Next Article in Journal
Microdrop Deposition Technique: Preparation and Characterization of Diluted Suspended Particulate Samples
Next Article in Special Issue
Fractal Dimension Analysis of High-Resolution X-Ray Phase Contrast Micro-Tomography Images at Different Threshold Levels in a Mouse Spinal Cord
Previous Article in Journal / Special Issue
Annealed Low Energy States in Frustrated Large Square Josephson Junction Arrays
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

On the Nanoscale Structure of KxFe2−yCh2 (Ch = S, Se): A Neutron Pair Distribution Function View

by
Panagiotis Mangelis
1,*,
Hechang Lei
2,†,
Marshall T. McDonnell
3,
Mikhail Feygenson
3,‡,
Cedomir Petrovic
2,
Emil S. Bozin
2 and
Alexandros Lappas
1
1
Institute of Electronic Structure and Laser, Foundation for Research and Technology—Hellas, Vassilika Vouton, 711 10 Heraklion, Greece
2
Condensed Matter Physics and Materials Science Department, Brookhaven National Laboratory, Upton, NY 11973, USA
3
Neutron Scattering Division, Oak Ridge National Laboratory, Oak Ridge, TN 37831, USA
*
Author to whom correspondence should be addressed.
Present Address: Department of Physics, Renmin University, Beijing 100872, China.
Present Address: Forschungszentrum Jülich, JCNS, D-52425 Jülich, Germany.
Condens. Matter 2018, 3(3), 20; https://doi.org/10.3390/condmat3030020
Submission received: 3 June 2018 / Revised: 26 June 2018 / Accepted: 27 June 2018 / Published: 3 July 2018

Abstract

:
Comparative exploration of the nanometer-scale atomic structure of KxFe2−yCh2 (Ch = S, Se) was performed using neutron total scattering-based atomic pair distribution function (PDF) analysis of 5 K powder diffraction data in relation to physical properties. Whereas KxFe2−ySe2 is a superconductor with a transition temperature of about 32 K, the isostructural sulphide analogue is not, which instead displays a spin glass semiconducting behavior at low temperatures. The PDF analysis explores phase separated and disordered structural models as candidate descriptors of the low temperature data. For both materials, the nanoscale structure is well described by the iron (Fe)-vacancy-disordered K2Fe5−yCh5 (I4/m) model containing excess Fe. An equally good description of the data is achieved by using a phase separated model comprised of I4/m vacancy-ordered and I4/mmm components. The I4/mmm component appears as a minority phase in the structure of both KxFe2−ySe2 and KxFe2−yS2, and with similar contribution, implying that the phase ratio is not a decisive factor influencing the lack of superconductivity in the latter. Comparison of structural parameters of the Fe-vacancy-disordered model indicates that the replacement of selenium (Se) by sulphur (S) results in an appreciable reduction in the Fe-Ch interatomic distances and anion heights, while simultaneously increasing the irregularity of FeCh4 tetrahedra, suggesting the more significant influence of these factors. Structural features are also compared to the non-intercalated FeSe and FeS parent phases, providing further information for the discussion about the influence of the lattice degrees of freedom on the observed properties in layered iron chalcogenides.

1. Introduction

Binary Fe1+ySe [1] has attracted considerable attention [2] due to its critical superconducting transition temperature (TC) up to 8 K at ambient pressure. Externally applied pressure increases the TC to 37 K at 7 GPa [3], whereas single FeSe layers grown on SrTiO3 show strikingly higher TCs (ca. 65 K) [4,5] than many other Fe-based superconductors. Their crystal structure is quasi two-dimensional, composed of slabs of FeSe edge-sharing tetrahedra held together by van der Waals interactions. Isovalent substitution with sulphur or tellurium for selenium does not change the carrier density [6,7], but substantially modifies the regularity of FeCh4 (Ch = Se, S, Te) tetrahedron, which is believed to be a crucial factor governing electronic and magnetic properties, having particular influence on electron pairing. Consequently, the chemistry and physics of these simple iron-based materials result in complex relationships in the Fe1+yTe1−xSex series [8].
The transition between itinerant antiferromagnetism and superconductivity can be controlled at elevated pressures, but also at ambient pressure by altering the crystal structure through substitutions either at the iron or chalcogen sites and favorably by intercalation of electron donating spacers (guests) between the FeCh sheets. Typically, alkali metals inserted between the FeCh layers expand the lattice and result in a ternary family of iron-superconductors, AxFe2Se2 (A = K, Rb, Cs or Tl), with TCs in the order of 30 K. The first example of these so-called 122-iron selenides, the K0.8Fe2Se2 derivative [9,10], was shown to adopt a ThCr2Si2-type structure. Much like the iron-pnicitides, with Fe tetrahedrally coordinated, superconductivity remains at the boundary with antiferromagnetism. However, whereas pnictides display superconductivity in the fully occupied, stoichiometric ThCr2Si2 structure, divalent Se2− (instead of trivalent As3−) leads to the presence of alkali and iron vacancies and issues of chemical instability that is promoted by the high-temperature solid-state synthesis of the KxFe2−ySe2 compounds. This entails microscale phase separation, with heterogeneity creating less than 20% of I4/mmm KxFe2Se2, adopting the iron-pnictide ThCr2Si2 superconducting crystal structure, and the majority I4/m K2Fe4Se5 Fe-vacancy-ordered antiferromagnetic phase (Figure 1) [11,12]. The origin of superconductivity and the precise stoichiometry of iron and potassium content that leads to the superconducting phase are still under debate owing to the intrinsic phase separation and inhomogeneity in this type of chemically complex material [13,14]. Interestingly, the substitution of Se for S in the KxFe2−ySe2−zSz (0 ≤ z ≤ 2) series suppresses the superconducting state (z = 1.6) [15] and eventually the sulphide end-member, KxFe2−yS2, is not superconducting but exhibits a spin-glass behavior at temperatures below 32 K [16]. Therefore, experimental efforts to understand superconductivity in iron chalcogenides must not only tackle the sensitive stoichiometry required to observe a sizeable TC, but also decipher the subtle differences in the atomic structure configuration that lead to such markedly different physical properties.
Here, we use neutron total scattering experiments combined with atomic pair distribution function (PDF) analysis [17] to explore the nanometer scale aspects of phase separation and inhomogeneity for the two end-members of KxFe2−ySe2−zSz (z = 0, 2) series, at 5 K. We probe their structural evolution based on the regularity of the FeCh4 tetrahedra and the Fe anion heights, which have been considered before as empirical parameters that mediate the evolution of the superconducting critical temperature in these intercalated materials.

2. Materials and Methods

Powdered samples of KxFe2−ySe2 and KxFe2−yS2 were prepared by pulverizing single crystals stored in a glove box. The single crystals were synthesized and characterized according to protocols reported previously [15]. Laboratory X-ray powder diffraction data of the samples used in this study were fully indexed within the I4/m space group. Low temperature magnetic susceptibility and electrical resistivity measurements of the materials were previously reported [15]. The selenium-containing sample exhibited bulk superconductivity at TC = 33 K, whereas the sulfur-containing sample was not superconducting and displayed magnetic semi-metallic behavior at low temperatures.
The neutron total scattering experiment was performed at the Nanoscale-Ordered Materials Diffractometer (NOMAD) [18] at the Spallation Neutron Source at Oak Ridge National Laboratory. Two pulverized samples, approximately 0.5 g each, were sealed in helium atmosphere in extruded cylindrical vanadium containers 6 mm in diameter, and mounted in the diffractometer equipped with an Orange cryostat sample environment. The instrument was calibrated using diamond standard. The data for each sample were collected at 5 K for 60 min. Data reduction was performed following standard protocols described elsewhere [19]. Pair distribution functions were obtained by Sine Fourier transformation of the measured reduced total scattering functions F(Q), where Q is the momentum transfer, over a range of QMIN = 0.5 Å−1 to QMAX = 26 Å−1. Refinement of the structural models against the experimental PDF data was completed over a 1 nm range via the small-box approach used within the PDFGUI software suite [20].

3. Results and Discussion

3.1. Qualitative Data Comparison

We began by observing and comparing the raw neutron powder diffraction data for KxFe2−ySe2 and KxFe2−yS2 shown in Figure 2, as seen by the forward scattering bank of detectors. Notably, this detector bank provided rather poor resolution featuring broad peaks and inherently lower statistics. This is a consequence of the construction of the NOMAD instrument, which is optimized for total scattering type measurements, where broad Q-range coverage and high resolution in backscattering banks are far more important, at the expense of the low angle banks. Nevertheless, for both, the data clearly identify the presence of, albeit weak, intensity associated with the (110) reflection that is a hallmark of I4/m and is absent in I4/mmm symmetry. This qualitative assessment revealed that both systems have, at least in part, I4/m structural character, in agreement with previous reports [15,16].
The neutron total scattering data collected over a broad range of momentum transfer Q were examined next. The findings are illustrated in Figure 3a,b for KxFe2−ySe2 and KxFe2−yS2, respectively, in the form of reduced total scattering structure function F(Q). Observable are appreciable arcs of diffuse scattering, particularly at higher momentum transfer values, indicative of appreciable disorder, as expected for this class of materials. Fourier transforms of F(Q) resulted in PDF and G(r), that are presented in Figure 3c,d, respectively. Some of the PDF peaks appear more intense in the KxFe2−ySe2 pattern than in that of KxFe2−yS2. In principle, this could be due to the differences in the underlying bond-length distributions reflecting the nanoscale structures and the complexity of underlying disorder. However, the coherent neutron scattering lengths, which scale the individual PDF contributions of any given pair of atoms in the structure, are nearly three times larger for selenium than for sulfur (b(Se) = 7.970 fm, whereas b(S) = 2.847 fm). The change in relative PDF peak intensity also originates from the renormalization due to different scattering properties of the samples.
The PDFs of the two samples studied were directly compared in Figure 3. Notably, the PDF pattern of the sulphide is shifted considerably to lower interatomic distances, r, as indicated by the block arrow in Figure 3d, compared to the selenide counterpart. This lattice contraction occurred due to the lower ionic size of S2− (ionic radius 1.84 Å) compared to that of Se2− (ionic radius 1.98 Å) [21]. By examining the very local scale (Figure 3c), observably narrower bond-length distribution pertaining to the FeCh4 tetrahedral environments was apparent in KxFe2−ySe2 compared with KxFe2−yS2. However, considering the appreciably different neutron scattering lengths of Se and S, as well as their effective radii, this assertion requires further confirmation from explicit structural modeling.

3.2. Model Dependent PDF Analysis

Structural information from experimental PDFs was extracted using the small box modeling approach based on small symmetrized unit cells (typically ~10 to 100 atoms) [17]. This approach provides a simple framework for crystal structure interpretation at the expense of being limited to the symmetry constraints of the underlying models. The approach was previously reported in detail [20].
Our modeling was based on the two phases described above and shown in Figure 1. The phase with I4/mmm symmetry, often referred to as metallic, features stoichiometric slabs of FeCh4 tetrahedra, whereas the phase with I4/m symmetry, considered insulating and antiferromagnetic, contains two symmetrically distinct Fe crystallographic sites, one of which features Fe-vacancies and is completely empty in the case of full vacancy order. Here, we refer to the latter as the Fe-vacancy-ordered (I4/m VO) model. In this model, the Fe and K occupancies at the low multiplicity sites, 4d and 2b, respectively, were fixed to zero under the assumption that all Fe and K atoms reside only at the high multiplicity sites, 16i and 8h, consistent with our room temperature Rietveld refinements reported earlier [22]. In addition, we considered a modified version of the I4/m model, in which the occupancy of the Fe-vacancy 4d site was allowed to vary, which we refer to as the vacancy-disordered (I4/m VD) model. We used these atomic configurations both individually in single phase fits and combined in two-phase refinements. All fractional coordinates were constrained to follow the respective space group symmetries. All atomic displacement parameters (ADP) were set to be isotropic and identical for all atomic species of the same type. The occupancy and ADP constraints were introduced to minimize the number of refined parameters and to enable the sensible semiquantitative comparison of the two systems within the realm of these models. Although various model combinations were tested, the two that showed the best agreement with the data were (1) two-phase mixture composed of I4/mmm and I4/m VO ingredient phases, and (2) single phase I4/m VD model. We limit our discussion to these two models.
The refinements of the structural models that result in the lowest fit residual (RW) are summarized in Figure 4a for selenide and in Figure 4b for sulphide. The refinements results revealed that, at the nanoscale, the structures of both KxFe2−ySe2 and KxFe2−yS2 within the sensitivity of our data, were consistent with both a two-phase mixture (metallic I4/mmm and insulating I4/m VO) model and with a single phase disordered I4/m VD model. Previous reports confirmed the coexistence of these two phases in both selenide [22] and sulphide analogue [23]. Notably, the single phase I4/m VD model is effectively equivalent to a phase separated model, albeit disordered. In the explicit two-phase modeling of the selenide and sulphide data, the majority phase was found to be the I4/m VO with a percentage of ca. 89(5)% and ca. 73(8)% by weight, respectively. Notably, according to this step in the analysis, the minority phase I4/mmm in the case of KxFe2−yS2 exhibited a higher percentage in the phase mixture than KxFe2−ySe2. Whereas the exact phase percentage is difficult to acquire and is somewhat dependent on the type of constraints used, the robust observation in all our fits was that the metallic I4/mmm phase contribution in non-superconducting KxFe2−yS2 was at least comparable or probably slightly larger than in superconducting KxFe2−ySe2. This implied that the presence or absence of this phase is not the sole factor that affects the superconductivity in this system. Furthermore, the phase separation picture on the nanoscale likely represents an oversimplification of the true complexity of the system, and may possibly reflect the limitations of the small box modeling approach.
We next explored the vacancy-disordered K2Fe5−yCh5 I4/m VD model in which the Fe occupancy at the 4d site is explicitly refined. The Fe occupancy at the high multiplicity 16i site was fixed to one, as an attempt to explicitly refine the marginal deviations, in agreement with previous reports. Although attempts to refine K-content were made, this resulted in fit instabilities and the correlation of refined parameters was indicative of over-parametrization, so the occupancy of K sites was adjusted in all models to reflect nominal 0.8 potassium content. In I4/m-type models, the K 2b site was empty. Importantly, the refinement results were qualitatively very similar to those of the phase mixture model. The refinements show finite and similar Fe occupancies at the 4d site in both materials, which increase the overall Fe content to K2Fe4.27(1)Se5 and K2Fe4.27(1)S5, indicating no substantial difference across the two systems in terms of phase-separation sensitive parameters and within the limitations of the applied modeling methodology. We interpreted this as an indicator that the phase separation aspect is not of fundamental importance for the lack of superconductivity in the sulphide, in the sense that non-superconducting sulphide does not account for a smaller fraction of the metallic phase.
In Table 1 and Table 2, we quantitatively compare various structural parameters derived from the single-phase I4/m VD model at 5 K, such as lattice parameters, interatomic Fe-Ch and Fe-Fe distances, anion heights, and ADP values. Most values are compared in reference to respective quantities obtained for non-intercalated FeSe [24] and FeS [25] parent materials. As expected, the lattice parameters and the interatomic distances (all Fe-Ch and the inter-cluster Fe-Fe lengths) are longer in the case of selenide, resulting in a large increase in the unit cell volume of ca. 11%. In the K-intercalated systems, two types of Fe-Fe distances are associated with the square Fe clusters: intracluster and intercluster, as defined within the I4/m model. In the FeSe reference, there are two distinct Fe-Fe distances; however, these are along the two distinct crystallographic directions, as defined by the orthorhombic Cmma model.
One prominent structural parameter known to be correlated with the superconducting transition temperature TC in iron-based superconductors is the anion height, defined as the c-axis normal distance between the chalcogen/pnictogen and the Fe-sheet. According to this empirical observation, the critical temperature, TC increases as the anion height approaches a value of ~1.38 Å, inferring the existence of an optimal bonding environment for maximizing the superconducting response [26]. The anion height values obtained for the samples studied here were generally consistent with the empirical rule established by Mizuguchi et al. [26]. For our superconducting K2Fe4.27(1)Se5 sample, with TC ≈ 32 K, the anion height was indeed relatively close to this ideal value, whereas for the insulating K2Fe4.27(1)S5 material, the anion height was appreciably further away from the optimal 1.38 Å. This is illustrated schematically in the inset of Figure 5. The replacement of Se by smaller S results in reduction of all interatomic cation-anion distances, which in turn gives rise to a decrease in the anion height to about ~1.28 Å. The anion height in the potassium-intercalated selenide can also be compared to the parent FeSe, which crystallizes in the Cmma structure at low temperatures. For the latter, the neutron powder diffraction data collected at 10 K on the same instrument were used [24]. Intercalation of K in between the FeSe layers induced an effective reduction in the anion height, shifting the value closer to the optimum, while simultaneously increasing the Fe-interlayer spacing from 5.316(1) Å to 7.001(1) Å, with a concomitant increase of TC from about 8 K to about 32 K. In the sulphide analogue, the K intercalation also increased the Fe interlayer spacing from 5.031(1) Å to 6.747(4) Å, but the anion height increased only slightly, and at 1.28 Å, is still substantially far from the optimum value for superconductivity. Whereas FeS is a superconductor with a TC ~5 K [25], the K-intercalated compound exhibited a spin-glass behavior.
Lastly, we considered the regularity of the FeCh4 tetrahedron, which represents yet another structural aspect of relevance to superconductivity [12,15,27]. Indeed, the superconducting critical temperature TC in some iron-based superconductors was maximized for structures featuring regular tetrahedral units, defined by the Ch-Fe-Ch angles all being equal to 109.47°, thus resulting in perfect tetrahedral geometry. Given this, the mean and standard deviation of the underlying bond angle distributions were estimated from the neutron scattering data analyzed for the systems of interest here. To facilitate the comparison, the data are schematically presented as Gaussian distributions in Figure 5, whose centroids reflect the average tetrahedral angle, whereas the widths represent the standard deviation of the angular distribution. For the samples studied, these values were obtained for the six tetrahedral angles of K2Fe5−yCh5 (Ch = Se, S) based on the I4/m VD model for 16i Fe site, and the three angles of FeSe (with multiplicity two) assuming an orthorhombic Cmma crystal type. In the case of FeS, the distribution was the narrowest (not shown), implying that the tetrahedral regularity is not a required condition for optimizing superconductivity. Nevertheless, PDF results evidenced a larger degree of tetrahedral irregularity in the case of K2Fe5−yS5 as compared with K2Fe5−ySe5. Replacement of Se by S substantially increased the irregularity of the FeCh4 tetrahedron on average, which, together with the substantial departure of the anion height from the optimum value, coincide with the observed decrease in superconductivity and the appearance of the spin glass state in the sulphide. These modeling results confirm the narrower distribution of interatomic distances in the selenide, determined from the direct comparison of the experimental PDFs. Importantly, the parent FeSe showed a slightly wider distribution compared with the K-intercalated counterpart, and consistently lower value of TC. As the anion heights of the selenide and sulfide samples studied here were positioned on the opposite sides of the optimal value of 1.38 Å, where the observed TC peaks in iron chalcogenides and pnictides, tracking this subtle structural parameter on the basis of the KxFe2−ySe2−zSz solid-solution testing ground would be valuable. In particular, systematically and quantitatively characterizing the evolution of the disorder across the phase diagram of KxFe2−ySe2−zSz at low temperatures would help correlate how the key building blocks relay the competition of magnetism with superconductivity (and vice versa), a concept of fundamental importance in Fe-based correlated electron systems.

4. Conclusions

In summary, we performed a neutron PDF analysis to explore the nanoscale atomic structures of two KxFe2−yCh2 materials (Ch = S, Se) at 5 K. The study was motivated by the phase separation concept in these compositions, which enabled the comparison of their structural features to those observed in their non-intercalated counterparts. Interestingly, the diffraction patterns of both systems displayed the (110) reflection, a hallmark of I4/m crystal symmetry. In the respective reduced total scattering structure functions F(Q), appreciable arcs of diffuse scattering, particularly at high momentum transfers Q, were indicative of the presence of substantial disorder in the system. The diffuse intensity appeared to be more pronounced for KxFe2−ySe2 than for KxFe2−yS2. The PDF analysis revealed that the data for both systems could be explained by either a mixture of majority I4/m vacancy-ordered and minority I4/mmm components, or by a single phase I4/m vacancy-disordered model. Both descriptions are effectively consistent with the underlying phase separation and complex disorder on the nanoscale. Quantitatively, in the two-phase description, the I4/mmm minority phase content did not decrease from KxFe2−ySe2 to KxFe2−yS2 as would be expected on the grounds that this phase is commonly associated with superconductivity. Similarly, the single-phase vacancy disordered model description resulted in comparable values of the occupancy of the Fe 4d site for the two systems. This indicates that the potassium intercalated FeCh is more complex than outlined by the simplistic phase separation scenario, and implies that the phase separation does not play a key role in the collapse of superconductivity in the intercalated sulphide system. An explanation might be provided by the fine details of disorder that are yet to be explored. Comparison of relevant structural parameters derived from the Fe-vacancy-disordered model demonstrated that replacement of larger Se by smaller S induced a dramatic reduction in the unit cell volume and interatomic Fe-Ch distances. Corresponding anion heights conformed to the empirical rule of Mizuguchi et al. [26]. Furthermore, the FeCh4 tetrahedra were more irregular in the non-superconducting sulphide than in the superconducting selenide. The same parameters obtained for the non-intercalated parent materials appear to also follow the same empirical trend. Overall, this neutron total scattering study highlights that deciphering the subtle differences in the atomic structure configuration of the K-intercalated end-members and their markedly different physical properties is a complex physical crystallography problem. Seeking additional insights from KxFe2−ySe2−zSz solid-solutions may be crucial for uncovering the mediating role of key structural building blocks in this puzzle.

Author Contributions

Project conceived by E.S.B., A.L. and C.P.; Sample preparation and characterization, H.L. and C.P.; Data acquisition, E.S.B. and M.F.; Data reduction, M.F., E.S.B. and M.T.M.; Data modeling and interpretation, P.M., A.L. and E.S.B.; Writing-Original Draft Preparation, P.M., A.L. and E.S.B.; Writing-Review, all co-authors; Writing-Final Editing, P.M., C.P., A.L. and E.S.B.

Funding

This research used resources at the Spallation Neutron Source, a U.S. Department of Energy Office of Science User Facility operated by the Oak Ridge National Laboratory. Alexandros Lappas acknowledges support by the U.S. Office of Naval Research Global, NICOP grant award No. N62909-17-1-2126. Work at Brookhaven National Laboratory was supported by U.S. Department of Energy, Office of Science, Office of Basic Energy Sciences (DOE-BES) under contract DE-SC0012704 and by the Center for Emergent Superconductivity, an Energy Frontier Research Center funded by the U.S. Department of Energy, Office for Basic Energy Science (Hechang Lei and Cedomir Petrovic).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Hsu, F.-C.; Luo, J.-Y.; Yeh, K.-W.; Chen, T.-K.; Huang, T.-W.; Wu, P.M.; Lee, Y.-C.; Huang, Y.-L.; Chu, Y.-Y.; Yan, D.-C.; et al. Superconductivity in the PbO-type structure α-FeSe. Proc. Natl. Acad. Sci. USA 2008, 105, 14262–14264. [Google Scholar] [CrossRef] [PubMed]
  2. Malavasi, L.; Margadonna, S. Structure–properties correlations in Fe chalcogenide superconductors. Chem. Soc. Rev. 2012, 41, 3897–3911. [Google Scholar] [CrossRef] [PubMed]
  3. Margadonna, S.; Takabayashi, Y.; Ohishi, Y.; Mizuguchi, Y.; Takano, Y.; Kagayama, T.; Nakagawa, T.; Takata, M.; Prassides, K. Pressure evolution of the low-temperature crystal structure and bonding of the superconductor FeSe (TC = 37 K). Phys. Rev. B 2009, 80, 064506. [Google Scholar] [CrossRef]
  4. Tan, S.; Zhang, Y.; Xia, M.; Ye, Z.; Chen, F.; Xie, X.; Peng, R.; Xu, D.; Fan, Q.; Xu, H.; et al. Interface-induced superconductivity and strain-dependent spin density waves in FeSe/SrTiO3 thin films. Nat. Mater. 2013, 12, 634–640. [Google Scholar] [CrossRef] [PubMed]
  5. Ge, J.-F.; Liu, Z.-L.; Liu, C.; Gao, C.-L.; Qian, D.; Xue, Q.-K.; Liu, Y.; Jia, J.-F. Superconductivity above 100 K in single-layer FeSe films on doped SrTiO3. Nat. Mater. 2015, 14, 285–289. [Google Scholar] [CrossRef] [PubMed]
  6. Fang, M.H.; Pham, H.M.; Qian, B.; Liu, T.J.; Vehstedt, E.K.; Liu, Y.; Spinu, L.; Mao, Z.Q. Superconductivity close to magnetic instability in Fe(Se1−xTex)0.82. Phys. Rev. B 2008, 78. [Google Scholar] [CrossRef]
  7. Yeh, K.-W.; Huang, T.-W.; Huang, Y.; Chen, T.-K.; Hsu, F.-C.; Wu, P.M.; Lee, Y.-C.; Chu, Y.-Y.; Chen, C.-L.; Luo, J.-Y.; et al. Tellurium substitution effect on superconductivity of the α-phase iron selenide. EPL Europhys. Lett. 2008, 84, 37002. [Google Scholar] [CrossRef] [Green Version]
  8. Katayama, N.; Ji, S.; Louca, D.; Lee, S.; Fujita, M.; Sato, T.J.; Wen, J.; Xu, Z.; Gu, G.; Xu, G.; et al. Investigation of the Spin-Glass Regime between the Antiferromagnetic and Superconducting Phases in Fe1+ySexTe1-x. J. Phys. Soc. Jpn. 2010, 79, 113702. [Google Scholar] [CrossRef]
  9. Guo, J.; Jin, S.; Wang, G.; Wang, S.; Zhu, K.; Zhou, T.; He, M.; Chen, X. Superconductivity in the iron selenide KxFe2Se2 (0 ≤ x ≤ 1.0). Phys. Rev. B 2010, 82. [Google Scholar] [CrossRef]
  10. Ye, F.; Chi, S.; Bao, W.; Wang, X.F.; Ying, J.J.; Chen, X.H.; Wang, H.D.; Dong, C.H.; Fang, M. Common Crystalline and Magnetic Structure of Superconducting A2Fe4Se5 (A = K, Rb, Cs, Tl) Single Crystals Measured Using Neutron Diffraction. Phys. Rev. Lett. 2011, 107, 137003. [Google Scholar] [CrossRef] [PubMed]
  11. Shoemaker, D.P.; Chung, D.Y.; Claus, H.; Francisco, M.C.; Avci, S.; Llobet, A.; Kanatzidis, M.G. Phase relations in KxFe2−ySe2 and the structure of superconducting KxFe2Se2 via high-resolution synchrotron diffraction. Phys. Rev. B 2012, 86. [Google Scholar] [CrossRef]
  12. Vivanco, H.K.; Rodriguez, E.E. The intercalation chemistry of layered iron chalcogenide superconductors. J. Solid State Chem. 2016, 242, 3–21. [Google Scholar] [CrossRef] [Green Version]
  13. Carr, S.V.; Louca, D.; Siewenie, J.; Huang, Q.; Wang, A.; Chen, X.; Dai, P. Structure and composition of the superconducting phase in alkali iron selenide KyFe1.6+xSe2. Phys. Rev. B 2014, 89, 134509. [Google Scholar] [CrossRef]
  14. Tanaka, M.; Yanagisawa, Y.; Denholme, S.J.; Fujioka, M.; Funahashi, S.; Matsushita, Y.; Ishizawa, N.; Yamaguchi, T.; Takeya, H.; Takano, Y. Origin of the Higher-Tc Phase in the KxFe2−ySe2 System. J. Phys. Soc. Jpn. 2016, 85, 044710. [Google Scholar] [CrossRef]
  15. Lei, H.; Abeykoon, M.; Bozin, E.S.; Wang, K.; Warren, J.B.; Petrovic, C. Phase Diagram of KxFe2−ySe2−zSz and the Suppression of its Superconducting State by an Fe2-Se/S Tetrahedron Distortion. Phys. Rev. Lett. 2011, 107. [Google Scholar] [CrossRef] [PubMed]
  16. Lei, H.; Abeykoon, M.; Bozin, E.S.; Petrovic, C. Spin-glass behavior of semiconducting KxFe2−yS2. Phys. Rev. B 2011, 83. [Google Scholar] [CrossRef]
  17. Billinge, S.J.L.; Egami, T. Underneath the Bragg Peaks: Structural Analysis of Complex Materials; Pergamon Materials Series; Elsevier: Oxford, UK, 2003; Volume 7. [Google Scholar]
  18. Neuefeind, J.; Feygenson, M.; Carruth, J.; Hoffmann, R.; Chipley, K.K. The Nanoscale Ordered MAterials Diffractometer NOMAD at the Spallation Neutron Source SNS. Nucl. Instrum. Methods Phys. Res. Sect. B 2012, 287, 68–75. [Google Scholar] [CrossRef]
  19. Peterson, P.F.; Gutmann, M.; Proffen, T.; Billinge, S.J.L. PDFgetN: A user-friendly program to extract the total scattering structure factor and the pair distribution function from neutron powder diffraction data. J. Appl. Crystallogr. 2000, 33, 1192. [Google Scholar] [CrossRef]
  20. Farrow, C.L.; Juhas, P.; Liu, J.W.; Bryndin, D.; Božin, E.S.; Bloch, J.; Proffen, T.; Billinge, S.J.L. PDFfit2 and PDFgui: Computer programs for studying nanostructure in crystals. J. Phys. Condens. Matter 2007, 19, 335219. [Google Scholar] [CrossRef] [PubMed]
  21. Shannon, R.D. Revised effective ionic radii and systematic studies of interatomic distances in halides and chalcogenides. Acta Crystallogr. Sect. A 1976, 32, 751–767. [Google Scholar] [CrossRef] [Green Version]
  22. Lazarević, N.; Abeykoon, M.; Stephens, P.W.; Lei, H.; Bozin, E.S.; Petrovic, C.; Popović, Z.V. Vacancy-induced nanoscale phase separation in KxFe2−ySe2 single crystals evidenced by Raman scattering and powder X-ray diffraction. Phys. Rev. B 2012, 86. [Google Scholar] [CrossRef]
  23. Tsuchiya, Y.; Ikeda, S.; Zhang, X.-W.; Kishimoto, S.; Kikegawa, T.; Hirao, N.; Kawaguchi, S.I.; Ohishi, Y.; Kobayash, H. Pressure-Induced Phase Transition in KxFe2−yS2. J. Phys. Soc. Jpn. 2017, 86. [Google Scholar] [CrossRef]
  24. Xu, Z.; Schneeloch, J.A.; Wen, J.; Božin, E.S.; Granroth, G.E.; Winn, B.L.; Feygenson, M.; Birgeneau, R.J.; Gu, G.; Zaliznyak, I.A.; et al. Thermal evolution of antiferromagnetic correlations and tetrahedral bond angles in superconducting FeTe1−xSex. Phys. Rev. B 2016, 93, 104517. [Google Scholar] [CrossRef]
  25. Lai, X.; Zhang, H.; Wang, Y.; Wang, X.; Zhang, X.; Lin, J.; Huang, F. Observation of Superconductivity in Tetragonal FeS. J. Am. Chem. Soc. 2015, 137, 10148–10151. [Google Scholar] [CrossRef] [PubMed]
  26. Mizuguchi, Y.; Hara, Y.; Deguchi, K.; Tsuda, S.; Yamaguchi, T.; Takeda, K.; Kotegawa, H.; Tou, H.; Takano, Y. Anion height dependence of Tc for the Fe-based superconductor. Supercond. Sci. Technol. 2010, 23, 054013. [Google Scholar] [CrossRef]
  27. Lee, C.-H.; Iyo, A.; Eisaki, H.; Kito, H.; Fernandez-Diaz, M.T.; Ito, T.; Kihou, K.; Matsuhata, H.; Braden, M.; Yamada, K. Effect of Structural Parameters on Superconductivity in Fluorine-Free LnFeAsO1-y (Ln = La, Nd). J. Phys. Soc. Jpn. 2008, 77, 083704. [Google Scholar] [CrossRef]
Figure 1. Atomic structure models for KFe2Ch2: (a) I4/mmm featuring fully occupied Fe-layers and (b) I4/m, exhibiting Fe-vacancy order. Combinations of these models (not shown) were also used in the study.
Figure 1. Atomic structure models for KFe2Ch2: (a) I4/mmm featuring fully occupied Fe-layers and (b) I4/m, exhibiting Fe-vacancy order. Combinations of these models (not shown) were also used in the study.
Condensedmatter 03 00020 g001
Figure 2. Neutron powder diffraction intensities for forward scattering 2θ = 15° NOMAD detector bank over selected d-spacing range. For easier comparison, the patterns were normalized to the intensities of (002) reflections (shaded): (a) KxFe2−ySe2 and (b) KxFe2−yS2. Arrows mark the (110) reflection that is allowed in the I4/m but not in the I4/mmm symmetry.
Figure 2. Neutron powder diffraction intensities for forward scattering 2θ = 15° NOMAD detector bank over selected d-spacing range. For easier comparison, the patterns were normalized to the intensities of (002) reflections (shaded): (a) KxFe2−ySe2 and (b) KxFe2−yS2. Arrows mark the (110) reflection that is allowed in the I4/m but not in the I4/mmm symmetry.
Condensedmatter 03 00020 g002
Figure 3. Neutron total scattering data for KxFe2−yCh2 (Ch = S, Se) at 5 K. Reduced neutron total scattering structure function, F(Q), for (a) KxFe2−ySe2 (red) and (b) KxFe2−yS2 (blue). Corresponding PDF data for KxFe2−ySe2 (red) and KxFe2−yS2 (blue) over a 5-nm PDF field of view are displayed below. (c) The low-r region is shown on an expanded scale, as compared to (d), which features intermediate distance range. The block arrow indicates the volume reduction trend from Ch = Se to S.
Figure 3. Neutron total scattering data for KxFe2−yCh2 (Ch = S, Se) at 5 K. Reduced neutron total scattering structure function, F(Q), for (a) KxFe2−ySe2 (red) and (b) KxFe2−yS2 (blue). Corresponding PDF data for KxFe2−ySe2 (red) and KxFe2−yS2 (blue) over a 5-nm PDF field of view are displayed below. (c) The low-r region is shown on an expanded scale, as compared to (d), which features intermediate distance range. The block arrow indicates the volume reduction trend from Ch = Se to S.
Condensedmatter 03 00020 g003
Figure 4. Refinements of neutron total scattering PDF data of (a) KxFe2−ySe2 and (b) KxFe2−yS2 at 5 K for the two structural models discussed in the text: (1) two-phase mixture of I4/mmm KFe2Ch2 and I4/m VO K2Fe4Ch5 (bottom row), and (2) single phase I4/m VD K2Fe5−yCh5 (top row). Observed (blue open symbols), calculated (red solid line), and difference (green solid line, offset for clarity) PDF profiles are illustrated. Fit residuals RW are also shown.
Figure 4. Refinements of neutron total scattering PDF data of (a) KxFe2−ySe2 and (b) KxFe2−yS2 at 5 K for the two structural models discussed in the text: (1) two-phase mixture of I4/mmm KFe2Ch2 and I4/m VO K2Fe4Ch5 (bottom row), and (2) single phase I4/m VD K2Fe5−yCh5 (top row). Observed (blue open symbols), calculated (red solid line), and difference (green solid line, offset for clarity) PDF profiles are illustrated. Fit residuals RW are also shown.
Condensedmatter 03 00020 g004
Figure 5. Schematic of the tetrahedral angle distribution (solid lines) extracted from the tetrahedral angles of the Fe 16i site in K2Fe5−yCh5 I4/m VD PDF model at 5 K and for FeSe in the Cmma symmetry at 10 K. The vertical dashed line marks the ideal tetrahedral angle of 109.47°. The inset (red bar) illustrates the placement of the average anion height hCh (black lines) for the material systems considered relative to the optimal value of ~1.38 Å (dashed line).
Figure 5. Schematic of the tetrahedral angle distribution (solid lines) extracted from the tetrahedral angles of the Fe 16i site in K2Fe5−yCh5 I4/m VD PDF model at 5 K and for FeSe in the Cmma symmetry at 10 K. The vertical dashed line marks the ideal tetrahedral angle of 109.47°. The inset (red bar) illustrates the placement of the average anion height hCh (black lines) for the material systems considered relative to the optimal value of ~1.38 Å (dashed line).
Condensedmatter 03 00020 g005
Table 1. Lattice parameters, interatomic Fe-Ch distances (d) and anion heights obtained from PDF refinements of I4/m VD model to the K2Fe5−yCh5 5 K data (Ch = Se, S). Neutron powder diffraction results for FeSe at 10 K [24] and x-ray powder diffraction results for FeS at 300 K [25] are also shown as a reference.
Table 1. Lattice parameters, interatomic Fe-Ch distances (d) and anion heights obtained from PDF refinements of I4/m VD model to the K2Fe5−yCh5 5 K data (Ch = Se, S). Neutron powder diffraction results for FeSe at 10 K [24] and x-ray powder diffraction results for FeS at 300 K [25] are also shown as a reference.
I4/ma = b (Å)c (Å)d1 (Å) 1d2 (Å) 2d3 (Å) 2d4 (Å) 2hCh (Å)
K2Fe4.27(1)Se58.683(1)14.001(3)2.337(8)2.446(2)2.451(3)2.512(5)1.396(9)
K2Fe4.27(1)S58.395(2)13.437(2)2.283(5)2.321(8)2.332(5)2.37(1)1.282(8)
Cmmaa (Å)b (Å)c (Å)d (Å) × 4--hCh (Å)
FeSe5.3147(5)5.3367(5)5.4855(3)2.383(1)--1.461(2)
P4/nmma = b (Å)c (Å)d (Å) × 4---hCh (Å)
FeS3.6802(5)5.0307(7)2.235(1)---1.270(1)
1 Fe-Ch(1) distance, where Ch(1) belongs to 4e site; 2 Fe-Ch(2) distances, where Ch(2) belongs to 16i site.
Table 2. Interatomic Fe-Fe distances and isotropic ADPs from the PDF refinements of I4/m VD model to the K2Fe5−yCh5 data (Ch = Se, S) at 5 K and neutron powder diffraction refinement for FeSe at 10 K.
Table 2. Interatomic Fe-Fe distances and isotropic ADPs from the PDF refinements of I4/m VD model to the K2Fe5−yCh5 data (Ch = Se, S) at 5 K and neutron powder diffraction refinement for FeSe at 10 K.
I4/mIntra-Cluster Fe-Fe (Å)Inter-Cluster Fe-Fe (Å)K, Uiso (Å2)Fe, Uiso (Å2)Ch, Uiso (Å2)
K2Fe4.27(1)Se52.649(4)2.897(5)0.0231(7)0.0087(1)0.0067(1)
K2Fe4.27(1)S52.672(1)2.746(1)0.0236(2)0.0093(1)0.0032(1)
CmmaFe-Fe(1) (Å)Fe-Fe(2) (Å)
FeSe2.657(1)2.667(1)-0.0043(7)0.0033(9)

Share and Cite

MDPI and ACS Style

Mangelis, P.; Lei, H.; McDonnell, M.T.; Feygenson, M.; Petrovic, C.; Bozin, E.S.; Lappas, A. On the Nanoscale Structure of KxFe2−yCh2 (Ch = S, Se): A Neutron Pair Distribution Function View. Condens. Matter 2018, 3, 20. https://doi.org/10.3390/condmat3030020

AMA Style

Mangelis P, Lei H, McDonnell MT, Feygenson M, Petrovic C, Bozin ES, Lappas A. On the Nanoscale Structure of KxFe2−yCh2 (Ch = S, Se): A Neutron Pair Distribution Function View. Condensed Matter. 2018; 3(3):20. https://doi.org/10.3390/condmat3030020

Chicago/Turabian Style

Mangelis, Panagiotis, Hechang Lei, Marshall T. McDonnell, Mikhail Feygenson, Cedomir Petrovic, Emil S. Bozin, and Alexandros Lappas. 2018. "On the Nanoscale Structure of KxFe2−yCh2 (Ch = S, Se): A Neutron Pair Distribution Function View" Condensed Matter 3, no. 3: 20. https://doi.org/10.3390/condmat3030020

Article Metrics

Back to TopTop