- Art Gallery -

Copernicium

Copernicium (pronounced /koʊpərˈnɪsiəm/ koe-pər-NIS-ee-əm, with the first C hard and the second soft) is a synthetic radioactive chemical element with the symbol Cn and atomic number 112. The element was previously known by the IUPAC systematic element name ununbium (pronounced /uːnˈuːnbiəm/ ( listen)[1] oon-OON-bee-əm), with the symbol Uub. It was first created in 1996 by the Gesellschaft für Schwerionenforschung (GSI). It is named after the astronomer Copernicus.

Copernicium is currently the highest-numbered element to be officially recognised by the International Union of Pure and Applied Chemistry (IUPAC). The most stable isotope discovered to date is 285Cn with a half-life of ≈30 s, although evidence exists that 285Cn may have a nuclear isomer with a much longer half-life of 8.9 min. In total, about 75 atoms of copernicium have been detected using various nuclear reactions.[2] Recent experiments suggest that copernicium behaves as a typical member of group 12, demonstrating properties consistent with a volatile metal.[3]

History
Official discovery

Copernicium was first created on February 9, 1996, at the Gesellschaft für Schwerionenforschung (GSI) in Darmstadt, Germany by Sigurd Hofmann, Victor Ninov et al.[4] This element was created by firing accelerated zinc-70 nuclei at a target made of lead-208 nuclei in a heavy ion accelerator. A single atom (the second has subsequently been dismissed) of copernicium was produced with a mass number of 277.[4]

20882Pb + 7030Zn → 278112Cn → 277112Cn + 10n

In May 2000, the GSI successfully repeated the experiment to synthesise a further atom of Cn-277.[5][6] This reaction was repeated at RIKEN using the GARIS set-up in 2004 to synthesise two further atoms and confirm the decay data reported by the GSI team.[7]

The IUPAC/IUPAP Joint Working Party (JWP) assessed the claim of discovery by the GSI team in 2001[8] and 2003.[9] In both cases, they found that there was insufficient evidence to support their claim. This was primarily related to the contradicting decay data for the known isotope 261Rf. However, between 2001 and 2005, the GSI team studied the reaction 248Cm(26Mg,5n)269Hs, and were able to confirm the decay data for 269Hs and 261Rf. It was found that the existing data on 261Rf was for an isomer,[10] now designated 261a Rf.

In May 2009, the JWP reported on the claims of discovery of element 112 again and officially recognised the GSI team as the discoverers of element 112.[11] This decision was based on recent confirmation of the decay properties of daughter nuclei as well as the confirmatory experiments at RIKEN.[12]
Naming

After acknowledging their discovery, the IUPAC asked the discovery team at GSI to suggest a permanent name for ununbium.[13][14] On 14 July 2009, they proposed copernicium with the element symbol Cp, after Nicolaus Copernicus "to honor an outstanding scientist, who changed our view of the world."[15] IUPAC delayed the official recognition of the name, pending the results of a six-month discussion period among the scientific community.[16][17]

Alternative spellings had been suggested to Hofmann, namely "copernicum", "copernium", and "kopernikium" (Kp), and Hofmann has said that the team had discussed the possibility of "copernicum" or "kopernikum", but that they had agreed on "copernicium" in order to comply with current IUPAC rules, which allow only the suffix -ium for new elements.[18][citation needed]

However, it was pointed out that the symbol Cp was previously associated with the name cassiopeium (cassiopium), now known as lutetium (Lu).[19][20] Furthermore, the symbol Cp is also used in organometallic chemistry to denote the ligand cyclopentadiene. For this reason, the IUPAC disallowed the use of Cp as a future symbol, prompting the GSI team to put forward the symbol Cn as an alternative. On February 19, 2010, the 537th anniversary of Copernicus' birth, IUPAC officially accepted the proposed name and symbol.[21][22]
Isotopes and nuclear properties
Nucleosynthesis
Target-projectile combinations leading to Z=112 compound nuclei

The table below contains various combinations of targets and projectiles which could be used to form compound nuclei with Z=112.

Target Projectile CN Attempt result
208Pb 70Zn 278Cn Successful reaction
232Th 50Ti 282Cn Reaction yet to be attempted
238U 48Ca 286Cn Successful reaction
244Pu 40Ar 284Cn Reaction yet to be attempted
248Cm 36S 284Cn Reaction yet to be attempted
249Cf 30Si 279Cn Reaction yet to be attempted


Cold fusion

This section deals with the synthesis of nuclei of copernicium by so-called "cold" fusion reactions. These are processes which create compound nuclei at low excitation energy (~10–20 MeV, hence "cold"), leading to a higher probability of survival from fission. The excited nucleus then decays to the ground state via the emission of one or two neutrons only.

208Pb(70Zn,xn)278-xCn (x=1)

The team at GSI first studied this reaction in 1996 and reported the detection of two decay chains of 277Cn.[4] In a review of the data in 2000, the first decay chain was retracted. In a repeat of the reaction in 2000 they were able to synthesise a further atom. They attempted to measure the 1n excitation function in 2002 but suffered from a failure of the Zn-70 beam. The unofficial discovery of 277Cn was confirmed in 2004 at RIKEN, where researchers detected a further two atoms of the isotope and were able to confirm the decay data for the entire chain.

208Pb(68Zn,xn)276-xCn

After the successful synthesis of 277Cn, the GSI team performed a reaction using a 68Zn projectile in 1997 in an effort to study the effect of isospin (neutron richness) on the chemical yield. The experiment was initiated after the discovery of a yield enhancement during the synthesis of darmstadtium isotopes using 62Ni and 64Ni ions. No decay chains of 275Cn were detected leading to a cross section limit of 1.2 pb. However, the revision of the yield for the 70Zn reaction to 0.5 pb does not rule out a similar yield for this reaction.

184W(88Sr,xn)272-xCn

In 1990, after some early indications for the formation of isotopes of copernicium in the irradiation of a tungsten target with multi-GeV protons, a collaboration between GSI and the University of Jerusalem studied the foregoing reaction. They were able to detect some spontaneous fission activity and a 12.5 MeV alpha decay, both of which they tentatively assigned to the radiative capture product 272Cn or the 1n evaporation residue 271Cn. Both the TWG and JWP have concluded that a lot more research is required to confirm these conclusions.
Hot fusion

This section deals with the synthesis of nuclei of copernicium by so-called "hot" fusion reactions. These are processes which create compound nuclei at high excitation energy (~40–50 MeV, hence "hot"), leading to a reduced probability of survival from fission and quasi-fission. The excited nucleus then decays to the ground state via the emission of 3–5 neutrons. Fusion reactions utilizing 48Ca nuclei usually produce compound nuclei with intermediate excitation energies (~30–35 MeV) and are sometimes referred to as "warm" fusion reactions. This leads, in part, to relatively high yields from these reactions.

238U(48Ca,xn)286-xCn (x=3,4)

In 1998, the team at the Flerov Laboratory of Nuclear Research began a research program using Ca-48 nuclei in "warm" fusion reactions leading to superheavy elements (SHE's). In March 1998, they claimed to have synthesised the element (two atoms) in this reaction. The product, 283Cn, had a claimed half-life of 5 min, decaying by spontaneous fission (SF).[23]

The long lifetime of the product initiated first chemical experiments on the gas phase atomic chemistry of element 112. In 2000, Yuri Yukashev at Dubna repeated the experiment but was unable to observe any spontaneous fission from 5 min activities. The experiment was repeated in 2001 and an accumulation of eight fragments resulting from spontaneous fission were found in the low-temperature section, indicating that copernicium had radon-like properties. However, there is now some serious doubt about the origin of these results.

In order to confirm the synthesis, the reaction was successfully repeated by the same team in Jan 2003, confirming the decay mode and half life. They were also able to calculate an estimate of the mass of the spontaneous fission activity to ~285 lending support to the assignment.[24]

The team at LBNL entered the debate and performed the reaction in 2002. They were unable to detect any spontaneous fission and calculated a cross section limit of 1.6 pb for the detection of a single event.[25]

The reaction was repeated in 2003–2004 by the team at Dubna using a slightly different set-up, the Dubna Gas Filled Recoil Separator (DGFRS). This time, 283Cn was found to decay by emission of a 9.53 MeV alpha-particle with a half-life of 4 seconds. 282Cn was also observed in the 4n channel.[26]

In 2003, the team at GSI entered the debate and performed a search for the five-minute SF activity in chemical experiments. Like the Dubna team, they were able to detect seven SF fragments in the low temperature section. However, these SF events were uncorrelated, suggesting they were not from actual direct SF of copernicium nuclei and raised doubts about the original indications for radon-like properties.[27] After the announcement from Dubna of different decay properties for 283Cn, the GSI team repeated the experiment in September 2004. They were unable to detect any SF events and calculated a cross section limit of ~ 1.6 pb for the detection of one event, not in contradiction with the reported 2.5 pb yield by Dubna.

In May 2005, the GSI performed a physical experiment and identified a single atom of 283Cn decaying by SF with a short lifetime suggesting a previously unknown SF branch.[28] However, initial work by Dubna had detected several direct SF events but had assumed that the parent alpha decay had been missed. These results indicated that this was not the case.

In 2006, the new decay data on 283Cn was confirmed by a joint PSI-FLNR experiment aimed at probing the chemical properties of copernicium. Two atoms of 283Cn were observed in the decay of the parent 287Uuq nuclei. The experiment indicated that contrary to previous experiments, copernicium behaves as a typical member of group 12, demonstrating properties of a volatile metal.[3]

Finally, the team at GSI successfully repeated their physical experiment in Jan 2007 and detected three atoms of 283Cn, confirming both the alpha and SF decay modes.[29]

As such, the 5 min SF activity is still unconfirmed and unidentified. It is possible that it refers to an isomer, namely 283bCn, whose yield is dependent upon the exact production methods.

233U(48Ca,xn)281-xCn

The team at FLNR studied this reaction in 2004. They were unable to detect any atoms of element 112 and calculated a cross section limit of 600 fb. The team concluded that this indicated that the neutron mass number for the compound nucleus had an effect on the yield of evaporation residues.[26]
As a decay product

Copernicium has also been observed as decay products of elements 114, 116, and 118 (see ununoctium).
Evaporation Residue Observed Cn isotope

Evaporation Residue Observed Cn isotope
293Uuh, 289Uuq 285Cn
292Uuh, 288Uuq 284Cn
291Uuh, 287Uuq 283Cn
294Uuo, 290Uuh, 286Uuq 282Cn

As an example, in May 2006, the Dubna team (JINR) identified 282Cn as a final product in the decay of ununoctium via the alpha decay sequence.

294118Uuo → 290116Uuh → 286114Uuq → 282112Cn

It was found that the final nucleus undergoes spontaneous fission.[30]
Retracted isotopes
281Cn

In the claimed synthesis of 293Uuo in 1999 (see ununoctium) the isotope 281Cn was identified as decaying by emission of a 10.68 MeV alpha particle with half-life 0.90 ms. The claim was retracted in 2001 and hence this copernicium isotope is currently unknown or unconfirmed.
Chronology of isotope discovery

Isotope Year discovered discovery reaction
277Cn 1996 208Pb(70Zn,n)
278Cn unknown
279Cn unknown
280Cn unknown
281Cn unknown
282Cn 2004 238U(48Ca,4n)
283Cn 2002 244Pu(48Ca,5n)
283bCn ?? 1998 238U(48Ca,3n)
284Cn 2002 244Pu(48Ca,4n)
285Cn 1999 244Pu(48Ca,3n)
285bCn ? 1999 244Pu(48Ca,3n)


Nuclear isomerism

285a,bCn

In the synthesis of 289Uuq and 293Uuh, a 8.63 MeV alpha-decaying activity has been detected with a half-life of 8.9 minutes. Although unconfirmed in recent experiments, it is highly possible that this is associated with an isomer, namely 285bCn.

283a,bCn

First experiments on the synthesis of 283Cn produced a SF activity with half-life ~5 min. This activity was also observed from the alpha decay of 287Uuq. The decay mode and half-life were also confirmed in a repetition of the first experiment. However, more recently,283Cn has been observed to undergo 9.52 MeV alpha decay and SF with a half-life of 3.9 s. These results suggest the assignment of the two activities to two different isomeric levels in 283Cn, creating 283aCn and 283bCn. Further research is required to address these discrepancies.
Chemical yields of isotopes
Cold fusion

The table below provides cross-sections and excitation energies for cold fusion reactions producing copernicium isotopes directly. Data in bold represent maxima derived from excitation function measurements. + represents an observed exit channel.

Projectile Target CN 1n 2n 3n
70Zn 208Pb 278Cn 0.5 pb, 10.0, 12.0 MeV
68Zn 208Pb 276Cn <1.2 pb, 11.3, 12.8 MeV


Hot fusion

The table below provides cross-sections and excitation energies for hot fusion reactions producing copernicium isotopes directly. Data in bold represents maxima derived from excitation function measurements. + represents an observed exit channel.

Projectile Target CN 3n 4n 5n
48Ca 238U 286Cn 2.5 pb, 35.0 MeV 0.6 pb
48Ca 233U 281Cn <0.6 pb, 34.9 MeV


Fission of compound nuclei with Z=112

Several experiments have been performed between 2001 and 2004 at the Flerov Laboratory of Nuclear Reactions in Dubna studying the fission characteristics of the compound nucleus 286Cn. The nuclear reaction used is 238U+48Ca. The results have revealed how nuclei such as this fission predominantly by expelling closed shell nuclei such as 132Sn (Z=50, N=82). It was also found that the yield for the fusion-fission pathway was similar between 48Ca and 58Fe projectiles, indicating a possible future use of 58Fe projectiles in superheavy element formation.[31]
Theoretical calculations
Evaporation residue cross sections

The below table contains various targets-projectile combinations for which calculations have provided estimates for cross section yields from various neutron evaporation channels. The channel with the highest expected yield is given.

DNS = Di-nuclear system; σ = cross section

Target Projectile CN Channel (product) σmax Model Ref
208Pb 70Zn 278Cn 1n (277Cn) 1.5 pb DNS [32]
208Pb 67Zn 275Cn 1n (274Cn) 2 pb DNS [32]
238U 48Ca 286Cn 4n (282Cn) 0.2 pb DNS [33]
244Pu 40Ar 284Cn 4n (280Cn) 0.1 pb DNS [33]
250Cm 36S 286Cn 4n (282cn) 5 pb DNS [33]
252Cf 30Si 282Cn 3n (279Cn) 10 pb DNS [33]


Chemical properties
Extrapolated chemical properties
Oxidation states

Copernicium is the last member of the 6d series of transition metals and the heaviest member of group 12 (IIB) in the Periodic Table, below zinc, cadmium and mercury. Each of the members of this group show a stable +2 oxidation state. In addition, mercury(I), Hg2+2, is also well known. Copernicium is therefore expected to form a stable +2 state.
Chemistry

The known members of group 12 all react with oxygen and sulfur directly to form the oxides and sulfides, MO and MS, respectively. Mercury(II) oxide, HgO, can be decomposed by heat to the liquid metal. Mercury also has a well known affinity for sulfur. Therefore, copernicium should form an analogous oxide CnO and sulfide CnS.

In their halogen chemistry, all the metals form the ionic difluoride MF2 upon reaction with fluorine. The other halides are known but for mercury, the soft nature of the Hg(II) ion leads to a high degree of covalency and HgCl2, HgBr2 and HgI2 are low-melting, volatile solids. Therefore, copernicium is expected to form an ionic fluoride, CnF2, but volatile halides, CnCl2, CnBr2 and CnI2.

In addition, mercury is well known for its alloying properties, with the concomitant formation of amalgams, especially with gold and silver. It is also a volatile metal and is monatomic in the vapour phase. Copernicium is therefore also predicted to be a volatile metal which readily combines with gold to form a Au-Cn metal-metal bond.

Experimental chemistry
Atomic gas phase

Copernicium has the ground state electron configuration [Rn]5f14 6d10 7s2 and thus belongs to group 12 of the Periodic Table. As such, it should behave as the heavier homologue of mercury (Hg) and form strong binary compounds with noble metals like gold. Experiments probing the reactivity of copernicium have focused on the adsorption of atoms of element 112 onto a gold surface held at varying temperatures, in order to calculate an adsorption enthalpy. Due to possible relativistic stabilisation of the 7s electrons, leading to radon-like properties, experiments were performed with the simultaneous formation of mercury and radon radioisotopes, allowing a comparison of adsorption characteristics.

The first experiments were conducted using the 238U(48Ca,3n)283Cn reaction. Detection was by spontaneous fission of the claimed 5 min parent isotope. Analysis of the data indicated that copernicium was more volatile than mercury and had noble-gas properties. However, the confusion regarding the synthesis of 283Cn has cast some doubt on these experimental results.

Given this uncertainty, between April-May 2006 at the JINR, a FLNR-PSI team conducted experiments probing the synthesis of this isotope as a daughter in the nuclear reaction 242Pu(48Ca,3n)287Uuq. In this experiment, two atoms of 283Cn were unambiguously identified and the adsorption properties indicated that copernicium is a more volatile homologue of mercury, due to formation of a weak metal-metal bond with gold, placing it firmly in group 12.

In April 2007 this experiment was repeated and a further three atoms of 283Cn were positively identified. The adsorption property was confirmed and indicated that copernicium has adsorption properties completely in agreement with being the heaviest member of group 12.[34]
See also

* Island of stability

References

1. ^ J. Chatt (1979). "Recommendations for the Naming of Elements of Atomic Numbers Greater than 100". Pure Appl. Chem. 51: 381–384. doi:10.1351/pac197951020381.
2. ^ See references in this article relating to 277Cn, 282Cn and 283Cn, as well as references in ununquadium, ununhexium and ununoctium regarding observed daughter nuclei
3. ^ a b Eichler, R; Aksenov, NV; Belozerov, AV; Bozhikov, GA; Chepigin, VI; Dmitriev, SN; Dressler, R; Gäggeler, HW et al. (2007). "Chemical Characterization of Element 112". Nature 447 (7140): 72–75. doi:10.1038/nature05761. PMID 17476264.
4. ^ a b c S. Hofmann, et al. (1996). "The new element 112". Zeitschrift für Physik: A Hadrons and Nuclei 354 (1): 229–230. doi:10.1007/BF02769517.
5. ^ Hofmann et al. (2002). "New Results on Element 111 and 112". European Physical Journal A Hadrons and Nuclei 14 (2): 147–57. doi:10.1140/epja/i2001-10119-x.
6. ^ Hofmann et al. (2000). "New Results on Element 111 and 112". GSI Scientific Report 2000. http://www.gsi.de/informationen/wti/library/scientificreport2000/Nuc_St/7/ar-2000-z111-z112.pdf.
7. ^ K. Morita (2004). "Decay of an Isotope 277112 produced by 208Pb + 70Zn reaction". Proceedings of the International Symposium. Exotic Nuclei (EXON2004). World Scientific. pp. 188–191. doi:10.1142/9789812701749_0027.
8. ^ Karol, P. J; Nakahara, H; Petley, B. W; Vogt, E (2001). "On the Discovery of the Elements 110–112" (IUPAC Technical Report). Pure Appl. Chem. 73 (6): 959–967. doi:10.1351/pac200173060959. http://www.iupac.org/publications/pac/2001/pdf/7306x0959.pdf.
9. ^ Karol, P. J; Nakahara, H; Petley, B. W; Vogt, E (2003). "On the Claims for Discovery of Elements 110, 111, 112, 114, 116 and 118" (IUPAC Technical Report). Pure Appl. Chem. 75 (10): 1061–1611. doi:10.1351/pac200375101601. http://www.iupac.org/publications/pac/2003/pdf/7510x1601.pdf.
10. ^ R. Dressler; A. Türler (2001). "Evidence for Isomeric States in 261Rf". Annual Report 2001. Paul Scherrer Institute. http://lch.web.psi.ch/pdf/anrep01/B-02heavies.pdf.
11. ^ [1][dead link]
12. ^ Barber, R.C; Gaeggeler, H.W; Karol, P.J; Nakahara, H; Vardaci, E; Vogt, E (2009). "Discovery of the element with atomic number 112" (IUPAC Technical Report). Pure Appl. Chem. 81: 1331. doi:10.1351/PAC-REP-08-03-05. http://media.iupac.org/publications/pac/asap/pdf/PAC-REP-08-03-05.pdf.
13. ^ "New Chemical Element In The Periodic Table". www.sciencedaily.com. http://www.sciencedaily.com/releases/2009/06/090611210039.htm.
14. ^ Barber, Robert C.; Gäggeler; Karol; Nakahara; Vardaci; Vogt (2009). "Discovery of the element with atomic number 112 (IUPAC Technical Report)". Pure and Applied Chemistry 81: 1331. doi:10.1351/PAC-REP-08-03-05.
15. ^ "Element 112 shall be named "copernicium"". www.gsi.de. July 14, 2009. http://www.gsi.de/portrait/Pressemeldungen/14072009_e.html.
16. ^ New element named 'copernicium', BBC News, Thu 16 July 2009
17. ^ "News: Start of the Name Approval Process for the Element of Atomic Number 112". IUPAC. http://www.iupac.org/web/nt/2009-07-21_Naming_Element_112.
18. ^ private email from Hofmann
19. ^ Meija, J (2009). "The need for a fresh symbol to designate copernicium". Nature 461 (7262): 341. doi:10.1038/461341c. PMID 19759598.
20. ^ "71. Lutetium - Elementymology & Elements Multidict". Elements.vanderkrogt.net. http://elements.vanderkrogt.net/element.php?sym=Lu. Retrieved 2010-02-22.
21. ^ "Science & Environment | New element named 'copernicium'". BBC News. 2009-07-16. http://news.bbc.co.uk/2/hi/science/nature/8153596.stm. Retrieved 2010-02-22.
22. ^ "[IUPAC]Element 112 is Named Copernicium". iupac.org. doi:10.1351/PAC-REP-08-03-05. http://www.iupac.org/web/nt/2010-02-20_112_Copernicium. Retrieved 2010-02-22.
23. ^ Oganessian et al. (1999). "Search for new isotopes of element 112 by irradiation of 238U with 48Ca". Eur. Phys. J. A 5 (1): 63–68. doi:10.1007/s100500050257.
24. ^ Yu Ts Oganessian et al. (2004). "Second Experiment at VASSILISSA separator on the synthesis of the element 112". Eur. Phys. J. A 19 (1): 3–6. doi:10.1140/epja/i2003-10113-4.
25. ^ W. Loveland, K. E. Gregorich, J. B. Patin, D. Peterson, C. Rouki, P. M. Zielinski, and K. Aleklett (2002). "Search for the production of element 112 in the 48Ca+238U reaction". Phys. Rev. C 66 (4): 044617. doi:10.1103/PhysRevC.66.044617.
26. ^ a b Yu. Ts. Oganessian et al. (2004). "Measurements of cross sections and decay properties of the isotopes of elements 112, 114, and 116 produced in the fusion reactions 233,238U, 242Pu, and 248Cm+48Ca". Phys. Rev. C 70: 064609. doi:10.1103/PhysRevC.70.064609.
27. ^ S. Soverna (2003). Indication for a gaseous element 112. 2003. GSI Scientific Report. p. 187. http://www.gsi.de/informationen/wti/library/scientificreport2003/files/167.pdf.
28. ^ S. Hofmann, et al. (2005). Search for Element 112 Using the Hot Fusion Reaction 48Ca + 238U. 2005. GSI Scientific Report. p. 191. http://www.gsi.de/informationen/wti/library/scientificreport2005/PAPERS/NUSTAR-SHE-PHYS-01.pdf.
29. ^ S. Hofmann et al. (2007). "The reaction 48Ca + 238U -> 286112* studied at the GSI-SHIP". Eur. Phys. J. A 32 (3): 251–260. doi:10.1140/epja/i2007-10373-x.
30. ^ Oganessian, Yu. Ts.; et al. (2006). "Synthesis of the isotopes of elements 118 and 116 in the 249Cf and 245Cm+48Ca fusion reactions". Physical Review C 74 (4): 044602. doi:10.1103/PhysRevC.74.044602.
31. ^ see Flerov lab annual reports 2001–2004
32. ^ a b Feng, Zhao-Qing (2007). "Formation of superheavy nuclei in cold fusion reactions". Physical Review C 76: 044606. doi:10.1103/PhysRevC.76.044606. http://arxiv.org/pdf/0707.2588.
33. ^ a b c d Influence of entrance channels on formation of superheavy nuclei in massive fusion reactions. http://arxiv.org/pdf/0904.2994.
34. ^ H. W. Gäggeler (2007). "Gas Phase Chemistry of Superheavy Elements". Paul Scherrer Institute. http://lch.web.psi.ch/files/lectures/TexasA&M/TexasA&M.pdf.

External links

* WebElements.com: Copernicium

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

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