Difference between revisions of "User:Physchim62/Helium"

From WikiChem
Jump to: navigation, search
(Liquid helium)
(Liquid helium)
Line 108: Line 108:
 
{| class="wikitable" align=right style="margin:0 0 0 0.5em;"
 
{| class="wikitable" align=right style="margin:0 0 0 0.5em;"
 
|-
 
|-
!  
+
! rowspan=2 |  
 
! colspan=2 | [[Helium-4|<sup>4</sup>He]]
 
! colspan=2 | [[Helium-4|<sup>4</sup>He]]
 
! colspan=2 | [[Helium-3|<sup>3</sup>He]]
 
! colspan=2 | [[Helium-3|<sup>3</sup>He]]
Line 132: Line 132:
 
| 1.0 K
 
| 1.0 K
 
| 25 bar
 
| 25 bar
| 1 mK
+
| 0.3 K
 
| 29 bar
 
| 29 bar
 
|-
 
|-
| [[λ point]]
+
| [[λ point]] at<br/>saturated vapour pressure
|  
+
| 2.17K
|
 
 
|  
 
|  
 +
| 1 mK
 
|  
 
|  
 
|-
 
|-
Line 145: Line 145:
 
Liquid helium-4 can be condensed at 4.22&nbsp;K to a liquid phase now known as [[helium I]]. If cooling is continued (by allowing some of the liquid to evaporate), the physical properties of the liquid change drastically at about 2.2&nbsp;K (the exact value depends on the pressure). For a start, the visible bubbling of a boiling liquid disappears, although [[evaporation]] continues: the [[heat capacity]] increases ten-fold, the [[thermal conductivity]] increases a ''million-fold'' and the [[viscosity]] drops to virtually zero.
 
Liquid helium-4 can be condensed at 4.22&nbsp;K to a liquid phase now known as [[helium I]]. If cooling is continued (by allowing some of the liquid to evaporate), the physical properties of the liquid change drastically at about 2.2&nbsp;K (the exact value depends on the pressure). For a start, the visible bubbling of a boiling liquid disappears, although [[evaporation]] continues: the [[heat capacity]] increases ten-fold, the [[thermal conductivity]] increases a ''million-fold'' and the [[viscosity]] drops to virtually zero.
  
The low-temperature liquid phase is known as [[helium II]], and is a [[superfluid]] similar in nature to a [[Bose–Einstein condensate]], first discovered in 1938. The transition between the two liquid phases is a [[second-order phase transition]], and the transition temperature is known as the [[λ point]] after the shape of the plot of heat capacity vs. temperature, which resembles a Greek letter lambda.
+
The low-temperature liquid phase is known as [[helium II]], and is a [[superfluid]] similar in nature to a [[Bose–Einstein condensate]], first discovered in 1938. The transition between the two liquid phases is a [[second-order phase transition]], and the transition temperature is known as the [[λ point]] after the shape of the plot of heat capacity vs. temperature, which resembles a Greek letter lambda. Perhaps the best known property of helium&nbsp;II is its propensity to form Rollin films, layers approximately 30&nbsp;nm thick that will climb up surfaces cool enough to support helium&nbsp;II even against the force of gravity: as such, the confinement of liquid helium below the λ&nbsp;point is particularly difficult.
  
[[Helium-3]] also forms a superfluid phase, discovered in 1972, but only at temperatures of a few millikelvin above absolute zero.
+
[[Helium-3]] also forms a superfluid phase, discovered in 1972, but only at temperatures of one millikelvin or less. The difference in behaviour between the two isotopes is due to their nuclear spin: helium-4 atoms are [[boson]]s, with zero [[spin]], while helium-3 atoms are spin-½ [[fermion]]s, and the two types of particle obey different exclusion principles. For similar reasons, liquid helium-4 and liquid helium-3 are no longer miscible below about 0.9&nbsp;K.
  
Neither isotope can be solidified at pressures of 1&nbsp;atm. By raising the pressure to 25&nbsp;atm or higher, helium can be can be solidified at 1–1.5&nbsp;K.
+
Neither isotope can be solidified at pressures of 1&nbsp;atm. By raising the pressure to 25&nbsp;atm or higher, helium can be can be solidified at 1–1.5&nbsp;K. The solid form is unusually compressible, with a bulk modulus of around 50&nbsp;MPa (fifty-times more compressible than liquid water). Liquid and solid phases of helium have extremely low [[refractive index]]es, which makes it difficult to distinguish phases changes by visual means.
  
 
==Chemical properties==
 
==Chemical properties==

Revision as of 14:53, 19 March 2010

hydrogenheliumlithium


He

Ne
Atomic properties
Atomic number 2
Standard atomic weight 4.002 602(2)
Electron configuration 1s2
Physical properties[1]
Melting point see text
Boiling point 4.22 K (−268.93 °C)
Chemical properties
Solubility in water 8.61 cm3 kg−1 (101.325 kPa, 20 °C)
Ionization energies
1st 2372.3 kJ mol−1
2nd 5250.5 kJ mol−1
Thermodynamic properties[1][2]
Standard entropy 126.153(2) J K−1 mol−1
Enthalpy change of vaporization 0.08 kJ mol−1
Miscellaneous
CAS number 7440-59-7
Where appropriate, and unless otherwise stated, data are given for 100 kPa (1 bar) and 298.15 K (25 °C).

Helium (symbol: He) is a chemical element, the lightest of the noble gases.

History

Helium is the only element to have been discovered extraterrestrially before being found on Earth, specifically in the Sun.[2] A bright yellow spectral line (λ = 587.49 nm) was first observed by French astronomer Jules Janssen during the eclipse of 18 August 1868, which Janssen observed from Guntur in India, and independently by Norman Lockyer in London on 20 October 1868.[3][4][5] Lockyer, together with English chemist Edward Frankland, showed that the line could not be explained by any known element, and proposed the name helium, from the Greek ἥλιος (helios; the Sun).[6][7]

The first isolation of helium from a terrestrial source is usually credited to the British chemist William Ramsey, who isolated the helium that was occluded in the mineral cleveite (a uranium-containing mineral) and identified it by its spectrum.[8] The occluded gas had been noticed previously by American geochemist William Francis Hillebrand, but had been misidentified as nitrogen.[9] Helium had been identified (through its spectrum) on Earth as early as 1881 by Italian physicist Luigi Palmieri,[2] but Palmieri could not isolate a sample of the element. Finally, helium was isolated independently in 1895 by Per Teodor Cleve and Abraham Langlet in Uppsala, Sweden, who collected enough of the gas to accurately determine its atomic weight.[10] Nevertheless, it was Ramsey who was awarded the Nobel Prize for Chemistry (in 1904), and who is usually credited with the discovery.

Occurrence and production

Country Production
106 m3
[note 1]
United States 122*
Algeria 24
Qatar 15
Russia 7
Poland 2.5
Estimates for 2009 from the
U.S. Geological Survey[11]
*U.S. figure includes 42 million
cubic metres withdrawn from the
federal government stockpile.

Helium is the second most common element in the Universe (after hydrogen), accounting for 23% of all atoms. However, the Earth's gravitational field is not strong enough to retain helium in the atmosphere for long periods, and almost all the Earth's primordial helium is believed to have escaped.[12] The helium currently present on Earth has been formed from the α-decay of radioactive nuclides:[note 2] most of this helium escapes to the atmosphere, where the current volume fraction is 5.24 ppm, and then into space: some of it, however, can be trapped underground by impermeable rock formations, often associated with natural gas deposits.

The commercial production of helium is based around its extraction from natural gas, which is economically viable when the helium fraction is greater than about 0.3%.[2] The United States has historically been the predominant producer, although its de novo production is declining and Algeria and Qatar are gaining importance.

The U.S. Bureau of Land Management operates a stockpile of crude (~80%) helium, the National Helium Reserve at Cliffside Field, Potter County, Texas, and a crude-helium pipeline from Bushton, Kansas, passing through the Reichel Field (Kansas) and the Keyes Field (Oklahoma) to the Cliffside Field.[13] The National Helium Reserve is being run down under the terms of the Helium Privatization Act of 1996 (Pub. L. 104–273), and about one third of the helium produced in the United States is refined from stored helium rather than being extracted from natural gas.[11]

Use

Total U.S. consumpton (2009)
52.1 million cubic metres
Cryogenics 32%
Pressurizing and purging 18%
Controlled atmospheres 18%
Welding cover gas 13%
Leak detection 4%
Breathing mixtures 2%
Other uses 13%
Estimates for 2009 from the U.S. Geological Survey[11]
The breakdown of usage in other countries may be significantly different: see text.

The main use of helium is in cryogenics, where helium is essential for temperatures below −256 °C (−429 °F),[11] the approximate boiling point of liquid hydrogen. Cooling with liquid helium allows an operating temperature of around −269 °C (−452 °F), just four kelvins. Such low temperatures are required for superconducting magnets to operate; these are used in a variety of applications, including NMR spectrometers and magnetic resonance imaging (MRI) scanners.

Uses for pressurizing and purging, in controlled atmospheres (e.g., glove boxes) and as a welding cover gas are only important in the United States, where helium is relatively cheap, and are declining even there along with a general fall in U.S. helium consumption (−42% over the period 2000–2009).[14][15][note 3] In other countries, argon is used for these purposes.

There are several niche uses of helium, including:[2]

Physical properties

Helium has a number of unique physical properties. It is the chemical substance with the lowest boiling point (4.22 K) and the only substance that cannot be solidified by cooling alone at atmospheric pressure. It is also the only substance without a triple point, a temperature and pressure at which solid, liquid and gas are all in equilibrium.[2]

Helium has an extremely low molar heat capacity and, for a gas, a relatively high thermal conductivity (0.1430 W m−1 K−1 at 0 °C):[2] both of these are due to the low mass of the atoms which make up helium gas (H2 is even more exceptional on both counts), and are important in its uses in gas chromatography and as a heat transfer agent.

Liquid helium

  4He 3He
T p T p
Critical point 5.2 K 3.3 K
Boiling point 4.2 K 1 atm 3.2 K 1 atm
Melting point at
minimum freezing pressure
1.0 K 25 bar 0.3 K 29 bar
λ point at
saturated vapour pressure
2.17K 1 mK

Liquid helium-4 can be condensed at 4.22 K to a liquid phase now known as helium I. If cooling is continued (by allowing some of the liquid to evaporate), the physical properties of the liquid change drastically at about 2.2 K (the exact value depends on the pressure). For a start, the visible bubbling of a boiling liquid disappears, although evaporation continues: the heat capacity increases ten-fold, the thermal conductivity increases a million-fold and the viscosity drops to virtually zero.

The low-temperature liquid phase is known as helium II, and is a superfluid similar in nature to a Bose–Einstein condensate, first discovered in 1938. The transition between the two liquid phases is a second-order phase transition, and the transition temperature is known as the λ point after the shape of the plot of heat capacity vs. temperature, which resembles a Greek letter lambda. Perhaps the best known property of helium II is its propensity to form Rollin films, layers approximately 30 nm thick that will climb up surfaces cool enough to support helium II even against the force of gravity: as such, the confinement of liquid helium below the λ point is particularly difficult.

Helium-3 also forms a superfluid phase, discovered in 1972, but only at temperatures of one millikelvin or less. The difference in behaviour between the two isotopes is due to their nuclear spin: helium-4 atoms are bosons, with zero spin, while helium-3 atoms are spin-½ fermions, and the two types of particle obey different exclusion principles. For similar reasons, liquid helium-4 and liquid helium-3 are no longer miscible below about 0.9 K.

Neither isotope can be solidified at pressures of 1 atm. By raising the pressure to 25 atm or higher, helium can be can be solidified at 1–1.5 K. The solid form is unusually compressible, with a bulk modulus of around 50 MPa (fifty-times more compressible than liquid water). Liquid and solid phases of helium have extremely low refractive indexes, which makes it difficult to distinguish phases changes by visual means.

Chemical properties

Helium has the highest first ionization energy of any element (2372.3 kJ mol−1) and has almost unprecedented chemical inertness. No chemical compounds of helium are known.

The low solubility of helium in water (8.61 cm3 kg−1 at 101.325 kPa, 20 °C)[2] is important in its use as a component of breathing mixtures: helium, unlike nitrogen at high pressure, will not dissolve in blood serum, and so poses fewer problems of degassing of the blood stream during decompression.

Notes and references

Notes

  1. Most U.S. Geological Survey sources quote amounts of helium in millions of cubic metres, measured at 101.325 kPa and 15 °C. Older U.S. sources may quote values in cubic feet, measured at 14.7 psi and 70 °F: one cubic metre (101.325 kPa, 15 °C) = 36.053 cubic feet (14.7 psi, 70 °F). 106 m3 (101.325 kPa, 15 °C) = 169.29 tonnes.
  2. The helium-3 isotope (x = 1.34(3) ppm in the atmosphere) is formed from the β-decay of tritium (31H): the existence of helium from volcanic rocks or associated geothermal springs and gases with x(3He) > 10 ppm indicates that at least some of this helium-3 is primordial.
  3. U.S. consumption of helium in 2000 was 89.8 million cubic metres, of which: cryogenics 24%, pressurizing and purging 20%, welding cover gas 18%, controlled atmospheres 16%, leak detection 6%, breathing mixtures 3%, other uses (chromatography/lifting gas/heat transfer) 13%.

References

  1. 1.0 1.1 Helium. In NIST Chemistry WebBook; National Institute for Standards and Technology, <http://webbook.nist.gov/cgi/inchi/InChI%3D1S/He>. (accessed 19 March 2010).
  2. 2.0 2.1 2.2 2.3 2.4 2.5 2.6 2.7 Greenwood, Norman N.; Earnshaw, A. Chemistry of the Elements; Pergamon: Oxford, 1984; pp 1042–59. ISBN 0-08-022057-6.
  3. Cortie, A. L. Sir Norman Lockyer, 1836–1920. Astrophys. J. 1921, 53 (4), 233–48, <http://articles.adsabs.harvard.edu/cgi-bin/nph-iarticle_query?1921ApJ....53..233C&data_type=PDF_HIGH&whole_paper=YES&type=PRINTER&filetype=.pdf>.
  4. Leggett, Hadley Aug. 18, 1868: Helium Discovered During Total Solar Eclipse; wired.com, August 18, 2009, <http://www.wired.com/thisdayintech/2009/08/dayintech_0818/>. (accessed 18 March 2010).
  5. C. R. Hebd. Acad. Sci. Paris 1868, 67, 836–41, <http://gallica.bnf.fr/ark:/12148/bpt6k3024c.image.r=comptes-rendus+hebdomadaires+Acad%C3%A9mie+des+Sciences.f836.langFR>.
  6. Helium. In Oxford English Dictionary; Oxford University Press, 2008, <http://dictionary.oed.com/cgi/entry/50104457?>. (accessed 20 July 2008).
  7. "Frankland and Lockyer find the yellow prominences to give a very decided bright line not far from D, but hitherto not identified with any terrestrial flame. It seems to indicate a new substance, which they propose to call Helium." Thomson, W. Rep. Brit. Assoc. 1872, 99.
  8. Ramsay, William On a Gas Showing the Spectrum of Helium, the Reputed Cause of D3, One of the Lines in the Coronal Spectrum. Preliminary Note. Proc. Roy. Soc. London 1895, 58, 65–67. DOI: 10.1098/rspl.1895.0006. Ramsay, William Helium, a Gaseous Constituent of Certain Minerals. Part I. Proc. Roy. Soc. London 1895, 58, 80–89. DOI: 10.1098/rspl.1895.0010. Ramsay, William Helium, a Gaseous Constituent of Certain Minerals. Part II. Proc. Roy. Soc. London 1895, 59, 325–30. DOI: 10.1098/rspl.1895.0097.
  9. W. F. Hillebrand (1853–1925), geochemist and US Bureau of Standards administrator. In American National Biography; Garraty, John A.; Carnes, Mark C., Eds.; Oxford University Press, 1999; Vol. 10-11, pp 227–28, 808–9.
  10. Langlet, N. A. Das Atomgewicht des Heliums. Z. Anorg. Chem. 1895, 10 (1), 289–92. DOI: 10.1002/zaac.18950100130.
  11. 11.0 11.1 11.2 11.3 Pacheco, Norbert; Thomas, Diedre S. Helium. In Mineral Commodities Summaries; U.S. Geological Survey, January 2010, <http://minerals.usgs.gov/minerals/pubs/commodity/helium/mcs-2010-heliu.pdf>.
  12. Atomic weights of the elements. Review 2000. Pure Appl. Chem., 75 (6), 683–800. DOI: 10.1351/pac200375060683.
  13. Pacheco, Norbert Helium. In 2008 Minerals Yearbook; U.S. Geological Survey, October 2009, <http://minerals.usgs.gov/minerals/pubs/commodity/helium/myb1-2008-heliu.pdf>.
  14. Kelly, T. D.; Matos, G. R. Helium statistics. In Historical statistics for mineral and material commodities in the United States; U.S. Geological Survey, 2003. U.S. Geological Survey Data Series 140, <http://minerals.usgs.gov/ds/2005/140/helium-use.pdf>. (accessed 19 March 2010).
  15. Kelly, T. D.; Matos, G. R. Helium statistics. In Historical statistics for mineral and material commodities in the United States; U.S. Geological Survey, 2008. U.S. Geological Survey Data Series 140, <http://minerals.usgs.gov/ds/2005/140/helium.pdf>. (accessed 19 March 2010).

External links