First-principles study of negative thermal expansion mechanism in A-site-ordered perovskite SrCu3Fe4O12

Hongping Lia, Shuai Liua, Lin Chena, Jiandong Zhaoa, Beibei Chena, Zhongchang Wang*b, Jian Mengc and Xiaojuan Liu*c
aInstitute for Advanced Materials, School of Materials Science and Engineering, Jiangsu University, Zhenjiang, 212013, P. R. China
bAdvanced Institute for Materials Research, Tohoku University, 2-1-1 Katahira, Aoba-ku, Sendai 980-8577, Japan. E-mail: zcwang@wpi-aimr.tohoku.ac.jp
cState Key Laboratory of Rare Earth Resources Utilization, Changchun Institute of Applied Chemistry, Chinese Academy of Sciences, Changchun, 130022, P. R. China. E-mail: lxjuan@ciac.ac.cn

Received 14th August 2014 , Accepted 26th November 2014

First published on 27th November 2014


Abstract

Negative thermal expansion (NTE) materials offer a tremendous opportunity for fundamental as well as applied research, yet the origin remains difficult to understand. Here we perform a systematic first-principles calculation to investigate electrical and magnetic properties of an A-site-ordered perovskite SrCu3Fe4O12 and clarify its NTE mechanism. We find that SrCu3Fe4O12 is an antiferromagnetic metal, and its magnetic ordering can be demonstrated as C-type within Fe ions of mixed valence at the B-site and paramagnetic Cu3+ at the A-site. Electronic structure analysis reveals that there occurs temperature-induced Cu–Fe intersite charge transfer, which is mediated by the corner-sharing O atoms. Meanwhile, the magnetic interaction is found to undergo a transition from antiferromagnetism to ferrimagnetism. Further phonon calculations demonstrate that the SrCu3Fe4O12 is thermodynamically stable both at low and high temperature, and that there appear degenerated phonon branches, indicating that it is easy to transfer energy between these modes. We also find that the large NTE of SrCu3Fe4O12 originates from intermetallic charge transfer induced by temperature, which relaxes the Sr–O and Fe–O bonding units in the oxide.


Introduction

Negative thermal expansion (NTE) materials, which exhibit volume contraction upon heating, have aroused extensive attention, largely because their unique thermal properties are relevant for many practical applications such as elasticity-tuned sensors and switching devices. Moreover, they can be mixed with positive thermal expansion materials so as to tune the thermal expansion coefficient of materials. To date, several NTE materials with a large linear thermal expansion coefficient (α) of an order of 10−5 k−1, have been reported,1–3 for instance, the Ge-doped anti-perovskite manganese nitrides Mn3AN (A = Cu, Zn, Ga) show a giant α of up to −2.5 × 10−5 k−1.1 In addition to their thermal property, NTE materials can often simultaneously show intriguing electronic and physical properties, rendering them promising candidates for many technological applications.

Much effort has been devoted recently to the study of the NTE phenomena in the A-site-ordered perovskites A′A3B4O12 due to their unique structures and for their wide range of intriguing physical properties. For instance, LaCu3Fe4O12 and BiCu3Fe4O12 are reported to show temperature-induced Cu–Fe intermetallic charge transfer (CT) in the form of 3Cu3+ + 4Fe3+ → 3Cu2+ + 4Fe3.75+ from low-temperature La3+Cu33+Fe43+O12 (Bi3+Cu33+Fe43+O12) with Cu3+ at the A-site to high-temperature La3+Cu32+Fe43.75+O12 (Bi3+Cu32+Fe43.75+O12) with Fe3.75+ at the B-site, which is accompanied by a first-order isostructural transition from antiferromagnetic (AFM) insulator to paramagnetic metal with a remarkable volume shrinkage.4–6 In contrast, perovskite YCu3Fe4O12 shows a vastly different behavior: the instability of its high-valence Fe3.75+ atoms can be relieved by a charge disproportionation (CD) via the form of 8Fe3.75+ → 5Fe3+ + 3Fe5+, resulting in a charge ordering and a ferrimagnetic (FIM) ordering at low temperature without any NTE transition.7 To investigate the mutual interaction between strain and electronic phases, Yamada et al.8 conducted an systematic study of the LnCu3Fe4O12 (Ln: lanthanide) by adjusting A′-site ion size via choosing different lanthanides, and found that the oxides with a large Ln ion size show the intersite CT and isostructural NTE transition, while that those with a small Ln ion size show the CD. These findings indicate that both the bond strain and the amount of charges at the A-site play an important role in triggering the intersite CT or CD. Theoretically, the strength of crystal-field splitting and relative energy ordering between Cu 3dxy and Fe 3d states are reported to be key parameters in triggering the intersite CT and CD.9

Moreover, A-site-ordered perovskites with divalent alkaline-earth metals can also be successfully prepared and exhibit interesting properties. For instance, CaCu3Fe4O12 undergoes a phase transition to CD (2FeIIIL → FeIII + FeIIIL2) accompanied by a volume contraction at ∼210 K, and its ferrimagnetism has been found to originate from the AFM coupling of the Cu and Fe sublattices.10 On the contrary, SrCu3Fe4O12 is an antiferromagnet with TN of 180 K, and exhibits a large NTE with a linear expansion coefficient of −2.26 × 10−5 K−1.11,12 Furthermore, it has been reported that a partial replacement of Ca2+ with Sr2+ in CaCu3Fe4O12 can alter significantly its electrical and magnetic properties. To understand the origins of the magnetic coupling interactions and the large NTE, here, we investigate systematically the electronic and magnetic properties of SrCu3Fe4O12 by first-principles calculations, and demonstrate that this oxide is intrinsically an AFM metal. The large NTE is identified to originate from the temperature-induced Cu–Fe intersite CT, opening up an additional avenue in clarifying mechanisms of NTE phenomenon in transition-metal oxides.

Computational details

Calculations of energies and electronic structures were performed with the full-potential linearized augmented plane wave plus local orbitals methods, which were implemented in WIEN2K package.13,14 The generalized gradient approximation (GGA) of Perdew–Burke–Ernzerhof (PBE)15 was employed to describe the exchange-correlation functional. The values of atomic sphere radii (RMT) were chosen to be 2.41, 1.83, 1.93, and 1.62 a.u. for Sr, Cu, Fe and O, respectively. Wave function was expanded using the plane wave with a cutoff of 7.0 and the density and potential a value of 14. Sampling of irreducible Brillouin zone was performed with k points of 1000 using the modified tetrahedron method.16 Self consistence was achieved once the total energy was converged to less than 10−5 Ry/f.u. Electron correlation effect was taken into account by the effective Hubbard parameter Ueff = UJ,17 where U and J are the on-site Coulomb repulsion and Hund exchange constant, respectively. For simplicity, the U is used hereafter to represent the effective parameter Ueff. A series of U values were chosen from 2.0 to 7.0 eV for Cu (UCu) and from 1.0 to 5.0 eV for Fe (UFe). The spin–orbital coupling was also taken in account due to the significant orbital contribution in CaCu3Fe4O12,18 yet was found to impose negligible influence on our calculated results. In addition, the phonon properties were calculated using the PHONOPY package19 combined with Vienna Ab initio Simulation Package (VASP).20 The cut-off energy of plane wave was set to be 400 eV and the finite displacement method was adopted. The geometrical structures were fully relaxed until the change in total energy was less than 10−8 eV per atom.

Results and discussion

Crystal structure of SrCu3Fe4O12

SrCu3Fe4O12 crystallizes in the Im[3 with combining macron] cubic lattice (shown in Fig. 1),11 in which the FeO6 octahedra are fairly rigid yet markedly tilted, transforming the original twelvefold coordinated A-site Cu ions to square-coordinated CuO4 units that are perpendicular to each other. Its structure falls in three ranges (listed in Table 1): an ordinary positive thermal expansion range when temperature is either lower than 170 K (range 1) or higher than 270 K (range 3), and a large NTE range when temperature is in between 170 and 270 K (range 2). In ranges 1 and 3, its lattice constant and bond length expand with the rise of temperature, while in range 2, the Fe–O and Sr–O bonds shrink by ∼0.6% and ∼0.4%, respectively, due to its large NTE. However, the Cu–O distance is elongated significantly by ∼1.4% from 1.878 Å at 170 K to 1.901 Å at 270 K (Table 1). Moreover, its lattice constant relies primarily on the Fe-related parameters, i.e. Fe–O bond length and Fe–O–Fe bond angle, in which the contraction of Fe–O bond contraction is critical to the volume contraction in range 2, albeit that the Fe–O–Fe bond angle is increased by ∼1.5°.
image file: c4ra08652j-f1.tif
Fig. 1 Crystal structure of A-site-ordered double perovskite SrCu3Fe4O12 with Im[3 with combining macron] symmetry. A′- and A-site ions are ordered at the ratio of 1[thin space (1/6-em)]:[thin space (1/6-em)]3 and B-site ions form a heavily tilted FeO6 octahedron to optimize Sr–O and Fe–O bond lengths.
Table 1 The structural parameters of SrCu3Fe4O12 at the transfer temperature for the three ranges
  Range 1 Range 2 Range 3
80 K 170 K 270 K 450 K
a (Å) 7.356 7.358 7.348 7.357
Sr–O (Å) 2.615 2.620 2.607 2.601
Fe–O (Å) 1.979 1.979 1.968 1.970
Cu–O (Å) 1.878 1.878 1.901 1.905
Fe–O–Fe (°) 136.676 136.672 138.028 138.048


Magnetic ground state of SrCu3Fe4O12

Two experimentally observed structures of SrCu3Fe4O12 are adopted to calculate the magnetic ground state, i.e. 80 and 270 K, the former of which is low-temperature structure below TN and the latter of which is the most contracted one above TN.11 We construct several magnetic configurations by taking two likely magnetic sublattices of Cu and Fe into account: ferromagnetism (FM), A-type and G-type AFM within Cu sublattice, and the FM, A-type, C-type and G-type AFM within Fe sublattice. Furthermore, we also consider the Cu–Fe FIM, in which the Cu and Fe sublattices are arranged in an antiparallel manner.

To gain insights into how magnetic structures correlate with electron localization, we perform GGA + U calculations using different U values (2.0 to 7.0 eV for UCu and 1.0 to 5.0 eV for UFe) as well as GGA calculations for comparison. For the 80 K states, the magnetic moment of Cu is calculated to be nearly zero when magnetic interaction within Fe sublattice is AFM coupled, whichever the GGA or GGA + U method is used. Hence, we mainly analyze the results of the AFM coupling between Fe sublattice and Cu–Fe FIM for comparison. Table 2 lists their relative energies ΔE as a function of U (taking the A-type AFM within Fe sublattice, which hereafter is named AFM-A, as a reference). It is clear that the CC-type AFM (named AFM-C) is energetically favorable than the A-type and G-type AFM within Fe sublattice (named AFM-G), though the energy of AFM-G configuration is lower than that of AFM-C when the UCu and UFe are adopted as (3.0, 2.0) eV and (4.0, 3.0) eV, respectively. Moreover, the lowest energy state calculated using the GGA is FIM (see Table 2), yet switches to AFM-A when the UCu and UFe are adopted as 4.0 and 3.0 eV, respectively, consistent with the AFM observations below 180 K.11 In particular, the total energy difference between the AFM-C and FIM configurations is increased markedly with the rise of U (Table 2), implying that the former is affected more severely by the electron correlation. As for the high-temperature structure at 270 K, the FIM configuration is always found to be most stable, despite of the calculation method. In addition, the low-temperature ground state is energetically more preferred than the high-temperature one.

Table 2 Relative energy difference (in meV per formula unit) between the different magnetic configurations at 80 K. Calculations are conducted using the GGA + U with the U ranging from 0.0 to 7.0 eV for Cu and from 0.0 to 5.0 eV for Fe
  UCu 0.0 2.0 3.0 4.0 5.0 6.0 7.0
UFe 0.0 1.0 2.0 3.0 4.0 5.0 5.0
AFM-A 0.0 0.0 0.0 0.0 0.0 0.0 0.0
AFM-C 343.4 93.4 −78.7 −79.3 −45.5 −5.1 −10.6
AFM-G 508.3 24.4 −172.2 −88.8 2.23 82.3 88.9
FIM −684.4 −817.6 −380.0 69.0 330.6 496.9 339.1


For the 80 K states, the magnetic moment of Fe increases as the UFe increases from 2.0 to 4.0 eV, and maintains at ∼4.0 μB (Table 3). Meanwhile, the calculated total magnetic moment for each formula unit nearly maintains zero, whichever the GGA or GGA + U method is adopted, confirming the low-temperature AFM phase.11 Similarly, the magnetic moment of Cu or Fe is also increased with the rise of U in the case of 270 K. Based upon these results and our previous studies,21 the GGA + U method is adopted with UCu = 5.0 eV and UFe = 4.0 eV, unless otherwise stated.

Table 3 Calculated total magnetic (Mtotal) per formula unit and corresponding magnetic moments of Cu (MCu) and Fe (MFe) ions (in μB) for the AFM-C at 80 K and FIM at 270 K. Calculations are conducted using the GGA + U with the U ranging from 0.0 to 7.0 eV for Cu and from 0.0 to 5.0 eV for Fe
  UCu 0.0 2.0 3.0 4.0 5.0 6.0 7.0
UFe 0.0 1.0 2.0 3.0 4.0 5.0 5.0
AFM-C MCu 0.00 0.00 0.02 0.01 0.02 0.01 0.02
MFe 2.64 3.27 3.68 3.84 3.94 4.04 4.03
Mtotal 0.03 0.05 0.04 0.04 0.02 0.00 0.00
FIM MCu −0.21 −0.38 −0.48 −0.54 −0.52 −0.53 −0.67
MFe 2.72 3.12 3.37 3.54 3.75 3.88 3.82
Mtotal 11.8 13.0 13.4 13.7 14.9 15.5 14.2


Comparative analysis of electronic structures of SrCu3Fe4O12

To shed light on the relation between magnetic coupling and charge distribution, we conduct a comparative study of electronic structures of AFM-C at 80 K and FIM at 270 K. Fig. 2 shows total density of states (TDOS) and partial density of states (PDOS) for the AFM-C and FIM configurations calculated using both GGA and GGA + U. From the TDOS, a metallic behavior is observed for the two configurations with both spins crossing Fermi level (EF), independent of the calculation method. This indicates that low-temperature ground state of SrCu3Fe4O12 is AFM yet metallic, in sharp contrast to those of a conventional compound showing either the FM ordering with metallicity or AFM ordering with insulating states. Further GGA + U calculations reveal that the 3d orbitals below EF are pushed to low energy area, while those above the EF are pushed to the high energy area, indicating that the electron correlation plays an important role in 3d electrons in this system.
image file: c4ra08652j-f2.tif
Fig. 2 TDOS and PDOS plots of the Cu, Fe and O atom contributions for the AFM-C configuration calculated using the (a) GGA and (b) GGA + U method. The corresponding plots for the FIM configuration calculated using the (c) GGA and (d) GGA + U method. The inset in (b) enlarges the DOS around Fermi level, highlighting the metallic nature. The Fermi level is aligned to zero and indicated by a vertical dashed line.

For the low-temperature AFM-C configuration, the contribution to the DOS at EF comes mainly from Fe and O atoms, as shown in inset of Fig. 2(b), implying that they are crucial to realizing the metallicity. Importantly, the PDOSs of Fe, Cu and O span a broad energy range from −7.6 eV to 4.5 eV, indicating a strong hybridization between them. Moreover, the A-site Cu atoms are nonmagnetic as there is a symmetry in the spin-up and spin-down channel (Fig. 2(b)), consistent with the zero magnetic moment (Table 3). Further PDOS analyses reveal that the 3dxz orbital of Cu is unoccupied in both the spin-up and the spin-down channels, while the other four 3d orbitals are all occupied (Fig. 3(a)), a typical characteristic of d8 electronic configuration for the Cu3+, similar to what was seen in the low-temperature LaCu3Fe4O12 and BiCu3Fe4O12.4,6 On the other hand, the spin-up channel of Fe is almost filled (Fig. 3(b)), while its spin-down channel is utterly empty, demonstrating that it has a high-spin d5 electronic configuration. The O atoms show the p6 electronic states in view of the nearly full occupation of 2p orbitals (Fig. 3(c)). In light of charge neutrality, one can conclude that the charge combination is Sr2+Cu33+Fe43.25+O122− at low temperature. This mixed valence for the Fe3.25+ is close to the experimental 57Fe Mössbauer spectrum observation, where the isomer shift of B-site Fe is divided into a pair of components with a ratio of Fe3+[thin space (1/6-em)]:[thin space (1/6-em)]Fe5+ to 4[thin space (1/6-em)]:[thin space (1/6-em)]1.11 This is, however, reasonable because the Fe3+ and Fe5+ cations are randomly distributed in SrCu3Fe4O12.


image file: c4ra08652j-f3.tif
Fig. 3 PDOS plots of the (a) Cu 3d, (b) Fe 3d, and (c) O 2p orbitals for the AFM-C configuration. Corresponding plots for the FIM configuration are given in (d), (e) and (f), respectively. The Fermi level is aligned to zero and indicated by a vertical dashed line.

The Fe, Cu and O atoms all contribute significantly to the electronic states at EF for the high-temperature FIM configuration (Fig. 2(d)), in sharp contrast to the case of low-temperature AFM-C state. Evidently, the Cu atoms show magnetic nature, and the Cu2+ is formed with a d9 electronic configuration because its spin-down 3dxz orbital is shifted to low energy area (Fig. 2(d)). By comparing the d9 of Cu2+ with d8 electronic configuration of Cu3+ in low-temperature AFM-C, electrons are introduced to Cu sites. Consequently, B-site Fe atoms release electrons because their energetically degenerate 3dxz and 3dyz states are shifted to high energy (Fig. 3(e)). In this respect, the valence states should be Sr2+Cu32+Fe44+O122− at the high temperature. The temperature-induced intersite CT actually takes place between the Cu and Fe in SrCu3Fe4O12 via a form of 3Cu3+ + 4Fe3.25+ → 3Cu2+ + 4Fe4+, in consistence with the experimental observations.11 In addition, the magnetic moment (Table 3) is calculated to be 3.94 μB for Fe in the AFM-C, and −0.52 and 3.75 μB for Cu and Fe in the FIM, respectively, all of which are somewhat smaller than their respective theoretical values (4.75 μB for Fe3.25+ in AFM-C, and 1 and 4 μB for Cu2+ and Fe4+ in FIM, respectively). Such differences can be attributed to the fact that a small amount of magnetic moment is distributed to the O atoms, suggestive of the formation of covalent bonding.

The PDOSs of Cu, Fe and O atoms span a broad energy range and are almost in the same energy region with a similar intensity. In particular, the O 2px and 2py share the energy range with the Cu 3dxz orbital and the O 2pz share the energy range with the degenerate Fe 3dxz and 3dyz orbitals, indicating a strong covalent hybridization between them. Consequently, the Cu3+ prefers the d9L (L: a ligand oxygen hole) rather than the d8 electronic configuration, while the Fe3.25+ and Fe4+ have the d5L0.25 and d5L electronic configuration, respectively. The Cu–Fe intersite CT in SrCu3Fe4O12 can be expressed as 3d9L + 4d5L0.25 → 3d9 + 4d5L, which can be realized by the redistribution of ligand holes from Fe–O to Cu–O bonds. The Cu–Fe intersite CT is, however, not energetically expensive because all the FeO6 octahedra and CuO4 square-planar units are linked by the corner-sharing O atoms in SrCu3Fe4O12. Meanwhile, holes are found to be injected into the O 2pz orbital (Fig. 3(f)), as confirmed by upshift of its spin-up channel. The Cu–Fe intersite CT is thus mediated by O 2p orbitals via the form of Cu 3dxz → O 2px, 2py → O 2pz → Fe 3dxz + 3dyz. To test this scenario, we present the charge density difference (CDD) plots near EF for both the AFM-C and FIM (Fig. 4). There is no CDD distributed around Cu in the AFM-C (Fig. 4(a)), whereas the 3dxz orbital of Cu is occupied in the spin-down channel (Fig. 4(b)) in the FIM state. The 2pz orbital of O atom spatially expands to its neighboring Fe in the AFM-C state, while the 2px and 2py orbitals of O expand to its nearest Cu in the FIM state. These verify that the temperature-induced intersite CT occurs through the form of Cu 3dxz → O 2px, 2py → O 2pz → Fe 3dxz + 3dyz when the magnetic configuration varies from the low-temperature AFM-C to the high-temperature FIM.


image file: c4ra08652j-f4.tif
Fig. 4 Three-dimensional CDD plots for the (a) AFM-C and (b) FIM configurations of SrCu3Fe4O12 calculated using the GGA + U method. The spin-up and spin-down electrons are indicated in red and blue, respectively.

Comparative analysis of phonon property of SrCu3Fe4O12

Since vibrating properties are important for evaluating its thermal and dynamical stability, we also conducted the phonon calculations of SrCu3Fe4O12. The calculated phonon dispersion curves along the high symmetry lines in the Brillouin zone and the corresponding total phonon DOS are shown in Fig. 5 and 6, respectively. Obviously, there is no soft mode, i.e. negative frequencies in the phonon dispersion relations, at any wave vectors both in the low-temperature (Fig. 5(a)) and in high-temperature (Fig. 5(b)) states, confirming the dynamical stabilities of SrCu3Fe4O12. It is noteworthy that the phonon spectra can be divided into three intervals with the wide band gaps in the low-temperature SrCu3Fe4O12, which is mainly attributed to the large mass difference between these atoms. However, for the high-temperature structure, the high-frequency and mid-frequency phonon bands are practically disappeared and merged in the low-frequency region, which is also confirmed in its phonon DOS plot (shown in Fig. 6). These phonon spectrums transfer should be ascribed to the weakening of chemical bonding between the corresponding cations and anions under the higher-temperature circumstance.22 Furthermore, one can notice from Fig. 5 that the phonon branches are severely degenerated at the high-symmetry R and G points, whereas at the adjacent M and X points all the phonon modes are not degenerate. The closeness between the frequencies of the intensive phonon modes suggests that the phonon–phonon coupling turns stronger, which should benefit the realization of its metallic nature. In addition, the frequencies of some optical modes are close to those of some acoustic modes, suggesting that it is not hard to transfer energy between these modes.
image file: c4ra08652j-f5.tif
Fig. 5 Phonon dispersion curves along the high symmetry directions for SrCu3Fe4O12 at (a) 80 K and (b) 270 K.

image file: c4ra08652j-f6.tif
Fig. 6 Total phonon DOS for SrCu3Fe4O12 at 80 K and 270 K.

Charge transfer and negative thermal expansion mechanism in SrCu3Fe4O12

As discussed above, the electrical and magnetic properties of SrCu3Fe4O12 differ from those of its analogue CaCu3Fe4O12.4,6 The discrepancy can primarily be ascribed to the large difference in bond strain arising from the different ionic size between Sr2+ and Ca2+ (1.44 Å for Sr2+ and 1.34 Å for Ca2+), as in the case of the LnCu3Fe4O12 compounds.8 This also implies that the difference in bond strain might be responsible for the remarkable electronic phase transitions. In addition, the difference can also be observed in iron-based perovskites CaFeO3 and SrFeO3, for example, there occurs the CD of 2Fe4+ → Fe5+ + Fe3+ (2Fe3+L → Fe3+L2 + Fe3+) in CaFeO3,23 while metallic conductivity and cubic symmetry are preserved in SrFeO3 even at low temperatures,24 which are amazingly similar to those shown in CaCu3Fe4O12 and SrCu3Fe4O12.10,11 In view of the fact that CuO4 square-planar units are spatially isolated from each other, electrical property of this system is dominated by the corner-sharing FeO6 octahedral networks. In this sense, A-site-ordered perovskite BaCu3Fe4O12 is believed to have a similar electrical behavior as the perovskite BaFeO3.25

Structurally, an obvious lattice expansion is seen owing to the larger ionic size of Sr2+ as compared to that of Ca2+. Meanwhile, Fe–O bonds are elongated from 1.979 Å in SrCu3Fe4O12 to 1.970 and 1.893 Å in CaCu3Fe4O12, which is accompanied by the elongation of the Sr–O bonds as well (from 2.620 Å in SrCu3Fe4O12 to 2.585 Å in CaCu3Fe4O12). As a result, the CD is not preferred because not only the compression stress of Sr–O and Fe–O bonds should be overcome, but also the crystal symmetry is lowered due to the distortion of FeO6 octahedra. Consequently, the intermetallic CT should account for the relief of its structural instability and the structural linkage of the FeO6 octahedra and the CuO4 square-planar units by the corner sharing is critical to the charge distribution. In addition, the Cu–O bond is very sensitive to temperature, showing an unexpected elongation in the range of NTE, which indicates that the Cu2+ is prone to be formed because the ionic radius of Cu2+ is larger than that of Cu3+. This also implies that once the Cu–O bonds are elongated to a certain level, the Cu3+ can be transformed to Cu2+ by electron gain from B-site Fe as Fe has a variable valence state, leading to the shrinking of the Fe–O bonds. The O atoms hence play an important role in mediating active charge in this system. The unusual NTE of SrCu3Fe4O12 originates from the intersite CT, which relaxes the compression of Sr–O and Fe–O bonding units.

Conclusions

We have conducted a first-principles calculation to investigate the electrical and magnetic properties of A-site-ordered perovskite SrCu3Fe4O12, aimed at clarifying its NTE mechanism. Our calculations demonstrate that SrCu3Fe4O12 is intrinsically an AFM metal with an C-type magnetic configuration for Fe ions of mixed valence at the B-site and paramagnetic Cu3+ at the A-site. Electronic structure analysis reveals that there occurs temperature-induced intersite CT between the Cu and Fe in this oxide in the form of Cu 3dxz → O 2px, 2py → O 2pz → Fe 3dxz + 3dyz. The unusual NTE is found to originate from the temperature-induced intermetallic CT, which relaxes the Sr–O and Fe–O bonding units, opening up thereby an additional avenue to understand origins of the NTE in transition-metal oxides.

Acknowledgements

This work was supported by the National Natural Science Foundation of China (NSFC) under grant no. 21301075 and 51372244, Specialized Research Fund for the Doctoral Program of Higher Education under grant no. 20133227120003, the Open Project of State Key Laboratory of Rare Earth Resources Utilization, Changchun Institute of Applied Chemistry, Chinese Academy of Sciences (CAS) (grant no. RERU2014006), the Natural Science Foundation of Jiangsu Province (grant no. BK20140551) and Research Foundation for Advanced Talents of Jiangsu University (grant no. 12JDG096). Z.W. appreciates the financial supports from the Grant-in-Aid for Young Scientists (A) (grant no. 24686069), the NSFC (grant no. 11332013), the JSPS and CAS under the Japan-China Scientific Cooperation Program, and the Murata Science Foundation.

Notes and references

  1. K. Takenaka and H. Takagi, Appl. Phys. Lett., 2005, 87, 261902 CrossRef PubMed.
  2. T. A. Mary, J. S. O. Evans, T. Vogt and A. W. Sleight, Science, 1996, 272, 90 CAS.
  3. B. K. Greve, K. L. Martin, P. L. Lee, P. J. Chupas, K. W. Chapman and A. P. Wilkinson, J. Am. Chem. Soc., 2010, 132, 15496 CrossRef CAS PubMed.
  4. Y. W. Long, N. Hayashi, T. Saito, M. Azuma, S. Muranaka and Y. Shimakawa, Nature, 2009, 458, 60 CrossRef CAS PubMed.
  5. Y. W. Long, T. Kawakami, W. T. Chen, T. Saito, T. Watanuki, Y. Nakakura, Q. Q. Liu, C. Q. Jin and Y. Shimakawa, Chem. Mater., 2012, 24, 2235 CrossRef CAS.
  6. Y. W. Long, T. Saito, T. Tohyama, K. Oka, M. Azuma and Y. Shimakawa, Inorg. Chem., 2009, 48, 8489 CrossRef CAS PubMed.
  7. H. Etani, I. Yamada, K. Ohgushi, N. Hayashi, Y. Kusano, M. Mizumaki, J. Kim, N. Tsuji, R. Takahashi, N. Nishiyama, T. Inoue, T. Irifune and M. Takano, J. Am. Chem. Soc., 2013, 135, 6100 CrossRef CAS PubMed.
  8. I. Yamada, H. Etani, K. Tsuchida, S. Marukawa, N. Hayashi, T. Kawakami, M. Mizumaki, K. Ohgushi, Y. Kusano, J. Kim, N. Tsuji, R. Takahashi, N. Nishiyama, T. Inoue, T. Irifune and M. Takano, Inorg. Chem., 2013, 52, 13751 CrossRef CAS PubMed.
  9. N. Rezaei, P. Hansmann, M. S. Bahramy and R. Arita, Phys. Rev. B: Condens. Matter Mater. Phys., 2014, 89, 125125 CrossRef.
  10. I. Yamada, K. Takata, N. Hayashi, S. Shinohara, M. Azuma, S. Mori, S. Muranaka, Y. Shimakawa and M. Takano, Angew. Chem., Int. Ed., 2008, 47, 7032 CrossRef CAS PubMed.
  11. I. Yamada, K. Tsuchida, K. Ohgushi, N. Hayashi, J. Kim, N. Tsuji, R. Takahashi, M. Matsushita, N. Nishiyama, T. Inoue, T. Irifune, K. Kato, M. Takata and M. Takano, Angew. Chem., Int. Ed., 2011, 50, 6579 CrossRef CAS PubMed.
  12. I. Yamada, K. Shiro, K. Oka, M. Azuma and T. Irifune, J. Ceram. Soc. Jpn., 2013, 121, 912 CrossRef CAS.
  13. K. Schwarz and P. Blaha, Comput. Mater. Sci., 2003, 28, 259 CrossRef CAS.
  14. P. Blaha, K. Schwarz, G. K. H. Madsen, D. Kvasnicka and J. Luitz, WIEN2k, Vienna University of Technology, Austria, 2001 Search PubMed.
  15. J. P. Perdew, K. Burke and M. Ernzerhof, Phys. Rev. Lett., 1996, 77, 3865 CrossRef CAS.
  16. P. E. Blöchl, Phys. Rev. B: Condens. Matter Mater. Phys., 1994, 50, 17953 CrossRef.
  17. S. L. Dudarev, G. A. Botton, S. Y. Savrasov, C. J. Humphreys and A. P. Sutton, Phys. Rev. B: Condens. Matter Mater. Phys., 1998, 57, 1505 CrossRef CAS.
  18. M. Mizumaki, W. T. Chen, T. Saito, I. Yamada, J. P. Attifield and Y. Shimakawa, Phys. Rev. B: Condens. Matter Mater. Phys., 2011, 84, 094418 CrossRef.
  19. A. Togo, F. Oba and I. Tanaka, Phys. Rev. B: Condens. Matter Mater. Phys., 2008, 78, 134106 CrossRef.
  20. G. Kresse and J. Furthmuller, Phys. Rev. B: Condens. Matter Mater. Phys., 1996, 54, 11169 CrossRef CAS.
  21. H. P. Li, S. H. Lv, Z. C. Wang, Y. J. Xia, Y. J. Bai, X. J. Liu and J. Meng, J. Appl. Phys., 2012, 111, 103718 CrossRef PubMed.
  22. L. W. Shi, J. Hu, Y. Qin, Y. F. Duan, L. Wu, X. Q. Yang and G. Tang, J. Alloys Compd., 2014, 611, 210 CrossRef CAS PubMed.
  23. M. Takano, N. Nakanishi, Y. Takeda, S. Naka and T. Takada, Mater. Res. Bull., 1977, 12, 923 CrossRef CAS.
  24. J. B. Macchesney, R. C. Sherwood and J. F. Potter, J. Chem. Phys., 1965, 43, 1907 CrossRef CAS PubMed.
  25. N. Hayashi, T. Yamamoto, H. Kageyama, M. Nishi, Y. Watanabe, T. Kawakami, Y. Matsushita, A. Fujimori and M. Takano, Angew. Chem., Int. Ed., 2011, 50, 12547 CrossRef CAS PubMed.

This journal is © The Royal Society of Chemistry 2015
Click here to see how this site uses Cookies. View our privacy policy here.