Hostname: page-component-84b7d79bbc-5lx2p Total loading time: 0 Render date: 2024-07-30T23:53:51.925Z Has data issue: false hasContentIssue false

Superconductivity Near Ferromagnetism in MgCNi3

Published online by Cambridge University Press:  01 February 2011

H. Rosner
Affiliation:
Department of Physiscs, University of California, Davis CA 95616
R. Weht
Affiliation:
Department de Física, CNEA, Avda. General Paz y constiuyentes, 1650- san Martín, Argentina ICTP, P.O Box 586 34014 Trieste, Italy
M. Johannes
Affiliation:
Department of Physiscs, University of California, Davis CA 95616
W.E. Pickett
Affiliation:
Department of Physiscs, University of California, Davis CA 95616
E. Tosatti
Affiliation:
ICTP, P.O Box 586 34014 Trieste, Italy SISSA and INFM/SISSA, Via Beirut 2-4, 34014 Trieste, Italy
Get access

Abstract

Superconductivity and ferromagnetism have been believed to be incompatible over any extended temperature range until certain specific examples - RuSr2GdCu2O8 and UGe2- have arisen in the past 2-3 years. The discovery of superconductivity above 8 K in MgCNi3, which is primarily the ferromagnetic element Ni and is strongly exchange-enhanced, provides a probable new and different example. This compound is shown here to be near ferromagmnetism, requiring only hole-doping by 12% substitution of Mg by Na or Li. This system will provide the means to prode coupling, and possible coexistence, of these two forms of collective behavior without the requirement of pressure.

Type
Research Article
Copyright
Copyright © Materials Research Society 2002

Access options

Get access to the full version of this content by using one of the access options below. (Log in options will check for institutional or personal access. Content may require purchase if you do not have access.)

References

[1] Nagamatsu, J. et al., Nature 410, 63 (2001).Google Scholar
[2] Uji, S. et al., Nature 410, 907 (2001).Google Scholar
[3] Saxena, S.S. et al., Nature 406, 587 (2000).Google Scholar
[4] Huxley, A. et al., Phys. Rev. B 63, 144519 (2001).Google Scholar
[5] Matzdorf, R. et al., Science 289, 746 (2000).Google Scholar
[6] Shen, K. M. et al., cond-mat/0105487.Google Scholar
[7] He, T. et al., Nature 411, 54 (2001).Google Scholar
[8] We used two well established methods of calculation: WIEN97, See Blaha, P., Schwarz, K., and Luitz, J., Vienna University of Technology, 1997, improved and updated version of the original copyrighted WIEN code, which was published by P. Blaha, K. Schwarz, P. Sorantin, and S. B. Trickey, Comput. Phys. Commun. 59, 399 (1990); a full potential nonorthogonal minimum basis local orbital method described in K. Koepernik and H. Eschrig, Phys. Rev. B 59, 1743 (1999). This latter method gives a somewhat stronger tendency toward magnetism than the WIEN results that we quote in the text.Google Scholar
[9] Our band structure results differ somewhat from those presented by Dugdale, S. B. and Jarlborg, T., cond-mat/0105349, Probably because we have used full potential methods.Google Scholar
[10] Maeno, Y. et al., Nature 372, 532 (1994). I. I. Mazin and D. J. Singh, Phys. Rev. Lett. 79, 733 (1997).Google Scholar
[11] Schwarz, K. and Mohn, P., J. Phys. F 14, L129 (1984).Google Scholar
[12] Marcus, P. M. and Moruzzi, V. L., Phys. Rev. B 38, 6949 (1988).Google Scholar
[13] Mao, Z. Q. et al., cond-mat/0105280.Google Scholar
[14] Since doping simply due to the presence of an interface may occur at a tunnel junction, and its requires only a small amount of hole doping to introduce Ni moments, the effect of magnetic moments at the interface should not be overlooked.Google Scholar
[15] Mathur, N. D. et al., Nature 394, 39 (1998).Google Scholar
[16] Shick, A. B. and Pickett, W. E., Phys. Rev. Lett. 86 300 (2001).Google Scholar
[17] Karchev, N. I. et al., Phys. Rev. Lett. 86, 846 (2001).Google Scholar
[18] Wang, Z. et al., cond-mat/0104097.Google Scholar
[19] Fay, D. and Appel, J., Phys. Rev. B 22, 3173 (1980).Google Scholar
[20] Machida, K. and Ohmi, T., Phys. Rev. Lett. 86, 850 (2001).Google Scholar