Chemistry:Ytterbium(III) chloride

From HandWiki
Ytterbium(III) chloride
Ytterbium(III) chloride
Aluminium-trichloride-crystal-3D-balls.png
Names
IUPAC name
Ytterbium(III) chloride
Identifiers
3D model (JSmol)
ChemSpider
EC Number
  • 233-800-5
UNII
Properties
YbCl3
Molar mass 279.40 g/mol
Appearance White powder
Density 4.06 g/cm3 (solid)
Melting point 854 °C (1,569 °F; 1,127 K)[1]
Boiling point 1,453 °C (2,647 °F; 1,726 K)[1]
17 g/100 mL (25 °C)
Structure
Monoclinic, mS16
C12/m1, No. 12
Related compounds
Other anions
Ytterbium(III) oxide
Other cations
Terbium(III) chloride, Lutetium(III) chloride
Except where otherwise noted, data are given for materials in their standard state (at 25 °C [77 °F], 100 kPa).
☒N verify (what is ☑Y☒N ?)
Infobox references

Ytterbium(III) chloride (YbCl3) is an inorganic chemical compound. It reacts with NiCl2 to form a very effective catalyst for the reductive dehalogenation of aryl halides.[2] It is poisonous if injected, and mildly toxic by ingestion. It is an experimental teratogen, known to irritate the skin and eyes.

History

The synthesis of YbCl3 was first reported by Jan Hoogschagen in 1946.[3] It is now a commercially available source of Yb3+ ions and therefore of significant chemical interest.

Chemical properties

The valence electron configuration of Yb+3 (from YbCl3) is 4f135s25p6, which has crucial implications for the chemical behaviour of Yb+3. Also, the size of Yb+3 governs its catalytic behaviour and biological applications. For example, while both Ce+3 and Yb+3 have a single unpaired f electron, Ce+3 is much larger than Yb+3 because lanthanides become much smaller with increasing effective nuclear charge as a consequence of the f electrons not being as well shielded as d electrons.[4] This behavior is known as the lanthanide contraction. The small size of Yb+3 produces fast catalytic behavior and an atomic radius (0.99 Å) comparable to many biologically important ions.[4]

The gas-phase thermodynamic properties of this chemical are difficult to determine because the chemical can disproportionate to form [YbCl6]−3 or dimerize.[5] The Yb2Cl6 species was detected by electron impact (EI) mass spectrometry as (Yb2Cl5+).[5] Additional complications in obtaining experimental data arise from the myriad of low-lying f-d and f-f electronic transitions.[6] Despite these issues, the thermodynamic properties of YbCl3 have been obtained and the C3V symmetry group has been assigned based upon the four active infrared vibrations.[6]

Preparation

Anhydrous ytterbium(III) chloride can be produced by the ammonium chloride route.[7][8][9] In the first step, ytterbium oxide is heated with ammonium chloride to produce the ammonium salt of the pentachloride:

Yb2O3 + 10 NH4Cl → 2 (NH4)2YbCl5 + 6 H2O + 6 NH3

In the second step, the ammonium chloride salt is converted to the trichlorides by heating in a vacuum at 350-400 °C:

(NH4)2YbCl5 → YbCl3 + 2 HCl + 2 NH3

Reactions

YbCl3 is a paramagnetic Lewis acid, like many of the lanthanide chlorides. It gives rise to pseudocontact shifted NMR spectra, akin to NMR shift reagents

Applications in biology

Membrane biology has been greatly influenced by YbCl3, where39K+ and23Na+ ion movement is critical in establishing electrochemical gradients.[10] Nerve signaling is a fundamental aspect of life that may be probed with YbCl3 using NMR techniques. YbCl3 may also be used as a calcium ion probe, in a fashion similar to a sodium ion probe.[11]

YbCl3 is also used to track digestion in animals. Certain additives to swine feed, such as probiotics, may be added to either solid feed or drinking liquids. YbCl3 travels with the solid food and therefore helps determine which food phase is ideal to incorporate the food additive.[12] The YbCl3 concentration is quantified by inductively coupled plasma mass spectrometry to within 0.0009 μg/mL.[4] YbCl3 concentration versus time yields the flow rate of solid particulates in the animal's digestion. The animal is not harmed by the YbCl3 since YbCl3 is simply excreted in fecal matter and no change in body weight, organ weight, or hematocrit levels has been observed in mice.[11]

The catalytic nature of YbCl3 also has an application in DNA microarrays, or so called DNA “chips”.[13] YbCl3 led to a 50–80 fold increase in fluorescein incorporation into target DNA, which could revolutionize infectious disease detection (such as a rapid test for tuberculosis).[13]

References

  1. 1.0 1.1 Walter Benenson; John W. Harris; Horst Stöcker (2002). Handbook of Physics. Springer. p. 781. ISBN 0-387-95269-1. https://books.google.com/books?id=RbLE77b6eRUC&pg=PA781. 
  2. Zhang, Yuankui; Liao, Shijian; Xu, Yun; Yu, Daorong; Shen, Qi (1997). "Reductive Dehalogenation of Aryl Halides by the Nanometric Sodium Hydride Using Lanthanide Chloride as Catalyst". Synth. Commun. 27 (24): 4327–4334. doi:10.1080/00397919708005057. 
  3. Hoogschagen, J. (1946). "The light absorption in the near infra red region of praseodymium, samarium and ytterbium solutions". Physica 11 (6): 513–517. doi:10.1016/S0031-8914(46)80020-X. Bibcode1946Phy....11..513H. 
  4. 4.0 4.1 4.2 Evans, C.H. Biochemistry of the Lanthanides; Plenum: New York, 1990.
  5. 5.0 5.1 Chervonnyi, A.D.; Chervonnaya, N.A. (2004). "Thermodynamic Properties of Ytterbium Chlorides". Russ. J. Inorg. Chem. (Engl. Transl.) 49 (12): 1889–1897. 
  6. 6.0 6.1 Zasorin, E. Z. (1988). "Structure of the rare-earth element trihalide molecules from electron diffraction and spectral data". Russ. J. Phys. Chem. (Engl. Transl.) 62 (4): 441–447.  (Russian language version: Zh. Fiz. Khim. 62(4), pp. 883-895)
  7. Brauer, G., ed (1963). Handbook of Preparative Inorganic Chemistry (2nd ed.). New York: Academic Press. 
  8. Meyer, G. (1989). "The Ammonium Chloride Route to Anhydrous Rare Earth Chlorides—The Example of Ycl 3". The Ammonium Chloride Route to Anhydrous Rare Earth Chlorides-The Example of YCl3. Inorganic Syntheses. 25. pp. 146–150. doi:10.1002/9780470132562.ch35. ISBN 978-0-470-13256-2. 
  9. Edelmann, F. T.; Poremba, P. (1997). Herrmann, W. A.. ed. Synthetic Methods of Organometallic and Inorganic Chemistry. VI. Stuttgart: Georg Thieme Verlag. ISBN 978-3-13-103021-4. 
  10. Hayer, M.K.; Riddell, F.G. (1984). "Shift reagents for 39K NMR". Inorganica Chimica Acta 92 (4): L37–L39. doi:10.1016/S0020-1693(00)80044-4. 
  11. 11.0 11.1 Shinohara, A.; Chiba, M.; Inaba, Y. (2006). "Comparative study of the behavior of terbium, samarium, and ytterbium intravenously administered in mice". Journal of Alloys and Compounds 408–412: 405–408. doi:10.1016/j.jallcom.2004.12.152. 
  12. Ohashi, Y.; Umesaki, Y.; Ushida, K. (2004). "Transition of the probiotic bacteria, Lactobacillus casei strain Shirota, in the gastrointestinal tract of a pig". International Journal of Food Microbiology 96 (1): 61–66. doi:10.1016/j.ijfoodmicro.2004.04.001. PMID 15358506. 
  13. 13.0 13.1 Browne, K.A. (2002). "Metal ion-catalyzed nucleic acid alkylation and fragmentation". Journal of the American Chemical Society 124 (27): 7950–7962. doi:10.1021/ja017746x. PMID 12095339.