Krypton is a chemical element with the symbolKr and atomic number36, classified as a noble gas in group 18 of the periodic table.[1] It is a colorless, odorless, and tasteless gas at standard temperature and pressure, occurring in trace amounts in Earth's atmosphere at approximately 1 part per million by volume.[2] Discovered in 1898 by British chemists Sir William Ramsay and Morris William Travers through the fractional distillation of liquefied air, krypton derives its name from the Greek word kryptos, meaning "hidden," reflecting its rarity and elusive nature.[2]As one of the six noble gases—alongside helium, neon, argon, xenon, and radon—krypton exhibits low chemical reactivity due to its full outer electron shell, though it can form compounds like krypton difluoride (KrF₂) under specific conditions.[3] Its physical properties include a boiling point of -153.4°C and a melting point of -157.4°C, making it denser than air with a density of about 3.75 grams per liter at standard conditions.[1] Krypton has six stable isotopes, with krypton-84 being the most abundant at 57%, and it emits brilliant green and orange spectral lines when electrically excited, which historically aided in its identification.[3]Commercially, krypton is obtained as a byproduct of air separation processes and finds applications in specialized lighting, such as high-intensity discharge lamps and photographic flash tubes, due to its high light output efficiency.[2] It is also used in excimer lasers for medical procedures like eye surgery and in ion thrusters for spacecraft propulsion.[2] Despite its inertness, krypton poses no significant toxicity, though the radioactive isotope krypton-85, produced in nuclear reactors, requires monitoring for environmental release.[1] Overall, krypton's rarity—estimated at only 0.0001% of the atmosphere—limits its abundance, yet its unique properties make it valuable in scientific and industrial contexts.[3]
History
Discovery
Krypton was discovered in 1898 as part of the ongoing exploration of noble gases, which began with argon's identification in 1894 by Lord Rayleigh and William Ramsay, followed by helium's confirmation on Earth in 1895 by Ramsay. In 1898, Ramsay and Morris Travers went on to discover krypton in May, followed by neon and xenon later that year.[4] These discoveries highlighted the existence of a family of inert gases in the atmosphere, prompting further investigation into potential heavier members to fill gaps in the periodic table.[5]Ramsay and Travers isolated krypton through fractional distillation of liquefied air, a method that allowed them to separate components based on differing boiling points after removing nitrogen, oxygen, argon, and other known gases.[6] By evaporating the residue from this process, they obtained a small quantity of a new gas, which they estimated to be present in air at about 1 part per million.[7] This isolation occurred on May 30, 1898, at University College London, marking krypton as the heaviest noble gas identified at the time after argon.[8]The presence of this new element was confirmed spectroscopically when the gas was excited in a discharge tube, revealing distinct emission lines in the green and yellow regions of the visible spectrum that did not match any known substance.[9] Ramsay and Travers announced their findings in a paper presented to the Royal Society on June 9, 1898, describing the gas's unique spectral signature as evidence of a novel constituent of atmospheric air. They named it krypton, derived from the Greek word kryptos meaning "hidden," reflecting its rarity and elusive detection despite being a permanent component of the atmosphere.[6]
Etymology and Naming
The name krypton derives from the Greek word kryptos (κρυπτός), meaning "hidden," a reference to the element's extreme rarity in Earth's atmosphere and the challenges involved in its detection and isolation.[10] This etymology underscores the elusive nature of the gas, which constitutes only about 1 part per million of air.[2]The name was proposed by British chemists Sir William Ramsay and Morris William Travers shortly after their discovery of the element on May 30, 1898, through the fractional distillation of liquid air residues remaining after argon extraction.[11] They first announced the name in their preliminary account presented to the Royal Society on June 9, 1898, and it was formally accepted without controversy by the scientific community, aligning with the emerging convention for naming noble gases using Greek roots ending in "-on" to denote their inert, "otherworldly" character—such as argon ("lazy") and neon ("new").[12][13]Krypton's placement as element 36 in the modern periodic table reflects its atomic number, determined by the number of protons in its nucleus, a concept solidified by Henry Moseley's work in 1913; this position filled a gap in the zero group foreseen by Dmitri Mendeleev's periodic law after the discovery of argon prompted revisions to include inert gases.[14] While Mendeleev did not explicitly predict krypton, the periodic system's structure anticipated elements in this sequence, with krypton's properties matching the expected trends for the family.[15]Minor discussions in the late 19th century centered on standardizing nomenclature for the noble gases as a class, with Ramsay advocating Greek-derived names to emphasize their chemical inertness and distinction from reactive elements, a practice that influenced subsequent namings like xenon ("stranger") in 1898.[16] No significant controversies arose specifically over krypton's name, as it was ratified swiftly by bodies like the International Union of Pure and Applied Chemistry's predecessors.
Physical Properties
Atomic and Electronic Structure
Krypton is a chemical element with atomic number 36 and chemical symbol Kr, positioned in group 18 (the noble gases) and period 4 of the periodic table.[1] As a member of the noble gas group, it possesses a stable electronic structure that accounts for its low reactivity under standard conditions.[2]The ground-state electron configuration of krypton is [ \ce{Ar} ] 4s^2 3d^{10} 4p^6, where the core electrons are those of argon (Z = 18), followed by filled 4s and 3d subshells, and a completely occupied 4p subshell with six electrons.[1] This full valence shell octet in the outermost energy level provides krypton with high stability, contributing to the general inertness characteristic of noble gases.[17]Krypton's atomic radii reflect its position among the heavier noble gases. The covalent (atomic) radius is 116 ± 4 pm, while the van der Waals radius, relevant for non-bonded interactions, measures 202 pm. Ionic radii are not typically reported for neutral krypton due to its reluctance to form ions.[18]The ionization energies of krypton underscore the energy required to remove electrons from its tightly bound shells. The first ionization energy, corresponding to the removal of one 4p electron to form Kr⁺, is 1350.8 kJ/mol (or 13.9996 eV).[17] Successive ionization energies increase progressively: the second (to form Kr²⁺) is 2350.4 kJ/mol, the third (to Kr³⁺) is 3565 kJ/mol, and higher ones exceed 5000 kJ/mol, reflecting the greater difficulty in removing electrons from increasingly positively charged ions.Nuclear properties of krypton are dominated by its most common stable isotope, ^{84}Kr, which constitutes about 57% of naturally occurring krypton and features a nucleus with 36 protons and 48 neutrons.[1] This isotope's configuration contributes to the element's average atomic mass of 83.798 u.[19]
Thermodynamic and Optical Properties
Krypton exhibits a density of 3.733 g/L at standard temperature and pressure (STP), reflecting its relatively high atomic mass of 83.798 u among the noble gases. As a monatomic ideal gas under ambient conditions, it displays low thermal conductivity of 9.43 × 10^{-3} W/(m·K) and a specific heat capacity at constant pressure (C_p) of approximately 20.8 J/(mol·K), consistent with the theoretical value of (5/2)R for monatomic gases.[20]The element transitions through distinct phases at cryogenic temperatures: its triple point occurs at 115.78 K and 73.2 kPa, marking the coexistence of solid, liquid, and gas phases. The melting point of solid krypton is -157.37 °C (115.78 K), while the normal boiling point of the liquid is -153.415 °C (119.735 K); above the critical point of 209.48 K and 5.50 MPa, krypton exists as a supercritical fluid with no distinct liquid-gas boundary.[21] Liquid krypton has a density of about 2.413 g/cm³ near its boiling point, and the solid adopts a face-centered cubic lattice structure below the triple point temperature.[22]Krypton's solubility in water is low, at 224 mg/L (or 0.060 cm³/cm³) under 1 atmpartial pressure at 20 °C, decreasing with increasing temperature due to its nonpolar nature.[23]Optically, krypton gas has a refractive index of 1.000427 at visible wavelengths and STP, slightly higher than air due to its electronic polarizability.[24] Its atomic emission spectrum features prominent lines, including a strong green emission at 557.0 nm (from the 5p[3/2]_2 to 5s[3/2]_1^o transition in Kr I), which is utilized in spectroscopic calibration and lighting applications.
Chemical Properties
Reactivity and Bonding
Krypton exhibits low chemical reactivity characteristic of noble gases, primarily due to its electron configuration [Ar] 3d¹⁰ 4s² 4p⁶, which provides a stable octet in the valence shell, and its high first ionization energy of 1350.8 kJ/mol.[25] This full outer shell minimizes the tendency to gain, lose, or share electrons, rendering krypton inert under standard conditions and resistant to forming chemical bonds with most elements.[26] The high ionization energy reflects the strong binding of valence electrons, further contributing to its chemical stability.[25]Under high-energy excitation, such as in electrical discharges or laser irradiation, krypton can form transient excimers like Kr₂* in the gas phase. These diatomic species are bound by van der Waals forces in the ground state but stabilize through excited electronic states, emitting vacuum ultraviolet light upon relaxation; this process is key to krypton excimer lasers.[27] Such excimers represent a rare exception to krypton's inertness, occurring only under non-equilibrium conditions without forming stable ground-state molecules.Although no stable krypton compounds exist at standard conditions, rare oxidation states such as +2 have been observed in fluorides like KrF₂, first synthesized in 1963 under extreme conditions involving fluorine and electric discharge.[14]Molecular orbital theory explains this limited reactivity: the large energy separation between krypton's filled 4p orbitals and unoccupied 5s orbitals results in weak σ-bonding interactions and low bond dissociation energies, typically on the order of 50 kJ/mol for known species.[28] Compared to other noble gases, krypton shows intermediate reactivity—less than xenon, which benefits from greater polarizability and lower ionization energy (1170 kJ/mol), but more than neon due to its larger atomic radius facilitating better orbital overlap in potential bonds.[26]
Known Compounds
Krypton difluoride (KrF₂) is the most prominent and stable compound of krypton, synthesized by the direct combination of krypton and fluorine gases at low temperatures, typically using UV photolysis or electric discharge to initiate the reaction. The initial preparation in matrix isolation was reported in 1963 by photolyzing a mixture of Kr and F₂ in an argon matrix at 20 K, yielding the compound as a white solid upon warming. Bulk synthesis was achieved shortly thereafter by electron beam irradiation of Kr and F₂ mixtures at −150 °C, producing colorless crystals. KrF₂ adopts a linear molecular structure with Kr–F bond lengths of 1.890 Å, as determined by gas-phase electron diffraction and confirmed by low-temperature X-ray crystallography of the α-phase. It serves as a strong oxidizing agent due to its endothermic formation (ΔH_f = +60 kJ/mol), and its thermal stability is limited, with decomposition to the elements occurring above −10 °C and a rate of about 10% per hour at room temperature.[29][30]Other krypton halides are transient and less stable than KrF₂, observed primarily in matrix isolation or gas-phase experiments. Krypton dichloride (KrCl₂) has been detected as a short-lived species through photolysis of HCl in a krypton matrix at cryogenic temperatures or by Kr-sensitized reactions with Cl₂, exhibiting a bent structure with a Kr–Cl bond length of approximately 2.2 Å, but it decomposes rapidly upon warming above 20 K. Similar transient behavior is noted for KrBr₂ and KrI₂, formed under analogous conditions, with decomposition temperatures below 40 K and lifetimes on the order of seconds in the gas phase. Krypton oxides are even rarer, with no stable neutral species known; however, an unstable krypton peroxide-like intermediate has been proposed in matrix studies, decomposing immediately at temperatures above 50 K. The only verified krypton–oxygen bonded compound is Kr(OTeF₅)₂, prepared in 1989 by reacting KrF₂ with B(OTeF₅)₃ in HF solvent at −45 °C, featuring two nearly linear Kr–O bonds (1.72 Å) and decomposing above −20 °C to KrF₂ and TeF₅O–F.[31][32]Krypton also forms van der Waals compounds and clathrates, which are inclusion complexes rather than covalent bonds. The krypton clathrate hydrate Kr·6H₂O consists of krypton atoms trapped in cages formed by hydrogen-bonded water molecules, synthesized under high pressure (above 1 kbar) and low temperature (below 12 °C), with a cubic structure (space group Pm3n). This compound is stable up to its decomposition temperature of approximately 0 °C at atmospheric pressure but persists to higher temperatures (up to 20 °C) under elevated pressures (e.g., 15 kbar); it dissociates into gaseous Kr and ice Ih upon warming or pressure release. Similar clathrates form with other hosts, such as Kr·(THF)·nH₂O, exhibiting enhanced stability due to the tetrahydrofuran guest.[33]Post-2000 discoveries have expanded krypton's coordination chemistry, particularly cationic species and mixed noble gas compounds. Salts of the KrF⁺ cation, such as [KrF][AsF₆] and [KrF][SbF₆], are prepared by fluoride abstraction from KrF₂ using strong Lewis acids like AsF₅ or SbF₅ in HF solvent at −78 °C, yielding colorless solids with linear KrF⁺ units (Kr–F bond 1.837 Å) that are stable up to −40 °C before decomposing to KrF₂ and F₂. The related [Kr₂F₃]⁺ salts, featuring a bent F–Kr–F–Kr–F chain, are synthesized similarly and decompose above −30 °C. In matrix isolation, the mixed noble gas species HArKrF has been observed since 2000 through photolysis of HF/Ar/Kr mixtures at 8 K, forming a weakly bound complex with H–Ar and Kr–F linkages that persists below 20 K before photodissociation. In 2014, the high-pressure van der Waals compound Kr(H₂)₄ was discovered, consisting of krypton atoms surrounded by hydrogen molecules. In 2021, crystals of mixed krypton-xenon compounds, such as [KrXe][F₅(AsF₆)] and [Xe₂F₃][AsF₆], were prepared, further demonstrating heteronuclear bonding under extreme conditions. These compounds highlight krypton's ability to form higher oxidation state and heteronuclear bonds under extreme conditions.[30][34][35][36]
Occurrence and Production
Natural Occurrence
Krypton occurs naturally in trace amounts primarily in Earth's atmosphere, where it constitutes approximately 1 part per million (ppm) by volume, making it one of the rarer noble gases.[2] This atmospheric presence ranks krypton as the 83rd most abundant element in the Earth's crust, with a concentration of about 0.0001 ppm.[37] The element's primordial origins trace back to stellar nucleosynthesis processes, such as the slow neutron capture (s-process) in asymptotic giant branch stars, contributing to its cosmic abundance of approximately 0.0015 ppm relative to hydrogen in the universe.[38]In addition to the atmosphere, krypton is found in minute quantities within certain minerals and in igneous rocks at levels around 0.0001 ppm.[39] Volcanic gases also serve as a natural source, emanating from mantle plumes like those at Yellowstone, where krypton isotopes reflect deep Earth primordial signatures.[40] The atmospheric krypton is largely a result of primordial fractionation during Earth's formation, with minimal ongoing production from natural processes.The isotopic composition of atmospheric krypton emphasizes its stable isotopes, with ^{84}Kr comprising about 57% of the total, followed by ^{86}Kr at 17.3%, ^{82}Kr at 11.6%, ^{83}Kr at 11.5%, ^{80}Kr at 2.3%, and ^{78}Kr at 0.4%.[41] These ratios provide insights into the element's primordial distribution without significant alteration from decay.Krypton has been detected in meteorites, such as the Martian Chassigny meteorite, where its chondritic isotopic ratios indicate early solarsystem accretion processes.[42] In planetary atmospheres, it appears at about 0.3 ppm in Mars' thin atmosphere, measured by missions like Curiosity.[43] On Jupiter, krypton is enriched approximately 2.7 times relative to solar abundances, as determined by the Galileo probe, highlighting variations in giant planet formation.[44]
Commercial Production
Krypton is commercially produced primarily through fractional distillation of liquefied air in large-scale cryogenic air separation units (ASUs), where it is recovered as a valuable by-product alongside other noble gases like neon, argon, and xenon. Atmospheric air, containing approximately 1 ppm of krypton by volume, serves as the starting material for this process.[45]The production process involves several key steps: atmospheric air is first compressed and precooled, then further cooled to around -196°C using heat exchangers and expansion turbines to achieve liquefaction. The resulting liquid air mixture is introduced into a series of distillation columns operating under controlled pressure and temperature conditions. In these columns, components separate based on differing boiling points—nitrogen distills first at -196°C, followed by oxygen at -183°C and argon at -186°C—leaving a crude krypton-xenon fraction that is further refined in additional columns to isolate krypton, which boils at -153°C. Final purification occurs via selective adsorption or repeated distillation to remove trace impurities such as hydrocarbons and other gases.[5][46]Global annual production of krypton is estimated at around 100 tonnes as of 2024, reflecting its low atmospheric abundance and the limited number of large ASUs capable of economic recovery.[47] High-purity krypton (up to 99.999%) is achieved through these methods, suitable for demanding applications. Additionally, recycling efforts are increasing, particularly from semiconductormanufacturing processes where krypton is used in excimer lasers; recovery systems employ pumps and purification units to reclaim and reuse the gas, reducing dependency on primary production. The cost of high-purity krypton is approximately $20–100 per liter depending on quantity and market conditions, as of 2024, due to the energy-intensive nature of extraction and its scarcity.[48][49][50][51]
Isotopes
Stable Isotopes
Krypton has six stable isotopes: ⁷⁸Kr, ⁸⁰Kr, ⁸²Kr, ⁸³Kr, ⁸⁴Kr, and ⁸⁶Kr, which together comprise the element's natural occurrence in Earth's atmosphere. Their relative abundances are 0.35% for ⁷⁸Kr, 2.25% for ⁸⁰Kr, 11.6% for ⁸²Kr, 11.5% for ⁸³Kr, 57.0% for ⁸⁴Kr, and 17.3% for ⁸⁶Kr. These values yield a standard atomic weight of 83.798 ± 0.002.[52]The table below summarizes the key properties of these isotopes, including atomic masses and nuclear spins:
[53]The natural isotopic distribution of krypton shows minor variations due to mass-dependent fractionation, arising from processes like molecular diffusion in the atmosphere or solubility during air-sea gas exchange; these shifts are typically less than 1‰ and reflect environmental influences on noble gas partitioning.[54] Such variations are analyzed to trace geochemical cycles and atmospheric dynamics.Measurements of krypton isotope abundances were historically refined through mass spectrometry, with seminal work by Alfred O. C. Nier in 1950 providing high-precision ratios that established the foundational dataset for modern standards. Among the stable isotopes, ⁸³Kr's nuclear spin of 9/2 enables its use in nuclear magnetic resonance spectroscopy for probing atomic environments.[53]Stable krypton isotope ratios contribute to atmospheric studies by serving as inert tracers for air mass mixing and transport, particularly in assessing fractionation during ocean-atmosphere interactions that influence global circulation patterns.[54]
Radioactive Isotopes
Krypton possesses numerous radioactive isotopes, but only a few have half-lives long enough to be of practical significance, with krypton-85 (⁸⁵Kr) being the most prominent due to its anthropogenic production and environmental persistence. This isotope is generated as a byproduct of uranium and plutoniumfission in nuclear reactors, with a cumulative fission yield of approximately 0.3%, meaning three atoms of ⁸⁵Kr are produced per 1,000 fission events.[41] It decays via β⁻ emission to stable rubidium-85 (⁸⁵Rb), with a half-life of 10.76 years and a maximum β particle energy of 687 keV (average energy of 251 keV), releasing no significant γ radiation in over 99% of decays.[55] The decay chain is straightforward, transitioning directly from ⁸⁵Kr to stable ⁸⁵Rb without intermediate metastable states of note.Another key radioactive isotope is krypton-81 (⁸¹Kr), which occurs naturally through cosmic ray-induced spallation of heavier krypton isotopes in the upper atmosphere. With a half-life of 229,000 years, ⁸¹Kr primarily decays by electron capture to bromine-81 (⁸¹Br), emitting low-energy X-rays and Auger electrons but no β particles or γ rays of high yield. This long half-life makes it suitable for geochronological applications, particularly in dating ancient groundwater systems where recharge times extend up to 1 million years, as the isotope's conservative behavior in aquifers allows precise age determination without significant geochemical fractionation.[56]Shorter-lived isotopes, such as krypton-87 (⁸⁷Kr), are typically produced in nuclear reactions or fission but decay rapidly, limiting their environmental impact. ⁸⁷Kr has a half-life of 76.3 minutes and undergoes β⁻ decay to stable rubidium-87 (⁸⁷Rb), with endpoint β energy around 2.3 MeV, though its brief existence means it is rarely isolated or monitored outside laboratory settings.[57] In contrast to stable krypton isotopes, these radioactive variants highlight the element's role in nuclear processes and atmospheric tracing.The primary environmental concern for radioactive krypton stems from ⁸⁵Kr releases during nuclear fuel reprocessing and reactor operations, which have elevated its global atmospheric concentration to about 1.5 Bq/m³ in recent years, primarily in the Northern Hemisphere due to industrial activity.[58] This anthropogenic input overshadows natural production levels, serving as a tracer for air mass movements and a marker for verifying compliance with nuclear non-proliferation treaties, while ⁸¹Kr remains at trace cosmogenic levels unaffected by human sources.
Applications
Industrial and Technological Uses
Krypton is employed in the production of specialized incandescent bulbs, where it serves as a fill gas to extend filament life and produce a whiter light output compared to standard argon-filled bulbs. This enhancement occurs because krypton's higher atomic mass reduces filament evaporation at high temperatures, allowing bulbs to operate longer without significant blackening. For instance, krypton-filled halogen lamps are used in applications requiring high illumination quality, such as projectors, where they provide brighter and more efficient lighting. Krypton is also used in high-intensity discharge (HID) lamps and photographic flash tubes due to its high light output efficiency.[59][60][61][62]In laser technology, krypton-ion lasers are utilized for their precise emission lines, particularly at 647 nm (red) and 568 nm (orange), which enable accurate beam alignment in industrial and technological settings. These lasers operate by exciting krypton ions in a plasma discharge, producing stable, low-divergence beams suitable for alignment tasks in optical systems and manufacturing processes. Their application in scientific alignment supports high-precision tasks, such as in photolithography equipment calibration.[63][64][65]Krypton's low thermal conductivity makes it valuable in energy-efficient double-glazed windows, where it fills the space between panes to minimize heat transfer. Compared to air, krypton reduces thermal conductivity by approximately 64%, and relative to argon, it offers about 30% better insulation performance, enhancing overall window U-values in building envelopes. This property is particularly beneficial in thinner glazing units, improving energy efficiency without increasing frame size.[66][67][68]As a shielding gas, krypton is used in specialized arc welding processes, particularly for stainless steel and electronic components, where its inert nature protects the weld pool from oxidation and atmospheric contamination. In these applications, krypton helps maintain arc stability and prevents heat-induced defects in sensitive materials.[69][70]Krypton serves as a propellant in ion thrusters for spacecraft propulsion, offering a cost-effective alternative to xenon due to its availability and similar ionization properties. It has been tested in thrusters like Hall-effect types, providing efficient thrust for long-duration missions.[71][72]Trace amounts of krypton are incorporated in semiconductor manufacturing, specifically in plasma etching processes, to enhance etch selectivity and uniformity during circuit patterning. Its use in reactive ion etching mixtures, such as for DRAM and 3D NAND devices, leverages krypton's ionization properties to improve process control and reduce defects in silicon wafers.[49][73][74]
Scientific and Medical Uses
Krypton plays a significant role in astrophysical research through its spectral lines, which enable the study of elemental abundances in stellar atmospheres and nebulae. In particular, emission lines from ionized krypton, such as those from [Kr IV], have been identified in the spectra of planetary nebulae, which are the ejected envelopes of asymptotic giant branch (AGB) stars. These observations provide insights into the s-process nucleosynthesis occurring in AGB stars, where krypton isotopes are produced via slow neutron capture, contributing to the chemical evolution of galaxies.[75][76]In mass spectrometry, the isotope ^{84}Kr serves as a key reference standard for noble gas analysis due to its abundance and lack of radiogenic interference. This isotope is routinely measured in laboratory protocols to calibrate instruments and normalize data for heavier noble gases like xenon, ensuring high precision in quantifying atmospheric, oceanic, and geological samples. For instance, in groundwaterdating and paleoclimate studies, ^{84}Kr ratios help trace recharge conditions and noble gas recharge temperatures.[77][78]Medically, the short-lived isotope ^{81m}Kr is employed in ventilation imaging for assessing lung function, particularly in detecting pulmonary embolism. With a half-life of 13 seconds, ^{81m}Kr allows for continuous inhalation from a rubidium generator, enabling real-time dynamic scans with low radiation exposure and high spatial resolution compared to longer-lived alternatives like xenon-133. This technique visualizes regional ventilation patterns, aiding in the diagnosis of obstructive lung diseases. Radioactive isotopes like ^{81m}Kr are thus valuable in nuclear medicine diagnostics.[79][80]Krypton's low boiling point of 119.8 K makes it suitable for cryogenic applications, including experiments involving superfluid media. In matrix isolation studies, krypton aggregates are embedded in superfluid helium to stabilize impurities like nitrogen atoms, allowing electron spin resonance (ESR) spectroscopy at temperatures below 2 K to probe quantum properties and atomic interactions. This setup leverages krypton's inertness and cryogenic properties for low-temperature physics research.[81]Krypton has niche applications in quantum technologies and advanced optics. As a buffer gas in ion traps, krypton facilitates cooling of trapped ions, such as magnesium or ytterbium, by thermalizing their motion through collisions, which is crucial for maintaining coherence in quantum computing experiments. This approach enhances ion crystal formation and gate fidelity in scalable qubit arrays. Additionally, KrF excimer lasers, operating at 248 nm, are used in deep-ultraviolet photolithography for fabricating nanoscale structures, supporting research in semiconductor physics and materials science.[82][83][84]
Safety and Biological Role
Health Effects
Krypton, a noble gas, is chemically inert and exhibits no known biological toxicity in its stable form, posing risks primarily as a simple asphyxiant when present in high concentrations that displace oxygen in the breathing air. Krypton has no known biological role in living organisms, consistent with its chemical inertness as a noble gas. Exposure to concentrations exceeding 50% can lead to oxygen deprivation, resulting in symptoms such as dizziness, nausea, vomiting, loss of consciousness, and death if not addressed promptly.[85][86][87]The gas has low solubility in blood, with a blood-gas partition coefficient (Ostwald solubility coefficient) of approximately 0.105 vol/vol at 37°C, facilitating rapid diffusion into the bloodstream and subsequent exhalation, which minimizes accumulation during short-term exposure.[88] Although the U.S. Occupational Safety and Health Administration (OSHA) does not establish a specific permissible exposure limit for krypton due to its inert nature, it is managed as a simple asphyxiant, with workplace guidelines emphasizing maintenance of oxygen concentrations above 19.5% to prevent hypoxic effects during 8-hour exposures.The radioactive isotope ^{85}Kr, a beta-emitter with a half-life of 10.76 years, presents an internal radiationhazard primarily through inhalation, though its low bloodsolubility limits retention time in the body to minutes. As of 2025, global atmospheric concentrations of ^{85}Kr are approximately 1.5 Bq/m³. Due to its low solubility and rapid exhalation, the annual effective dose from environmental exposure is negligible (less than 0.001 μSv/year), posing no significant health risk.[89][90]Epidemiological and environmental monitoring data indicate no documented adverse health effects from chronic low-level exposure to either stable krypton or ^{85}Kr in the general population. Research on noble gas effects under hyperbaric conditions has demonstrated that krypton can induce narcosis, similar to other inert gases, at elevated partial pressures, but such scenarios are irrelevant to ambient atmospheric exposures.[91][92]
Handling Precautions
Krypton, as a compressed inert gas, is typically stored in high-pressure steel cylinders rated up to 300bar to prevent rupture from overpressurization, and these should be kept in a cool, dry, well-ventilated area away from flammable materials, heat sources, ignition points, and incompatible substances such as oxidizing agents.[93] Cylinders must be secured upright with chains or straps to avoid falling, protected from physical damage, and separated from empty containers to minimize hazards during handling or potential leaks.[94] For cryogenic liquid krypton, storage occurs in insulated Dewar flasks designed to maintain low temperatures around -153°C, with pressure-relief valves to manage vaporization and prevent overpressurization.[95]Safe handling requires using only equipment compatible with high-pressure inert gases, such as pressure-rated piping and regulators, and cylinders should be moved via hand trucks rather than dragged or rolled to avoid valve damage.[96] As krypton is colorless and odorless, leaks are detected using soap solution applied to joints and fittings for bubble formation or electronic oxygen sensors to monitor for displacement in enclosed spaces.[93] Personal protective equipment includes work gloves (EN 388 standard for mechanical protection) and safety shoes for cylinder handling, safety glasses or face shields, and self-contained breathing apparatus (SCBA) in confined or poorly ventilated areas to mitigate asphyxiation risks from oxygen displacement. Adequate ventilation must be ensured during use to maintain oxygen levels above 19.5%.[94]Transport of compressed krypton complies with UN 1056 classification as a non-flammable compressed gas (Class 2.2), requiring labeling as "Non-Flammable Gas" and secure packaging to withstand pressures without leakage.[97] In the United States, the Department of Transportation (DOT) designates it under Hazard Class 2.2, with proper shipping name "Krypton, Compressed," prohibiting transport in the same compartment as foodstuffs or oxidizers.[96]In emergencies, such as a cylinder leak, evacuate the area immediately, ventilate to dilute the gas with fresh air, and monitor oxygen concentrations; krypton poses no ignition or chemical reactivity risks in its elemental form but can cause asphyxiation if concentrated.[93] For krypton compounds like KrF₂, additional precautions are needed due to its corrosivity, including use of chemical-resistant PPE and containment to avoid skin or material damage.[98]