Elsevier

Physica C: Superconductivity

Volume 469, Issues 9–12, 1 May–15 June 2009, Pages 370-380
Physica C: Superconductivity

Single crystals of LnFeAsO1−xFx (Ln = La, Pr, Nd, Sm, Gd) and Ba1−xRbxFe2As2: Growth, structure and superconducting properties

https://doi.org/10.1016/j.physc.2009.03.048Get rights and content

Abstract

A review of our investigations on single crystals of LnFeAsO1−xFx (Ln = La, Pr, Nd, Sm, Gd) and Ba1−xRbxFe2As2 is presented. A high-pressure technique has been applied for the growth of LnFeAsO1−xFx crystals, while Ba1−xRbxFe2As2 crystals were grown using a quartz ampoule method. Single crystals were used for electrical transport, structure, magnetic torque and spectroscopic studies. Investigations of the crystal structure confirmed high structural perfection and show incomplete occupation of the (O, F) position in superconducting LnFeAsO1−xFx crystals. Resistivity measurements on LnFeAsO1−xFx crystals show a significant broadening of the transition in high magnetic fields, whereas the resistive transition in Ba1−xRbxFe2As2 simply shifts to lower temperature. The critical current density for both compounds is relatively high and exceeds 2 × 109 A/m2 at 15 K in 7 T. The anisotropy of magnetic penetration depth, measured on LnFeAsO1−xFx crystals by torque magnetometry is temperature dependent and apparently larger than the anisotropy of the upper critical field. Ba1−xRbxFe2As2 crystals are electronically significantly less anisotropic. Point-Contact Andreev-Reflection spectroscopy indicates the existence of two energy gaps in LnFeAsO1−xFx. Scanning Tunneling Spectroscopy reveals in addition to a superconducting gap, also some feature at high energy (∼20 meV).

Introduction

Since the first report on superconductivity at 26 K in F-doped LaFeAsO at the end of February 2008, the superconducting transition temperature has been quickly raised to about 55 K and several new superconductors of a general formula LnFeAsO1−xFx (Ln = La, Ce, Pr, Nd, Sm, Gd, Tb, Dy), abbreviated as Ln1111, have been synthesized [1], [2], [3], [4], [5], [6], [7], [8]. These compounds crystallize with the tetragonal layered ZrCuSiAs structure, in the space group P4/nmm. The structure consists of alternating LnO and FeAs layers, which are electrically charged represented as (LnO)(FeAs)−δ (Fig. 1). Covalent bonding is dominant in the layers, while ionic bonding dominates between layers. Electron carriers can be introduced by substituting F for O or by oxygen deficiency [1], [2], [3], [4], [5], [6], [7], [8]. By substituting Sr2+ for La3+ in La1111, holes are introduced [9].

More recently, superconductivity in AFe2As2 (A = Ca, Sr, Ba) (called A122) with ThCr2Si2-type structure and maximum Tc = 38 K has been reported [10]. These compounds have a more simple crystal structure in which (Fe2As2)-layers, identical to those in Ln1111 are separated by single elemental A layers (Fig. 2). Up to date superconductivity has been found in hole-doped Sr1−xKxFe2As2 and Sr1−xCsxFe2As2 [11], Ca1−xNaxFe2As2 [12], Eu1−xKxFe2As2 [13], and Eu1−xNaxFe2As2 [14], as well as in electron-doped Co-substituted BaFe2As2 [15] and SrFe2As2 [16], and Ni-substituted BaFe2As2 [17]. Furthermore, pressure induced superconductivity has been also discovered in the parent compounds CaFe2As2 [18], [19], SrFe2As2 [20], [21], and BaFe2As2 [21].

Besides KFe2As2 and CsFe2As2, which are superconductors with Tc’s of 3.8 K and 2.6 K [11] respectively, RbFe2As2 is known to exist as well [22]. Therefore, it seemed natural to us to explore the BaFe2As2–RbFe2As2 system in order to search for superconductivity.

It is interesting to explore the important parameters which govern the superconducting properties of new superconductors. In the case of high-Tc cuprates it is the number of carriers doped into the CuO2 layers. In analogy in the pnictides it is the number of carriers doped into the FeAs layers. There are some similarities between the new pnictide superconductors and the cuprate superconductors due to the layered structure and the fact that both Fe and Cu are 3d-elements. However, there are important differences. First, doping on the Fe site in Sr122 or Ba122 by substitution of Fe by Co leads also to the appearance of superconductivity [15], [16] in contrast to cuprates, where substitution for Cu suppresses superconductivity. Second, in cuprates, introducing of one oxygen atom is equivalent to introducing of two fluorine atoms [23]. In Ln1111 there is a significant difference in carrier doping between oxygen deficiency and fluorine substitution. One expects that one oxygen atom deficiency provides two electrons while substitution of F for O2− provides one electron. However, according to [24], [25] oxygen deficiency is much less effective as a source of electrons than F-substitution. Structural parameters play also an important role for obtaining high Tc’s. There is dependence between Tc and the As–Fe–As bond angle of the FeAs4 tetrahedron: maximum Tc is achieved when As–Fe–As bond angle is close to 109.47° corresponding to an ideal tetrahedron [25].

Despite all these differences, the similarities due to the layered crystal structure are important as well. So far, all high-Tc superconductors have a layered crystal structure leading to pronounced anisotropic physical properties. All cuprate superconductors have been characterized by a well-defined effective mass anisotropy parameter γ [26]γ=mc/mab=λc/λab=ξab/ξc=Hc2ab/Hc2c,where mi denote the effective mass, λi the magnetic penetration depth, ξi the coherence length and Hc2IIi the upper critical field in the magnetic field direction i. Nevertheless, the understanding of high-temperature superconductivity was challenged by the observation of two distinctly different and temperature dependent anisotropies in MgB2 single crystals [27], [28], [29]:γλ=λc/λab,γH=Hc2ab/Hc2c.

A straight forward interpretation based on a two-band model was quickly developed, which also lead to a further understanding of the temperature and field dependence of the anisotropy parameters in MgB2, mirroring the complex inter- and intraband mechanism of the two superconducting gaps [30]. For having an overall comparison with the other high-temperature superconductors (e.g. cuprates and MgB2) a detailed knowledge of γ in the oxypnictides is required. Various attempts were made to estimate anisotropy in the oxypnictide superconductors [31], [32], [33], [34], [35], [36], [37], [38], [39], [40], [41], [42], leading to a wide range of results. Nevertheless, from both experimental and theoretical sides there is clear evidence that superconductivity in the pnictides involves more than one-band [31], [32], [33], [43], [44], [45], [46], [47], [48] and therefore it is expected that anisotropy is temperature dependent.

Not only the anisotropic properties are, obviously, best investigated on single crystals. Single crystals are also required for spectroscopic techniques such as Scanning Tunneling Spectroscopy (STS), Angle-Resolved Photoemission Spectroscopy (ARPES), Point-Contact Andreev-Reflection (PCAR) spectroscopy, optical spectroscopy, etc. A number of recent investigations of oxypnictides have focused on the multi-band superconductivity [31], [32], [33], [43], [44], [45], [46], [47], [48]. Answering the question of whether the different Fermi surface sheets are associated with different gaps is of crucial importance in order to identify the mechanism of superconductivity in these compounds. In this regard, experimental techniques such as ARPES, PCAR, and STS are very powerful methods since they allow a direct determination of the energy gap(s). STS is a suitable technique to study this issue since it probes the quasiparticle excitation spectrum in the superconducting state, a direct measure of the local density of states and therefore of the fundamental properties of the superconducting order parameter. This technique was successfully applied to MgB2, where multi-band superconductivity was unambiguously demonstrated through directional STS measurements [49].

We succeeded in growing of the first free standing FeAs-oxypnictide crystals (SmFeAsO1−xFx) using a high pressure technique and NaCl/KCl flux [50]. The NaCl/KCl flux has very low solubility at temperatures below 1000 °C used for processes in quartz ampoules, therefore crystal growth at this temperature is extremely slow [51]. In order to increase the solubility in NaCl/KCl flux for more efficient crystal growth higher temperature should be used, but Ln1111 becomes unstable. This trend can be counteracted by applying high pressure, which then tends to stabilize the structure of Ln1111 at high temperature.

Single crystals of AFe2As2 can be grown from Sn flux, similar to many other intermetallic compounds [52], [53]. Tin is practically the only metal that dissolves iron reasonably well and does not form stable unwanted compounds. Due to high solubility in Sn flux at temperatures compatible with quartz ampoules, large, millimeter-sized crystals of A122 have been grown, which allowed extensive measurements of their physical properties. The disadvantage of the Sn flux technique is that crystals usually contain ∼1 at.% Sn. Another method of growing AFe2As2 crystals is the high-temperature growth from FeAs flux [54].

Here, we report on the crystal growth using both the high-pressure, high-temperature method with NaCl/KCl flux for Ln1111 and the quartz ampoule method with Sn flux for A122 [55]. The results of structure investigations on series of Ln1111 crystals (Ln = Sm, Nd, Pr, La, Gd) and Ba1−xRbxFe2As2 are presented. Electrical resistivity measurements, investigations of the anisotropy parameter and spectroscopic studies are also summarized.

Section snippets

Crystal growth

For the synthesis of LnFeAsO1−xFx (Ln = La, Pr, Nd, Sm, Gd) polycrystalline samples and single crystals we used a cubic anvil high-pressure technique which has been successfully applied in our laboratory at ETH Zurich also for the single crystal growth of MgB2 and other superconductors. The mixture of LnAs, FeAs, Fe2O3, Fe, and LnF3 powders was used as a precursor. For the growth of single crystals we used additionally NaCl/KCl flux. The precursor-to-flux ratio was varied between 1:1 and 1:3. By

Crystal structure of LnFeAs(O, F)

All atomic positions were found using the direct method. The refinement was performed without any constraints. The oxygen and fluorine atoms occupy the same position and were treated as one atom because it is impossible to distinguish between them by X-ray diffraction. The results of the structure refinement are presented in the Table 1. The lanthanide contraction reflects itself in a systematic lattice parameters reduction across the series (Fig. 6, Fig. 7). The only two variable atomic

Conclusions

Single crystals of LnFeAsO1−xFx (Ln = La, Pr, Nd, Sm, Gd) have been grown using a cubic anvil high-pressure technique, and Ba1−xRbxFe2As2 crystals have been grown in quartz ampoules. Superconductivity in the Ba122 compound has been induced by Rb substitution for the first time. The availability of Ln1111 single crystals made it possible to determine several basic superconducting parameters, such as upper critical fields and their anisotropy γH and magnetic penetration depth anisotropy γλ. The

Acknowledgements

This work was supported by the Swiss National Science Foundation, by the NCCR program MaNEP, and partially supported by the Polish Ministry of Science and Higher Education under Research Project for the years 2007–2009 (No. N N202 4132 33). This work was a part of the research program of the Polish National Scientific Network “Materials with strongly correlated electrons”.

References (74)

  • M.R. Eskildsen et al.

    Physica C

    (2003)
  • L. Fang et al.

    J. Cryst. Growth

    (2009)
  • Y. Kamihara et al.

    J. Am. Chem. Soc.

    (2008)
  • H. Takahashi et al.

    Nature

    (2008)
  • X.H. Chen et al.

    Nature

    (2008)
  • G.F. Chen et al.

    Phys. Rev. Lett.

    (2008)
  • Z.-A. Ren et al.

    Europhys. Lett.

    (2008)
  • Z.-A. Ren et al.

    Europhys. Lett.

    (2008)
  • P. Cheng et al.

    Sci. China, Ser.

    (2008)
  • Z.-A. Ren et al.

    Mater. Res. Innov.

    (2008)
  • J.-W.G. Bos et al.

    Chem. Commun.

    (2008)
  • M. Rotter et al.

    Phys. Rev. Lett.

    (2008)
  • K. Sasmal et al.

    Phys. Rev. Lett.

    (2008)
  • G. Wu et al.

    J. Phys. Condens. Matter

    (2008)
  • H.S. Jeevan et al.

    Phys. Rev. B

    (2008)
  • Y. Qi et al.

    New J. Phys.

    (2008)
  • A.S. Sefat et al.

    Phys. Rev. Lett.

    (2008)
  • A. Leithe-Jasper et al.

    Phys. Rev. Lett.

    (2008)
  • L.J. Li et al.

    New J. Phys.

    (2009)
  • T. Park et al.

    J. Phys. Condens. Matter

    (2008)
  • M.S. Torikachvili et al.

    Phys. Rev. Lett.

    (2008)
  • K. Igawa et al.

    J. Phys. Soc. Jpn.

    (2009)
  • P.L. Alireza et al.

    J. Phys.: Condens. Matter.

    (2008)
  • P. Wenz et al.

    B Anorg. Chem. Org. Chem.

    (1984)
  • A.M. Abakumov et al.

    Phys. Rev. Lett.

    (1998)
  • C.-H. Lee et al.

    J. Phys. Soc. Jpn.

    (2008)
  • H. Eisaki et al.

    J. Phys. Soc. Jpn.

    (2008)
  • V.G. Kogan

    Phys. Rev. B

    (1981)
  • M. Angst et al.

    Phys. Rev. Lett.

    (2002)
  • M. Angst et al.
  • J.D. Fletcher et al.

    Phys. Rev. Lett.

    (2005)
  • A. Gurevich

    Phys. Rev. B

    (2003)
  • S. Weyeneth et al.

    J. Supercond. Nov. Magn.

    (2009)
  • L. Balicas, A. Gurevich, Y.-J. Jo, J. Jaroszynski, D.C. Larbalestier, R. H. Liu, H. Chen, X.H. Chen, N.D. Zhigadlo, S....
  • J. Jaroszynski et al.

    Phys. Rev. B

    (2008)
  • A. Dubroka et al.

    Phys. Rev. Lett.

    (2008)
  • L. Shan et al.

    Europhys. Lett.

    (2008)
  • Cited by (125)

    • Mixed anion materials

      2023, Comprehensive Inorganic Chemistry III, Third Edition
    View all citing articles on Scopus
    View full text