounds are toxic due to their strong oxidative properties, and explosive due to their tendency to break down into elemental xenon plus diatomic oxygen (O2), which contains much stronger chemical bonds than the xenon compounds.[152]

Xenon gas can be safely kept in normal sealed glass or metal containers at standard temperature and pressure. However, it readily dissolves in most plastics and rubber, and will gradually escape from a container sealed with such materials.[153] Xenon is non-toxic, although it does dissolve in blood and belongs to a select group of substances that penetrate the blood-brain barrier, causing mild to full surgical anesthesia when inhaled in high concentrations with oxygen.[152]

At 169 m/s, the speed of sound in xenon gas is slower than that in air[154] due to the slower average speed of the heavy xenon atoms compared to nitrogen and oxygen molecules. Hence, xenon lowers the resonant frequencies of the vocal tract when inhaled. This produces a characteristic lowered voice timbre, an effect opposite to the high-timbred voice caused by inhalation of helium. Like helium, xenon does not satisfy the body's need for oxygen. Xenon is both a simple asphyxiant and an anesthetic more powerful than nitrous oxide; consequently, many universities no longer allow the voice stunt as a general chemistry demonstration. As xenon is expensive, the gas sulfur hexafluoride, which is similar to xenon in molecular weight (146 versus 131), is generally used in this stunt, and is an asphyxiant without being anesthetic.[155]

It is possible to safely breathe heavy gases such as xenon or sulfur hexafluoride when they include a 20% mixture of oxygen. Although xenon at this concentration would be expected to rapidly produce the unconsciousness of general anesthesia, the lungs mix the gases very effectively and rapidly so that the heavy gases are purged along with the oxygen and do not accumulate at the bottom of the lungs.[156] There is, however, a danger associated with any heavy gas in large quantities: it may sit invisibly in a container, and if a person enters a container filled with an odorless, colorless gas, they may find themselves breathing it unknowingly. Xenon is rarely used in large enough quantities for this to be a concern, though the potential for danger exists any time a tank or container of xenon is kept in an unventilated space.[157]

See also

References

Cite error: Invalid <references> tag; parameter "group" is allowed only.

Use <references />, or <references group="..." />

External links

af:Xenon ar:زينون az:Ksenon bn:জেনন be:Ксенон be-x-old:Ксэнон bs:Ksenon bg:Ксенон ca:Xenó cv:Ксенон cs:Xenon co:Xenu cy:Senon da:Xenon de:Xenon et:Ksenoon el:Ξένο es:Xenón eo:Ksenono eu:Xenon fa:زنون fr:Xénon fur:Xenon ga:Xeanón gv:Xenon gl:Xenon xal:Ксенөн ko:제논 (원소) hy:Քսենոն hi:ज़ेनान hr:Ksenon io:Xenono id:Xenon is:Xenon it:Xeno he:קסנון kn:ಝೆನಾನ್ sw:Xenoni la:Xenon lv:Ksenons lb:Xenon lt:Ksenonas lij:Seno li:Xenon jbo:fangynavni hu:Xenon ml:സെനൊൺ mr:झेनॉन ms:Xenon nl:Xenon ja:キセノン no:Xenon nn:Xenon oc:Xenon uz:Ksenon nds:Xenon pl:Ksenon pt:Xenônio ro:Xenon qu:Senun ru:Ксенон stq:Xenon scn:Xenu simple:Xenon sk:Xenón sl:Ksenon sr:Ксенон sh:Ksenon fi:Ksenon sv:Xenon tl:Henon (elemento) ta:செனான் th:ซีนอน tr:Ksenon uk:Ксенон ur:زینون ug:كسېنون vi:Xenon war:Xenon yo:Xenon zh-yue:氙

zh:氙
  1. J. A. Simpson & E. S. C. Weiner, ed (1989). "Xenon". Oxford English Dictionary. 20 (2nd ed.). Oxford: Clarendon Press. ISBN 0-19-861232-X. 
  2. "Xenon". Dictionary.com Unabridged. 2010. Retrieved 2010-05-06. 
  3. Lide, David R. (2004). "Section 4, Properties of the Elements and Inorganic Compounds; Melting, boiling, triple, and critical temperatures of the elements". CRC Handbook of Chemistry and Physics (85th edition ed.). Boca Raton, Florida: CRC Press. ISBN 0849304857.  line feed character in |chapter= at position 72 (help)
  4. Magnetic susceptibility of the elements and inorganic compounds, in Handbook of Chemistry and Physics 81st edition, CRC press.
  5. Staff (2007). "Xenon". Columbia Electronic Encyclopedia (6th edition ed.). Columbia University Press. Retrieved 2007-10-23. 
  6. 6.0 6.1 Husted, Robert; Boorman, Mollie (December 15, 2003). "Xenon". Los Alamos National Laboratory, Chemical Division. Retrieved 2007-09-26. 
  7. Lua error in package.lua at line 80: module 'Module:Citation/CS1/Suggestions' not found.—National Standard Reference Data Service of the USSR. Volume 10.
  8. 8.0 8.1 Freemantel, Michael (August 25, 2003). "Chemistry at its Most Beautiful" (PDF). Chemical & Engineering News. Retrieved 2007-09-13. 
  9. 9.0 9.1 Kaneoka, Ichiro (1998). "Xenon's Inside Story". Science. 280 (5365): 851–852. doi:10.1126/science.280.5365.851b. 
  10. 10.0 10.1 Stacey, Weston M. (2007). Nuclear Reactor Physics. Wiley-VCH. p. 213. ISBN 3527406794. 
  11. 11.0 11.1 11.2 Burke, James (2003). Twin Tracks: The Unexpected Origins of the Modern World. Oxford University Press. p. 33. ISBN 0743226194. 
  12. 12.0 12.1 Mellor, David (2000). Sound Person's Guide to Video. Focal Press. p. 186. ISBN 0240515951. 
  13. Sanders, Robert D.; Ma, Daqing; Maze, Mervyn (2005). "Xenon: elemental anaesthesia in clinical practice". British Medical Bulletin. 71 (1): 115–135. doi:10.1093/bmb/ldh034. PMID 15728132. 
  14. 14.0 14.1 Lua error in package.lua at line 80: module 'Module:Citation/CS1/Suggestions' not found.
  15. 15.0 15.1 Lua error in package.lua at line 80: module 'Module:Citation/CS1/Suggestions' not found.
  16. 16.0 16.1 Ball, Philip (May 1, 2002). "Xenon outs WIMPs". Nature. Retrieved 2007-10-08. 
  17. 17.0 17.1 Lua error in package.lua at line 80: module 'Module:Citation/CS1/Suggestions' not found.
  18. Ramsay, W.; Travers, M. W. (1898). "On the extraction from air of the companions of argon, and neon". Report of the Meeting of the British Association for the Advancement of Science: 828. 
  19. Gagnon, Steve. "It's Elemental - Xenon". Thomas Jefferson National Accelerator Facility. Retrieved 2007-06-16. 
  20. Anonymous (1904). Daniel Coit Gilman, Harry Thurston Peck, Frank Moore Colby, ed. The New International Encyclopædia. Dodd, Mead and Company. p. 906. 
  21. Staff (1991). The Merriam-Webster New Book of Word Histories. Merriam-Webster, Inc. p. 513. ISBN 0877796033. 
  22. Ramsay, William (1902). "An Attempt to Estimate the Relative Amounts of Krypton and of Xenon in Atmospheric Air". Proceedings of the Royal Society of London. 71: 421–426. doi:10.1098/rspl.1902.0121. 
  23. Anonymous. "History". Millisecond Cinematography. Retrieved 2007-11-07. 
  24. Paschotta, Rüdiger (November 1, 2007). "Lamp-pumped lasers". Encyclopedia of Laser Physics and Technology. RP Photonics. Retrieved 2007-11-07. 
  25. Marx, Thomas; Schmidt, Michael; Schirmer, Uwe; Reinelt, Helmut (2000). "Xenon anesthesia" (PDF). Journal of the Royal Society of Medicine. 93 (10): 513–517. PMC 1298124Freely accessible. PMID 11064688. Retrieved 2007-10-02. 
  26. Bartlett, Neil; Lohmann, D. H. (1962). "Dioxygenyl hexafluoroplatinate (V), O2+[PtF6]". Proceedings of the Chemical Society. London: Chemical Society (3): 115. doi:10.1039/PS9620000097. 
  27. Bartlett, N. (1962). "Xenon hexafluoroplatinate (V) Xe+[PtF6]". Proceedings of the Chemical Society. London: Chemical Society (6): 218. doi:10.1039/PS9620000197. 
  28. Lua error in package.lua at line 80: module 'Module:Citation/CS1/Suggestions' not found.
  29. Lua error in package.lua at line 80: module 'Module:Citation/CS1/Suggestions' not found.; translation of Lehrbuch der Anorganischen Chemie, originally founded by A. F. Holleman, continued by Egon Wiberg, edited by Nils Wiberg, Berlin: de Gruyter, 1995, 34th edition, ISBN 3-11-012641-9.
  30. Steel, Joanna (2007). "Biography of Neil Bartlett". College of Chemistry, University of California, Berkeley. Retrieved 2007-10-25. 
  31. Bartlett, Neil (2003-09-09). "The Noble Gases". Chemical & Engineering News. American Chemical Society. 81 (36). Retrieved 2007-10-01. 
  32. Lua error in package.lua at line 80: module 'Module:Citation/CS1/Suggestions' not found.
  33. Lynch, C. T.; Summitt, R.; Sliker, A. (1980). CRC Handbook of Materials Science. CRC Press. ISBN 087819231X. 
  34. MacKenzie, D. R. (1963). "Krypton Difluoride: Preparation and Handling". Science. 141 (3586): 1171. doi:10.1126/science.141.3586.1171. PMID 17751791. 
  35. Paul R. Fields, Lawrence Stein, and Moshe H. Zirin (1962). "Radon Fluoride". Journal of the American Chemical Society. 84 (21): 4164–4165. doi:10.1021/ja00880a048. 
  36. "Xenon". Periodic Table Online. CRC Press. Archived from the original on April 10, 2007. Retrieved 2007-10-08. 
  37. Moody, G. J. (1974). "A Decade of Xenon Chemistry". Journal of Chemical Education. 51: 628–630. doi:10.1021/ed051p628. Retrieved 2007-10-16. 
  38. Williams, David R. (April 19, 2007). "Earth Fact Sheet". NASA. Retrieved 2007-10-04. 
  39. 39.0 39.1 Lua error in package.lua at line 80: module 'Module:Citation/CS1/Suggestions' not found.
  40. Lua error in package.lua at line 80: module 'Module:Citation/CS1/Suggestions' not found.
  41. Fontes, E. "Golden Anniversary for Founder of High-pressure Program at CHESS". Cornell University. Retrieved 2009-05-30. 
  42. Eremets, Mikhail I.; Gregoryanz, Eugene A.; Struzhkin, Victor V.; Mao, Ho-Kwang; Hemley, Russell J.; Mulders, Norbert; Zimmerman, Neil M. (2000). "Electrical Conductivity of Xenon at Megabar Pressures". Physical Review Letters. 85 (13): 2797–2800. doi:10.1103/PhysRevLett.85.2797. PMID 10991236. 
  43. Bader, Richard F. W. "An Introduction to the Electronic Structure of Atoms and Molecules". McMaster University. Retrieved 2007-09-27. 
  44. Talbot, John. "Spectra of Gas Discharges". Rheinisch-Westfälische Technische Hochschule Aachen. Retrieved 2006-08-10. 
  45. Watts, William Marshall (1904). An Introduction to the Study of Spectrum Analysis. London: Longmans, Green, and co. 
  46. Lua error in package.lua at line 80: module 'Module:Citation/CS1/Suggestions' not found.
  47. Kerry, Frank G. (2007). Industrial Gas Handbook: Gas Separation and Purification. CRC Press. pp. 101–103. ISBN 0849390052. 
  48. "Xenon - Xe". CFC StarTec LLC. August 10, 1998. Retrieved 2007-09-07. 
  49. 49.0 49.1 Singh, Sanjay (May 15, 2005). "Xenon: A modern anaesthetic". Indian Express Newspapers Limited. Retrieved 2007-10-10. 
  50. 50.0 50.1 Lua error in package.lua at line 80: module 'Module:Citation/CS1/Suggestions' not found.
  51. Arnett, David (1996). Supernovae and Nucleosynthesis. Princeton, New Jersey: Princeton University Press. ISBN 0-691-01147-8. 
  52. Lua error in package.lua at line 80: module 'Module:Citation/CS1/Suggestions' not found.
  53. Lua error in package.lua at line 80: module 'Module:Citation/CS1/Suggestions' not found.
  54. Lua error in package.lua at line 80: module 'Module:Citation/CS1/Suggestions' not found.
  55. Clayton, Donald D. (1983). Principles of Stellar Evolution and Nucleosynthesis. University of Chicago Press. ISBN 0226109534. 
  56. Heymann, D.; Dziczkaniec, M. (March 19–23, 1979). "Xenon from intermediate zones of supernovae". Proceedings 10th Lunar and Planetary Science Conference. Houston, Texas: Pergamon Press, Inc.. pp. 1943–1959. http://adsabs.harvard.edu/abs/1979LPSC...10.1943H. Retrieved 2007-10-02. 
  57. Lua error in package.lua at line 80: module 'Module:Citation/CS1/Suggestions' not found.
  58. Lua error in package.lua at line 80: module 'Module:Citation/CS1/Suggestions' not found.
  59. 59.0 59.1 59.2 Lua error in package.lua at line 80: module 'Module:Citation/CS1/Suggestions' not found.
  60. Rajam, J. B. (1960). Atomic Physics (7th ed.). Delhi: S. Chand and Co. ISBN 812191809X. 
  61. Barabash, A. S. (2002). "Average (Recommended) Half-Life Values for Two-Neutrino Double-Beta Decay". Czechoslovak Journal of Physics. 52 (4): 567–573. doi:10.1023/A:1015369612904. 
  62. Otten, Ernst W. (2004). "Take a breath of polarized noble gas". Europhysics News. 35 (1): 16. doi:10.1051/epn:2004109. 
  63. Lua error in package.lua at line 80: module 'Module:Citation/CS1/Suggestions' not found.
  64. Lua error in package.lua at line 80: module 'Module:Citation/CS1/Suggestions' not found.
  65. Lua error in package.lua at line 80: module 'Module:Citation/CS1/Suggestions' not found.
  66. von Schulthess, Gustav Konrad; Smith, Hans-Jørgen; Pettersson, Holger; Allison, David John (1998). "The Encyclopaedia of Medical Imaging". The Encyclopaedia of Medical Imaging. Taylor & Francis. p. 194. ISBN 1901865134. http://books.google.com/books?id=zvDY5unRC4oC&pg=PA194. 
  67. Lua error in package.lua at line 80: module 'Module:Citation/CS1/Suggestions' not found.
  68. Staff. "Hanford Becomes Operational". The Manhattan Project: An Interactive History. U.S. Department of Energy. Retrieved 2007-10-10. 
  69. Lua error in package.lua at line 80: module 'Module:Citation/CS1/Suggestions' not found.
  70. Laws, Edwards A. (2000). Aquatic Pollution: An Introductory Text. John Wiley and Sons. p. 505. ISBN 0471348759. 
  71. Staff (April 9, 1979). "A Nuclear Nightmare". Time. Retrieved 2007-10-09. 
  72. Clayton, Donald D. (1983). Principles of Stellar Evolution and Nucleosynthesis (2nd ed.). University of Chicago Press. p. 75. ISBN 0226109534. 
  73. Bolt, B. A.; Packard, R. E.; Price, P. B. (2007). "John H. Reynolds, Physics: Berkeley". The University of California, Berkeley. Retrieved 2007-10-01. 
  74. Williams, David R. (September 1, 2004). "Mars Fact Sheet". NASA. Retrieved 2007-10-10. 
  75. Schilling, James. "Why is the Martian atmosphere so thin and mainly carbon dioxide?". Mars Global Circulation Model Group. Retrieved 2007-10-10. 
  76. Zahnle, Kevin J. (1993). "Xenological constraints on the impact erosion of the early Martian atmosphere". Journal of Geophysical Research. 98 (E6): 10,899–10,913. doi:10.1029/92JE02941. 
  77. Lua error in package.lua at line 80: module 'Module:Citation/CS1/Suggestions' not found.
  78. 78.0 78.1 Harding, Charlie; Johnson, David Arthur; Janes, Rob (2002). Elements of the p block. Great Britain: Royal Society of Chemistry. pp. 93–94. ISBN 0854046909. 
  79. Weeks, James L.; Chernick, Cedric; Matheson, Max S. (1962). "Photochemical Preparation of Xenon Difluoride". Journal of the American Chemical Society. 84: 4612. doi:10.1021/ja00882a063. 
  80. Streng, L. V.; Streng, A. G. (1965). "Formation of Xenon Difluoride from Xenon and Oxygen Difluoride or Fluorine in Pyrex Glass at Room Temperature". Inorganic Chemistry. 4 (9): 1370–1371. doi:10.1021/ic50031a035. 
  81. 81.0 81.1 Tramšek, Melita; Žemva, Boris (December 5, 2006). "Synthesis, Properties and Chemistry of Xenon(II) Fluoride" (PDF). Acta Chimica Slovenica. 53 (2): 105–116. doi:10.1002/chin.200721209. Retrieved 2009-07-18. 
  82. Ogrin, Tomaz; Bohinc, Matej; Silvnik, Joze (1973). "Melting-point determinations of xenon difluoride-xenon tetrafluoride mixtures". Journal of Chemical and Engineering Data. 18 (4): 402. doi:10.1021/je60059a014. 
  83. 83.0 83.1 Scott, Thomas; Eagleson, Mary (1994). "Xenon Compounds". Concise encyclopedia chemistry. Walter de Gruyter. p. 1183. ISBN 3110114518. http://books.google.com/books?id=Owuv-c9L_IMC&pg=PA1183. 
  84. Proserpio, Davide M.; Hoffmann, Roald; Janda, Kenneth C. (1991). "The xenon-chlorine conundrum: van der Waals complex or linear molecule?". Journal of the American Chemical Society. 113: 7184. doi:10.1021/ja00019a014. 
  85. Richardson, Nancy A.; Hall, Michael B. (1993). "The potential energy surface of xenon dichloride". The Journal of Physical Chemistry. 97: 10952. doi:10.1021/j100144a009. 
  86. Zhou, M.; Zhao, Y.; Gong, Y.; Li, J. (2006). "Formation and Characterization of the XeOO+ Cation in Solid Argon". Journal of the American Chemical Society. 128 (8): 2504–2505. doi:10.1021/ja055650n. PMID 16492012. 
  87. Lua error in package.lua at line 80: module 'Module:Citation/CS1/Suggestions' not found.
  88. 88.0 88.1 Henderson, W. (2000). Main group chemistry. Great Britain: Royal Society of Chemistry. pp. 152–153. ISBN 0854046178. 
  89. 89.0 89.1 89.2 89.3 Mackay, Kenneth Malcolm; Mackay, Rosemary Ann; Henderson, W. (2002). Introduction to modern inorganic chemistry (6th ed.). CRC Press. pp. 497–501. ISBN 0748764208. 
  90. Smith, D. F. (1963). "Xenon Oxyfluoride". Science. 140 (3569): 899. doi:10.1126/science.140.3569.899. PMID 17810680. 
  91. K. O. Christe, D. A. Dixon, J. C. P. Sanders, G. J. Schrobilgen, S. S. Tsai, W. W. Wilson (1995). "On the Structure of the [XeOF5] Anion and of Heptacoordinated Complex Fluorides Containing One or Two Highly Repulsive Ligands or Sterically Active Free Valence Electron Pairs". Inorg. Chem. 34 (7): 1868–1874. doi:10.1021/ic00111a039. 
  92. K. O. Christe, C. J. Schack, D. Pilipovich (1972). "Chlorine trifluoride oxide. V. Complex formation with Lewis acids and bases". Inorg. Chem. 11 (9): 2205–2208. doi:10.1021/ic50115a044. 
  93. Holloway, John H.; Hope, Eric G. (1998). Advances in Inorganic Chemistry. Contributor A. G. Sykes. Academic Press. pp. 61–90. ISBN 012023646X. 
  94. Frohn, H (2004). "C6F5XeF, a versatile starting material in xenon–carbon chemistry". Journal of Fluorine Chemistry. 125: 981. doi:10.1016/j.jfluchem.2004.01.019. 
  95. Lua error in package.lua at line 80: module 'Module:Citation/CS1/Suggestions' not found.
  96. Li, Wai-Kee; Zhou, Gong-Du; Mak, Thomas C. W. (2008). Gong-Du Zhou; Thomas C. W. Mak, eds. Advanced Structural Inorganic Chemistry. Oxford University Press. p. 678. ISBN 0199216940. 
  97. Gerber, R. B. (2004). "Formation of novel rare-gas molecules in low-temperature matrices". Annual Review of Physical Chemistry. 55: 55–78. doi:10.1146/annurev.physchem.55.091602.094420. PMID 15117247. 
  98. Lua error in package.lua at line 80: module 'Module:Citation/CS1/Suggestions' not found.
  99. Lua error in package.lua at line 80: module 'Module:Citation/CS1/Suggestions' not found.
  100. Pauling, L. (1961). "A molecular theory of general anesthesia". Science. 134: 15–21. doi:10.1126/science.134.3471.15. PMID 13733483.  Reprinted as Pauling, Linus; Kamb, Barclay, ed. (2001). Linus Pauling: Selected Scientific Papers. 2. River Edge, New Jersey: World Scientific. pp. 1328–1334. ISBN 9810229402. 
  101. Henderson, W. (2000). Main group chemistry. Great Britain: Royal Society of Chemistry. p. 148. ISBN 0854046178. 
  102. Lua error in package.lua at line 80: module 'Module:Citation/CS1/Suggestions' not found.
  103. Lua error in package.lua at line 80: module 'Module:Citation/CS1/Suggestions' not found.
  104. Lua error in package.lua at line 80: module 'Module:Citation/CS1/Suggestions' not found.
  105. Lua error in package.lua at line 80: module 'Module:Citation/CS1/Suggestions' not found.
  106. Silfvast, William Thomas (2004). Laser Fundamentals. Cambridge University Press. ISBN 0521833450. 
  107. Webster, John G. (1998). The Measurement, Instrumentation, and Sensors Handbook. Springer. ISBN 3540648305. 
  108. Lua error in package.lua at line 80: module 'Module:Citation/CS1/Suggestions' not found.
  109. Staff (2007). "Xenon Applications". Praxair Technology. Retrieved 2007-10-04. 
  110. Lua error in package.lua at line 80: module 'Module:Citation/CS1/Suggestions' not found.
  111. Lua error in package.lua at line 80: module 'Module:Citation/CS1/Suggestions' not found.
  112. Anonymous. "The plasma behind the plasma TV screen". Plasma TV Science. Retrieved 2007-10-14. 
  113. Marin, Rick (March 21, 2001). "Plasma TV: That New Object Of Desire". The New York Times. Retrieved 2009-04-03. 
  114. Waymouth, John (1971). Electric Discharge Lamps. Cambridge, MA: The M.I.T. Press. ISBN 0262230488. 
  115. Lua error in package.lua at line 80: module 'Module:Citation/CS1/Suggestions' not found.
  116. Lua error in package.lua at line 80: module 'Module:Citation/CS1/Suggestions' not found.
  117. Bennett, Jr., W. R. (1962). "Gaseous optical masers". Applied Optics Supplement. 1: 24–61. 
  118. "Laser Output". University of Waterloo. Retrieved 2007-10-07. 
  119. Lua error in package.lua at line 80: module 'Module:Citation/CS1/Suggestions' not found.
  120. Tonner, P. H. (2006). "Xenon: one small step for anaesthesia…? (editorial review)". Current Opinion in Anaesthesiology. 19 (4): 382–384. doi:10.1097/01.aco.0000236136.85356.13. PMID 16829718. 
  121. Lua error in package.lua at line 80: module 'Module:Citation/CS1/Suggestions' not found.
  122. Lua error in package.lua at line 80: module 'Module:Citation/CS1/Suggestions' not found.
  123. Lua error in package.lua at line 80: module 'Module:Citation/CS1/Suggestions' not found.
  124. Lua error in package.lua at line 80: module 'Module:Citation/CS1/Suggestions' not found.
  125. Ma, D.; Wilhelm, S.; Maze, M.; Franks, N.P. (2002). "Neuroprotective and neurotoxic properties of the 'inert' gas, xenon". British Journal of Anaesthesia. 89 (5): 739–746. doi:10.1093/bja/aef258. PMID 12393773. 
  126. "First baby given xenon gas to prevent brain injury". BBC News. April 9, 2010. Retrieved 2010-04-09. 
  127. "First baby given xenon gas to prevent brain injury". BBC. 9 April 2010. 
  128. Van Der Wall, Ernst (1992). What's New in Cardiac Imaging?: SPECT, PET, and MRI. Springer. ISBN 0792316150. 
  129. Frank, John (1999). "Introduction to imaging: The chest". Student BMJ. 12: 1–44. Retrieved 2008-06-04. 
  130. Chandak, Puneet K. (July 20, 1995). "Brain SPECT: Xenon-133". Brigham RAD. Retrieved 2008-06-04. 
  131. Lua error in package.lua at line 80: module 'Module:Citation/CS1/Suggestions' not found.
  132. Irion, Robert (March 23, 1999). "Head Full of Xenon?". Science News. Archived from the original on January 17, 2004. Retrieved 2007-10-08. 
  133. Lua error in package.lua at line 80: module 'Module:Citation/CS1/Suggestions' not found.
  134. Lua error in package.lua at line 80: module 'Module:Citation/CS1/Suggestions' not found.
  135. Cleveland, Z.I.; Möller, H.E.; Hedlund, L.W.; Driehuys, B. (2009). "Continuously infusing hyperpolarized 129Xe into flowing aqueous solutions using hydrophobic gas exchange membranes". The journal of physical chemistry. 113 (37): 12489–99. doi:10.1021/jp9049582. PMC 2747043Freely accessible. PMID 19702286. 
  136. Lua error in package.lua at line 80: module 'Module:Citation/CS1/Suggestions' not found.
  137. Rubin, Seth M.; Spence, Megan M.; Goodson, Boyd M.; Wemmer, David E.; Pines, Alexander (August 15, 2000). "Evidence of nonspecific surface interactions between laser-polarized xenon and myoglobin in solution". Proceedings of the National Academy of Science USA. 97 (17): 9472–9475. doi:10.1073/pnas.170278897. PMC 16888Freely accessible. PMID 10931956. 
  138. Raftery, Daniel; MacNamara, Ernesto; Fisher, Gregory; Rice, Charles V.; Smith, Jay (1997). "Optical Pumping and Magic Angle Spinning: Sensitivity and Resolution Enhancement for Surface NMR Obtained with Laser-Polarized Xenon". Journal of the American Chemical Society. 119: 8746. doi:10.1021/ja972035d. 
  139. Gaede, H. C.; Song, Y. -Q.; Taylor, R. E.; Munson, E. J.; Reimer, J. A.; Pines, A. (1995). "High-field cross polarization NMR from laser-polarized xenon to surface nuclei". Applied Magnetic Resonance. 8: 373. doi:10.1007/BF03162652. 
  140. Galison, Peter Louis (1997). Image and Logic: A Material Culture of Microphysics. University of Chicago Press. p. 339. ISBN 0226279170. 
  141. Fontaine, J.-P.; Pointurier, F.; Blanchard, X.; Taffary, T. (2004). "Atmospheric xenon radioactive isotope monitoring". Journal of Environmental Radioactivity. 72 (1–2): 129–135. doi:10.1016/S0265-931X(03)00194-2. PMID 15162864. 
  142. Lua error in package.lua at line 80: module 'Module:Citation/CS1/Suggestions' not found.
  143. Schumann, Marc (October 10, 2007). "XENON announced new best limits on Dark Matter". Rice University. Retrieved 2007-10-08. 
  144. Lua error in package.lua at line 80: module 'Module:Citation/CS1/Suggestions' not found.
  145. Boyd, Jade (August 23, 2007). "Rice physicists go deep for 'dark matter'". Hubble News Desk. Retrieved 2007-10-08. 
  146. Zona, Kathleen (March 17, 2006). "Innovative Engines: Glenn Ion Propulsion Research Tames the Challenges of 21st century Space Travel". NASA. Retrieved 2007-10-04. 
  147. "Dawn Launch: Mission to Vesta and Ceres" (PDF). NASA. Retrieved 2007-10-01. 
  148. Brazzle, J. D.; Dokmeci, M. R.; Mastrangelo, C. H. (July 28-August 1, 1975). "Modeling and Characterization of Sacrificial Polysilicon Etching Using Vapor-Phase Xenon Difluoride". Proceedings 17th IEEE International Conference on Micro Electro Mechanical Systems (MEMS). Maastricht, Netherlands: IEEE. pp. 737–740. ISBN 9780780382657. 
  149. Staff (2007). "Powerful tool". American Chemical Society. Retrieved 2007-10-10. 
  150. Staff (December 21, 2004). "Protein Crystallography: Xenon and Krypton Derivatives for Phasing". Daresbury Laboratory, PX. Retrieved 2007-10-01. 
  151. Lua error in package.lua at line 80: module 'Module:Citation/CS1/Suggestions' not found.
  152. 152.0 152.1 Lua error in package.lua at line 80: module 'Module:Citation/CS1/Suggestions' not found.
  153. Lua error in package.lua at line 80: module 'Module:Citation/CS1/Suggestions' not found.
  154. 169.44 m/s in xenon (at 0°C and 107 KPa), compared to 344 m/s in air. See: Lua error in package.lua at line 80: module 'Module:Citation/CS1/Suggestions' not found.
  155. Spangler, Steve (2007). "Anti-Helium - Sulfur Hexafluoride". Steve Spangler Science. Retrieved 2007-10-04. 
  156. Lua error in package.lua at line 80: module 'Module:Citation/CS1/Suggestions' not found.
  157. Staff (August 1, 2007). "Cryogenic and Oxygen Deficiency Hazard Safety". Stanford Linear Accelerator Center. Archived from the original on June 9, 2007. Retrieved 2007-10-10.