- Art Gallery -

Dysprosium

Dysprosium is a chemical element with the symbol Dy and atomic number 66. It is a rare earth element with a metallic silver luster. Dysprosium is never found in nature as a free element, though it is found in various minerals, such as xenotime. Naturally occurring dysprosium is composed of 7 isotopes, the most abundant of which is 164Dy.

Dysprosium was first identified in 1886 by Paul Émile Lecoq de Boisbaudran, but was not isolated in pure form until the development of ion exchange techniques in the 1950s. Dysprosium is used for its high thermal neutron absorption cross-section in making control rods in nuclear reactors, for its high magnetic susceptibility to magnetization in data storage devices and as a component of Terfenol-D. Soluble dysprosium salts are mildly toxic, while the insoluble salts are considered non-toxic.

Characteristics

Physical
Dysprosium sample

Dysprosium is a rare earth element that has a metallic, bright silver luster. It is soft enough to be cut with a knife, and can be machined without sparking if overheating is avoided. Dysprosium's physical characteristics can be greatly affected even by small amounts of impurities.[1]

Dysprosium and holmium have the highest magnetic strengths of the elements,[2] especially at low temperatures.[3] Dysprosium has a simple ferromagnetic ordering at temperatures below 85 K. Above 85 K, it turns into an helical antiferromagnetic state in which all of the atomic moments in a particular basal plane layer are parallel, and oriented at a fixed angle to the moments of adjacent layers. This unusual antiferromagnetism transforms into a disordered (paramagnetic) state at 179 K.[4]

Chemical

Dysprosium metal tarnishes slowly in air and burns readily to form dysprosium(III) oxide:

4 Dy + 3 O2 → 2 Dy2O3

Dysprosium is quite electropositive and reacts slowly with cold water and quite quickly with hot water to form dysprosium hydroxide:

2 Dy (s) + 6 H2O (l) → 2 Dy(OH)3 (aq) + 3 H2 (g)

Dysprosium metal vigorously reacts with all the halogens at above 200 °C:

2 Dy (s) + 3 F2 (g) → 2 DyF3 (s) [green]
2 Dy (s) + 3 Cl2 (g) → 2 DyCl3 (s) [white]
2 Dy (s) + 3 Br2 (g) → 2 DyBr3 (s) [white]
2 Dy (s) + 3 I2 (g) → 2 DyI3 (s) [green]

Dysprosium dissolves readily in dilute sulfuric acid to form solutions containing the yellow Dy(III) ions, which exist as a [Dy(OH2)9]3+ complexes:[5]

2 Dy (s) + 3 H2SO4 (aq) → 2 Dy3+ (aq) + 3 SO2−4 (aq) + 3 H2 (g)


Compounds
See also: Dysprosium compounds

Dysprosium halides, such as DyF3 and DyBr3, tend to take on a yellow color. Dysprosium oxide, also known as dysprosia, is a white powder that is highly magnetic, more so than iron oxide.[3]

Dysprosium combines with various non-metals at high temperatures to form binary compounds with varying composition and oxidation states +3 and sometimes +2, such as DyN, DyP, DyH2 and DyH3; DyS, DyS2, Dy2S3 and Dy5S7; DyB2, DyB4, DyB6 and DyB12, as well as Dy3C and Dy2C3.[6]

Dysprosium carbonate, Dy2(CO3)3, and dysprosium sulfate, Dy2(SO4)3, result from similar reactions.[7] Most dysprosium compounds are soluble in water, though dysprosium carbonate tetrahydrate (Dy2(CO3)3·4H2O) and dysprosium oxalate decahydrate (Dy2(C2O4)3·10H2O) are both insoluble in water.[8][9]

Isotopes
Main article: Isotopes of dysprosium

Naturally occurring dysprosium is composed of 7 isotopes: 156Dy, 158Dy, 160Dy, 161Dy, 162Dy, 163Dy, and 164Dy. These are all considered stable, although 156Dy decays by alpha decay with a half-life of over 1×1018 years. Of the naturally occurring isotopes, 164Dy is the most abundant at 28%, followed by 162Dy at 26%. The least abundant is 156Dy at .06%.[10]

Twenty-nine radioisotopes have also been synthesized, ranging in atomic mass from 138 to 173. The most stable of these is 154Dy with a half-life of approximately 3 × 106 years, followed by 159Dy with a half-life of 144.4 days. The least stable is 138Dy with a half-life of 200 ms. Isotopes that are lighter than the stable isotopes tend to decay primarily by β+ decay, while those that are heavier tend to decay by β- decay, with some exceptions. 154Dy decays primarily by alpha decay, and 152Dy and 159Dy decay primarily by electron capture.[10] Dysprosium also has at least 11 metastable isomers, ranging in atomic mass from 140 to 165. The most stable of these is 165mDy, which has a half-life of 1.257 minutes. 149Dy has two excitation states, the second of which, 149m2Dy, has a half-life of 28 ns.[10]

History

In 1878, erbium ores were found to contain the oxides of holmium and thulium. French chemist Paul Émile Lecoq de Boisbaudran, while working with holmium oxide, separated dysprosium oxide from it in Paris in 1886.[11] His procedure for isolating the dysprosium involved dissolving dysprosium oxide in acid, then adding ammonia to precipitate the hydroxide. He was only able to isolate dysprosium from its oxide after more than 30 attempts at his procedure. Upon succeeding, he named the element dysprosium from the Greek dysprositos (δυσπρόσιτος), meaning "hard to get". However, the element was not isolated in relatively pure form until after the development of ion exchange techniques by Frank Spedding at Iowa State University in the early 1950s.[2]

In 1950, Glenn T. Seaborg, Albert Ghiorso, and Stanley G. Thompson bombarded 241Am with helium ions, which produced atoms with an atomic number of 97 and which closely resembled the neighboring lanthanide terbium. Because terbium was named after Ytterby, the city in which it and several other elements were discovered, this new element was named berkelium for the city in which it was synthesized. However, when the research team synthesized element 98, they could not think of a good analogy for dysprosium, and instead named the element californium in honor of the state in which it was synthesized. The research team went on to "point out that, in recognition of the fact that dysprosium is named on the basis of a Greek word meaning 'difficult to get at,' that the searchers for another element a century ago found it difficult to get to California."[12]

Occurrence
Xenotime

Dysprosium is never encountered as a free element, but is found in many minerals, including xenotime, fergusonite, gadolinite, euxenite, polycrase, blomstrandine, monazite and bastnäsite; often with erbium and holmium or other rare earth elements. Currently, most dysprosium is being obtained from the ion-adsorption clay ores of southern China.[13] In the high-yttrium version of these, dysprosium happens to be the most abundant of the heavy lanthanides, comprising up to 7–8% of the concentrate (as compared to about 65% for yttrium).[14][15] The concentration of Dy in the Earth crust is about 5.2 mg/kg and in sea water 0.9 ng/L.[6]

Production

Dysprosium is obtained primarily from monazite sand, a mixture of various phosphates. The metal is obtained as a by-product in the commercial extraction of yttrium. In isolating dysprosium, most of the unwanted metals can be removed magnetically or by a flotation process. Dysprosium can then be separated from other rare earth metals by an ion exchange displacement process. The resulting dysprosium ions can then react with either fluorine or chlorine to form dysprosium fluoride, DyF3, or dysprosium chloride, DyCl3. These compounds can be reduced using either calcium or lithium metals in the following reactions:[7]

3 Ca + 2 DyF3 → 2 Dy + 3 CaF2
3 Li + DyCl3 → Dy + 3 LiCl

The components are placed in a tantalum crucible and fired in a helium atmosphere. As the reaction progresses, the resulting halide compounds and molten dysprosium separate due to differences in density. When the mixture cools, the dysprosium can be cut away from the impurities.[7]

About 100 tonnes of dysprosium are produced worldwide each year,[16] with 99% of that total produced in China[17] Dysprosium prices have climbed nearly sevenfold since 2003, to $53 a pound.[13]

Applications

Dysprosium is used, in conjunction with vanadium and other elements, in making laser materials. Because of dysprosium's high thermal neutron absorption cross-section, dysprosium oxide-nickel cermets are used in neutron-absorbing control rods in nuclear reactors.[2][18] Dysprosium-cadmium chalcogenides are sources of infrared radiation which is useful for studying chemical reactions.[1] Because dysprosium and its compounds are highly susceptible to magnetization, they are employed in various data storage applications, such as in hard discs.[19]

Neodymium-iron-boron magnets can have up to 6% of the neodymium substituted with dysprosium[20] to raise the coercivity for demanding applications such as drive motors for hybrid electric vehicles. This substitution would require up to 100 grams of dysprosium per hybrid car produced. Based on Toyota's projected 2 million units per year, the use of dysprosium in applications such as this would quickly exhaust the available supply of the metal.[21] The dysprosium substitution may also be useful in other applications, as it improves the corrosion resistance of the magnets.[22]

Dysprosium is one of the components of Terfenol-D, along with iron and terbium. Terfenol-D has the highest room-temperature magnetostriction of any known material;[23] this property is employed in transducers, wide-band mechanical resonators,[24] and high-precision liquid fuel injectors.[25]
Nanofibers of dysprosium oxide fluoride

Dysprosium is used in dosimeters for measuring ionizing radiation. Crystals of calcium sulfate or calcium fluoride are doped with dysprosium. When these crystals are exposed to radiation, the dysprosium atoms become excited and luminescent. The luminescence can be measured to determine the degree of exposure to which the dosimeter has been subjected.[2]

Nanofibers of dysprosium compounds have high strength and large surface area; therefore, they can be used for reinforcement of other materials and as a catalyst. Fibers of dysprosium oxide fluoride can be produced by heating an aqueous solution of DyBr and NaF to 450 °C at 450 bar pressure for 17 hours. This material is remarkably robust, surviving over 100 hours in various aqueous solutions at temperatures exceeding 400 °C without re-dissolving or aggregating.[26][27][28]

Dysprosium iodide and dysprosium bromide are used in high intensity lighting. These compounds dissociate near the hot center of the lamp releasing isolated dysprosium atoms. The latter re-emit light in the green and red part of the spectrum thereby effectively producing bright light.[2][29]

Precautions

Like many powders, dysprosium powder may present an explosion hazard when mixed with air and when an ignition source is present. Thin foils of the substance can also be ignited by sparks or by static electricity. Dysprosium fires cannot be put out by water. It can react with water to produce flammable hydrogen gas.[30] Dysprosium chloride fires, however, can be extinguished with water,[31] while dysprosium fluoride and dysprosium oxide are non-flammable.[32][33] Dysprosium nitrate, Dy(NO3)3, is a strong oxidizing agent and will readily ignite upon contact with organic substances.[3]

Soluble dysprosium salts, such as dysprosium chloride and dysprosium nitrate, are mildly toxic when ingested. The insoluble salts, however, are non-toxic. Based on the toxicity of dysprosium chloride to mice, it is estimated that the ingestion of 500 grams or more could be fatal to a human.[2]

See also

* Lanthanoid


References

1. ^ a b Lide, David R., ed (2007–2008). "Dysprosium". CRC Handbook of Chemistry and Physics. 4. New York: CRC Press. pp. 11. ISBN 978-0-8493-0488-0.
2. ^ a b c d e f Emsley, John (2001). Nature's Building Blocks. Oxford: Oxford University Press. pp. 129–132. ISBN 0198503415. http://books.google.com/books?id=j-Xu07p3cKwC&pg=PA131.
3. ^ a b c Krebs, Robert E. (1998). "Dysprosium". The History and Use of our Earth's Chemical Elements. Greenwood Press. pp. 234–235. ISBN 0-313-30123-9.
4. ^ Jackson, Mike (2000). (PDF)IRM Quarterly (Institute for Rock Magnetism) 10 (3): 6. http://www.irm.umn.edu/quarterly/irmq10-3.pdf.
5. ^ "Chemical reactions of Dysprosium". Webelements. https://www.webelements.com/dysprosium/chemistry.html. Retrieved 2009-06-06.
6. ^ a b Patnaik, Pradyot (2003). Handbook of Inorganic Chemical Compounds. McGraw-Hill. pp. 289–290. ISBN 0070494398. http://books.google.com/books?id=Xqj-TTzkvTEC&pg=PA243. Retrieved 2009-06-06.
7. ^ a b c Heiserman, David L. (1992). Exploring Chemical Elements and their Compounds. TAB Books. pp. 236–238. ISBN 0-8306-3018-X.
8. ^ Perry, D. L. (1995). Handbook of Inorganic Compounds. CRC Press. pp. 152–154. ISBN 0-8492-8671-3.
9. ^ Jantsch, G.; Ohl, A. (1911). "Zur Kenntnis der Verbindungen des Dysprosiums". Berichte der deutschen chemischen Gesellschaft 44 (2): 1274–1280. doi:10.1002/cber.19110440215.
10. ^ a b c Audi, G. (2003). "Nubase2003 Evaluation of Nuclear and Decay Properties". Nuclear Physics A (Atomic Mass Data Center) 729: 3–128. doi:10.1016/j.nuclphysa.2003.11.001.
11. ^ Paul Émile Lecoq de Boisbaudran (1886). "L'holmine (ou terre X de M Soret) contient au moins deux radicaux métallique (Holminia contains at least two metal)" (in French). Comptes Rendus 143: 1003–1006. http://gallica.bnf.fr/ark:/12148/bpt6k3058f/f1001.chemindefer.
12. ^ Weeks, M. E. (1968). Discovery of the Elements (7 ed.). Journal of Chemical Education. pp. 848–849. ISBN 0848685792. OCLC 23991202.
13. ^ a b Bradsher, Keith (December 25, 2009). "Earth-Friendly Elements, Mined Destructively". The New York Times. http://www.nytimes.com/2009/12/26/business/global/26rare.html.
14. ^ Naumov, A. V. (2008). "Review of the World Market of Rare-Earth Metals". Russian Journal of Non-Ferrous Metals 49 (1): 14–22. http://www.springerlink.com/content/y8925j378w4u4175/.
15. ^ Gupta, C. K.; Krishnamurthy N. (2005). Extractive Metallurgy of Rare Earths. CRC Press. ISBN 9780415333405. http://books.google.com/books?id=F0Bte_XhzoAC.
16. ^ "Dysprosium (Dy) - Chemical properties, Health and Environmental effects". Lenntech Water treatment & air purification Holding B.V.. 2008. http://www.lenntech.com/Periodic-chart-elements/Dy-en.htm. Retrieved 2009-06-02.
17. ^ Bradsher, Keith (August 31, 2009). "China Tightens Grip on Rare Minerals". The New York Times. http://www.nytimes.com/2009/09/01/business/global/01minerals.html?pagewanted=all.
18. ^ Amit, Sinha; Sharma, Beant Prakash (2005). "Development of Dysprosium Titanate Based Ceramics". Journal of the American Ceramic Society 88 (4): 1064–1066. doi:10.1111/j.1551-2916.2005.00211.x.
19. ^ Lagowski, J. J., ed (2004). Chemistry Foundations and Applications. 2. Thomson Gale. pp. 267–268. ISBN 0-02-865724-1.
20. ^ Shi, Fang, X.; Shi, Y.; Jiles, D.C. (1998). "Modeling of magnetic properties of heat treated Dy-doped NdFeBparticles bonded in isotropic and anisotropic arrangements". IEEE Transactions on Magnetics 34 (4): 1291–1293. doi:10.1109/20.706525.
21. ^ Campbell, Peter (February 2008). "Supply and Demand, Part 2". Princeton Electro-Technology, Inc.. http://www.magnetweb.com/Col05.htm. Retrieved 2008-11-09.
22. ^ Yu, L. Q.; Wen, Y; Yan, M (2004). "Effects of Dy and Nb on the magnetic properties and corrosion resistance of sintered NdFeB". Journal of Magnetism and Magnetic Materials 283 (2-3): 353–356. doi:10.1016/j.jmmm.2004.06.006.
23. ^ "What is Terfenol-D?". ETREMA Products, Inc.. 2003. http://etrema-usa.com/core/terfenold/. Retrieved 2008-11-06.
24. ^ Kellogg, Rick; Flatau, Alison (May 2004). "Wide Band Tunable Mechanical Resonator Employing the ΔE Effect of Terfenol-D". Journal of Intelligent Material Systems & Structures (Sage Publications, Ltd) 15 (5): 355–368. doi:10.1177/1045389X04040649.
25. ^ Leavitt, Wendy (February 2000). "Take Terfenol-D and call me". Fleet Owner (RODI Power Systems Inc) 95 (2): 97. http://0-search.ebscohost.com.ilsprod.lib.neu.edu/login.aspx?direct=true&db=buh&AN=2869368&site=ehost-live. Retrieved 2008-11-06.
26. ^ "Supercritical Water Oxidation/Synthesis". Pacific Northwest National Laboratory. http://www.pnl.gov/supercriticalfluid/tech_oxidation.stm. Retrieved 2009-06-06.
27. ^ "Rare Earth Oxide Fluoride: Ceramic Nano-particles via a Hydrothermal Method". Pacific Northwest National Laboratory. http://availabletechnologies.pnl.gov/technology.asp?id=152. Retrieved 2009-06-06.
28. ^ M.M. Hoffman, J.S. Young, J.L. Fulton (2000). "Unusual dysprosium ceramic nano-fiber growth in a supercritical aqueous solution". J Mat. Sci. 35: 4177. doi:10.1023/A:1004875413406.
29. ^ Theodore Gray (2009). The Elements. Black Dog and Leventhal Publishers. pp. 152–153. ISBN 978-1-57912-814-2.
30. ^ Dierks, Steve (January 2003). "Dysprosium". Material Safety Data Sheets. Electronic Space Products International. http://www.espi-metals.com/msds's/Dysprosium.htm. Retrieved 2008-10-20.
31. ^ Dierks, Steve (January 1995). "Dysprosium Chloride". Material Safety Data Sheets. Electronic Space Products International. http://www.espi-metals.com/msds's/Dysprosium%20Chloride.htm. Retrieved 2008-11-07.
32. ^ Dierks, Steve (December 1995). "Dysprosium Fluoride". Material Safety Data Sheets. Electronic Space Products International. http://www.espi-metals.com/msds's/Dysprosium%20Fluoride.htm. Retrieved 2008-11-07.
33. ^ Dierks, Steve (November 1988). "Dysprosium Oxide". Material Safety Data Sheets. Electronic Space Products International. http://www.espi-metals.com/msds's/Dysprosium%20Oxide.htm. Retrieved 2008-11-07.


External links

* WebElements.com – Dysprosium
* It's Elemental – Dysprosium

Periodic table
H   He
Li Be   B C N O F Ne
Na Mg   Al Si P S Cl Ar
K Ca Sc   Ti V Cr Mn Fe Co Ni Cu Zn Ga Ge As Se Br Kr
Rb Sr Y   Zr Nb Mo Tc Ru Rh Pd Ag Cd In Sn Sb Te I Xe
Cs Ba La Ce Pr Nd Pm Sm Eu Gd Tb Dy Ho Er Tm Yb Lu Hf Ta W Re Os Ir Pt Au Hg Tl Pb Bi Po At Rn
Fr Ra Ac Th Pa U Np Pu Am Cm Bk Cf Es Fm Md No Lr Rf Db Sg Bh Hs Mt Ds Rg Cn Uut Uuq Uup Uuh Uus Uuo
Alkali metals Alkaline earth metals Lanthanoids Actinoids Transition metals Other metals Metalloids Other nonmetals Halogens Noble gases

Chemistry Index

Retrieved from "http://en.wikipedia.org/"
All text is available under the terms of the GNU Free Documentation License