high-pressure synthesis of osmium oxides with double-perovskite structure … · 2019. 3. 19. ·...

128
Instructions for use Title High-Pressure Synthesis of Osmium Oxides with Double-Perovskite Structure and Their Magnetic Properties Author(s) 馮, 海 Citation 北海道大学. 博士(理学) 甲第11587号 Issue Date 2014-09-25 DOI 10.14943/doctoral.k11587 Doc URL http://hdl.handle.net/2115/57240 Type theses (doctoral) File Information Hai_Feng.pdf Hokkaido University Collection of Scholarly and Academic Papers : HUSCAP

Upload: others

Post on 09-Feb-2021

2 views

Category:

Documents


0 download

TRANSCRIPT

  • Instructions for use

    Title High-Pressure Synthesis of Osmium Oxides with Double-Perovskite Structure and Their Magnetic Properties

    Author(s) 馮, 海

    Citation 北海道大学. 博士(理学) 甲第11587号

    Issue Date 2014-09-25

    DOI 10.14943/doctoral.k11587

    Doc URL http://hdl.handle.net/2115/57240

    Type theses (doctoral)

    File Information Hai_Feng.pdf

    Hokkaido University Collection of Scholarly and Academic Papers : HUSCAP

    https://eprints.lib.hokudai.ac.jp/dspace/about.en.jsp

  • High-Pressure Synthesis of Osmium Oxides with

    Double-Perovskite Structure and Their Magnetic

    Properties

    A Thesis

    Submitted by

    Hai FENG

    In fulfillment for the award of the degree of

    Doctor of Science

    Graduate School of Chemical Sciences and Engineering,

    Hokkaido University

    2014

  • I

    Abstract

    Double perovskite oxides A2BB′O6, in which B and B′ are usually 3d (or

    nonmagnetic) ion and 5d (or 4d) ion, respectively, have attracted great attention because of

    the half-metallic (HM) nature found in Sr2FeMoO6 and Sr2FeReO6. In addition, a series of

    exotic magnetic states such as spin singlet and spin freezing was discovered for A2BB′O6. In

    the HM state, 3d-t2g and 5d(4d)-t2g electrons in the B and B′ sites play a pivotal role in the

    highly spin-polarized conduction, which results from a generalized double exchange

    mechanism. However, the generalized double exchange mechanism was unable to account for

    the remarkably high critical temperature (TC) of ferrimagnetic (FIM) transition at 725 K in

    Sr2CrOsO6 because of the multi-orbital Mott state. The FIM insulating state was therefore

    distinguishable from the HM state: the high-TC FIM transition was argued to be a novel

    phenomenon. To further elucidate the mechanism of the high-TC FIM transition as well as to

    explore significant magnetic states, additional 5d double perovskite oxides have been highly

    desired.

    In my study, newly compositional 5d-3d hybrid double perovskite oxides

    Ca2FeOsO6 and Ba2CuOsO6 were synthesized and the fundamental properties were

    investigated. Ca2FeOsO6 crystallized into an ordered double-perovskite structure with a space

    group of P21/n and showed a long-range FIM transition at a temperature of ~320 K. Although

    Ca2FeOsO6 was not a band insulator, it appeared to be electrically insulating like Sr2CrOsO6

    (TC ~725 K): the electronic state of Ca2FeOsO6 was adjacent to a HM state as well as that of

    Sr2CrOsO6. Besides, the FIM state was found to be driven by lattice distortion, being

    observed for the first time among double-perovskite oxides. Ca2FeOsO6 as well as Sr2CrOsO6

    apparently established a new class of 5d-3d hybrid FIM insulators with high-TC, which could

    be useful for advanced scientific and spintronics applications.

  • II

    Magnetic studies of Ca2-xSrxFeOsO6 indicated that TC increases with decreasing the

    unit cell volume. Therefore, TC was reasonably expected to increase to some extent by further

    contraction of the cell. In this study, it was achieved by a partial substitution of Cd for Ca,

    resulting in successful increase of TC to 360 K (Ca1.9Cd0.1FeOsO6) from 320 K (Ca2FeOsO6).

    Ba2CuOsO6 crystallized under a certain high-pressure condition into a double

    perovskite structure with a space group of I4/m, in which Os (VI) and Cu (II) were ordered in

    the perovskite B-site. Ba2CuOsO6 was found to be electrically insulating and to show

    antiferromagnetic characteristics at temperatures of ~55 K and ~70 K. The Jahn-Teller

    distortion of CuO6 octahedra was argued to influence the magnetic transition via possible

    two-dimensional magnetic correlations. The first-principles study suggested that the

    spin-orbit interaction of Os (VI) plays a substantial role in the insulating state.

    The double perovskite oxides Ca2InOsO6 (Os5+

    , S = 3/2), Ca3OsO6 (Os6+

    , S = 1), and

    Sr2LiOsO6 (Os7+

    , S = 1/2) with a variety of the total spin quantum numbers were synthesized.

    Ca2InOsO6 and Ca3OsO6 crystallized in a monoclinic double perovskite structure (P21/n),

    whereas Sr2LiOsO6 crystallized in a tetragonal double perovskite structure (I4/m). The

    magnetic susceptibility and heat capacity measurements revealed that the compounds show

    manifest antiferromagnetic characters on cooling at temperatures of 14 K for Ca2InOsO6, 50

    K for Ca3OsO6, and 12 K for Sr2LiOsO6.

    In the last chapter, general conclusions and future prospects are discussed.

    Keywords:

    Osmium, Double perovskite, High-pressure synthesis, Crystal structure, Magnetic property.

  • III

    List of Abbreviations

    TGA Thermo-gravimetric analysis

    MPMS Magnetic property measurement system

    PPMS Physical property measurement system

    XRD X-ray diffraction

    SXRD Synchrotron X-ray diffraction

    FC Field cooling

    ZFC Zero field cooling

    TC Curie temperature

    TN Neel temperature

    TG Spin glass temperature

    HM Half-metallic

    FIM Ferrimagnetic

    DOS Density of states

    AF Antiferromagnetic

    FM Ferromagnetic

    Ea Activation energy

    VBG Valence bond glass

    NN Nearest neighbor

    t Tolerance factor

    ∏𝑐 Coulombic energy

    ∏𝑒 Exchange energy

    θ Weiss temperature

    C Curie constant

  • IV

    μeff Effective magnetic moment

    μS Spin only magnetic moment

    μS+L Magnetic moment from full spin and orbital motion

    BVS Bond valance sum

    ΘD Debye temperature

    γ Electronic specific heat coefficient

    Cp Heat capacity

    Clat Heat capacity of lattice contribution

    NIMS National Institute for Materials Science

    SO Spin-orbit

    EF Fermi energy

    J Nearest-neighbor exchange constant

    DFT Density functional theory

    MIT Metal-insulator transition

    λ Spin-orbit coupling constant

    2D Two-dimensional

    χ Magnetic susceptibility

    ρ Electrical resistivity

  • V

    High-Pressure Synthesis of Osmium Oxides with

    Double-Perovskite Structure and Their Magnetic Properties

    Contents

    Chapter 1. Introduction .............................................................................................................. 1

    1.1. Crystal structure .............................................................................................................. 1

    1.1.1. Perovskite and double perovskite structure .............................................................. 1

    1.1.2. Hexagonal-perovskite structure ................................................................................ 4

    1.2. Electronic structure .......................................................................................................... 5

    1.2.1. Crystal field considerations....................................................................................... 5

    1.2.2. The Jahn-Teller interaction ....................................................................................... 7

    1.3. Magnetism ....................................................................................................................... 8

    1.3.1. Origin of paramagnetic moments.............................................................................. 8

    1.3.2. Spin-orbit coupling ................................................................................................. 10

    1.3.3. Magnetic susceptibility and Currie-Weiss law ....................................................... 11

    1.3.4. Ferromagnetism and ferrimagnetism ...................................................................... 13

    1.3.5. Antiferromagnetism ................................................................................................ 14

    1.4. Theory of magnetic coupling......................................................................................... 15

    1.4.1. Superexchange ........................................................................................................ 15

    1.4.2. Double exchange ..................................................................................................... 17

    1.5. Magnetism in double perovskite oxides ........................................................................ 18

    1.5.1. Overview ................................................................................................................. 18

    1.5.2. High TC ferrimagnetic half-metal ............................................................................ 19

    1.5.3. High TC ferrimagnetic insulator .............................................................................. 21

    1.5.4. Other novel magnetic behavior ............................................................................... 22

    1.6. Objectives of this thesis ................................................................................................. 23

    References in chapter 1 ........................................................................................................ 24

    Chapter 2. Experimental methods ............................................................................................ 32

    2.1. Sample synthesis: high-pressure method ...................................................................... 32

  • VI

    2.2. X-ray diffraction ............................................................................................................ 33

    2.2.1. Powder X-ray diffraction ........................................................................................ 33

    2.2.2. Single crystal X-ray diffraction............................................................................... 33

    2.2.3. Synchrotron X-ray diffraction ................................................................................. 34

    2.3. Thermo-gravimetric analysis ......................................................................................... 34

    2.4. Magnetic properties measurement ................................................................................. 35

    2.5. Electrical properties measurement ................................................................................ 36

    2.6. Heat capacity ................................................................................................................. 36

    References in chapter 2 ........................................................................................................ 37

    Chapter 3. High-temperature ferrimagnetism of Ca2FeOsO6 and doping studies of

    Ca2-xSrxFeOsO6 and Ca2-xCdxFeOsO6...................................................................................... 38

    3.1. High-temperature ferrimagnetism driven by lattice distortion in double perovskite

    Ca2FeOsO6 ............................................................................................................................ 39

    3.1.1. Experimental details................................................................................................ 39

    3.1.2. Results and discussion ............................................................................................ 40

    3.2. Magnetic properties of Ca2-xSrxFeOsO6 and Ca2-xCdxFeOsO6 ...................................... 49

    3.2.1. Experimental details................................................................................................ 49

    3.2.2. Magnetic properties of Ca2-xSrxFeOsO6.................................................................. 50

    3.2.3. Magnetic properties of Ca2-xCdxFeOsO6 ................................................................ 56

    3.3. Summary of chapter 3 ................................................................................................... 59

    References in chapter 3 ........................................................................................................ 60

    Chapter 4. Crystal structure and magnetic properties of double perovskite oxide Ba2CuOsO6:

    a possible two-dimensional antiferromagnet ........................................................................... 64

    4.1. Experimental details ...................................................................................................... 65

    4.2. Results and discussion ................................................................................................... 66

    4.3. Summary of chapter 4 ................................................................................................... 78

    References in chapter 4 ........................................................................................................ 78

    Chapter 5. Crystal structure and magnetic properties of Ca2InOsO6, Ca3OsO6 and Sr2LiOsO6

    .................................................................................................................................................. 82

    5.1. Crystal structure and magnetic properties of Ca2InOsO6 .............................................. 82

  • VII

    5.1.1. Experimental details................................................................................................ 82

    5.1.2. Results and discussion ............................................................................................ 83

    5.2. Crystal structure and magnetic properties of Ca3OsO6 ................................................. 88

    5.2.1. Experimental details................................................................................................ 88

    5.2.2 Results and discussion ............................................................................................. 90

    5.3. Crystal structure and magnetic properties of Sr2LiOsO6 ............................................... 97

    5.3.1. Experimental details................................................................................................ 97

    5.3.2. Results and discussion ............................................................................................ 98

    5.4. Summary of chapter 5 ................................................................................................. 103

    References in chapter 5 ...................................................................................................... 105

    Chapter 6. General conclusions and future prospects ............................................................ 108

    6.1. General conclusions .................................................................................................... 108

    6.2. Future prospects .......................................................................................................... 111

    References in chapter 6 ...................................................................................................... 112

    List of appended publications ................................................................................................ 114

    Acknowledgement ................................................................................................................. 118

  • 1

    Chapter 1. Introduction

    1.1. Crystal structure

    1.1.1. Perovskite and double perovskite structure

    Perovskite oxides have the general formula ABO3, in which A is a large

    electropositive cation, B is a small transition metal or main group ion, and O is an oxygen ion

    [1]. The ideal perovskite structure is cubic with a space group of Pm3̅m, as shown in Figure

    1.1. The structure can be described as a frame of corner sharing BO6 octahedra, in which the

    center position is occupied by the A cation [2]. In the ideal structure, the equilibrium bond

    lengths (A-O) and (B-O) satisfy the relationship (𝑟𝐴 + 𝑟O) = √2(𝑟𝐵 + 𝑟O), where 𝑟𝐴, 𝑟𝐵 and

    𝑟O are effective ionic radii. In fact, the ideal cubic perovskite structure appears only in minor

    cases, and in major cases the bond lengths of (A-O) and (B-O) are geometrically incompatible,

    resulting in structure distortions and lower structure symmetry [3].

    Figure 1.1. ABO3 ideal perovskite structure [2].

  • 2

    To optimize the electric and magnetic properties, chemical substitution has been

    extensively studied, particularly for the B-site [4]. Generally, substitution of cation B′ to B

    leads to solid solution AB1-xB′xO3. But when x = 0.5, B-site cations may order when the

    charge and/or size of B and B′ atoms are sufficiently different. The formula then can be

    written as A2BB′O6 and the compound is commonly described as a double perovskite oxide

    when B and B′ are ordered. Through comprehensive survey of double perovskite oxides,

    Anderson et al. identified essential two types of B-sites ordering, rock-salt and layered

    ordering. The layered ordering type double perovskite is exemplified only by one compound

    [5]. Most of the double perovskite oxides are rock-salt type, in which cations B and B′ order

    into alternate octahedra. The ideal double perovskite oxide structure with rock-salt ordering

    (space group, Fm3̅m) is shown in Figure 1.2, which is obtained by doubled ideal perovskite

    by imposition of rock-salt ordering of B-sites. In this thesis, the structure of A2BB′O6 double

    perovskite oxides is concerned only with rock-salt ordering on the B-sites.

    Figure 1.2. A2BB′O6 ideal double perovskite structure with rock-salt B-sites ordering [1].

  • 3

    In the ideal double perovskite structure, the equilibrium bond lengths (A-O) and

    (B/B′-O) satisfy relationship (𝑟𝐴 + 𝑟O) = √2(𝑟𝐵+𝑟𝐵′

    2+ 𝑟O) , where 𝑟𝐴 , 𝑟𝐵 , 𝑟𝐵′ and 𝑟O are

    effective ionic radii. The double perovskite structure also allows a large degree of mismatch

    between the equilibrium (A-O) and (B/B′-O) bond lengths, which lowers structural symmetry.

    As a measure of deviation from the ideal situation, Goldschmidt introduced a tolerance factor

    (t), defined by equation: 𝑡 = 𝑟𝐴 + 𝑟O √2(𝑟𝐵+𝑟𝐵′

    2+ 𝑟O)⁄ , which are applicable to empirical

    ionic radii under ambient condition [1]. Clearly, in ideal cubic double perovskite structure the

    t-values should be very close to 1.

    Figure 1.3. A schematic diagram showing the group-subgroup relationships among the 12

    space groups for double perovskites. The dashed lines indicate the corresponding phase

    transition is first order as required by Landau theory [4].

    When the A cation is small, t < 1, the B-O and B′-O bonds suffer compressive stress

    and the A-O bonds suffer tensile stress. The stresses can be relieved by a cooperative rotation

    of BO6 and B′O6 octahedra [3], which deviates the B-O-B′ bond angles from 180º to 180 - Φ

    and lowers the coordination number of the A cations. Considering the combination with

  • 4

    ubiquitous BO6 (or B′O6) octahedral tilting, Howard et al. identified 12 different possible

    space groups by using a group-theoretical analysis [4]. The group-subgroup relationships

    among the 12 space group are schematically depicted in Figure 1.3, where the letters abc

    refer to tilts in the [100], [010], and [001] pseudo-cubic axes, and the superscript, 0,

    +, or

    -,

    indicates no tilt in an axis or tilts of successive octahedra in the same or opposite direction

    according to Glazer’s notation [6]. The repeated letters indicate equal tilts in the different

    pseudo-cubic axes.

    1.1.2. Hexagonal-perovskite structure

    When the tolerance factor of the perovskite structure decreases from unity, the cubic

    symmetry reduces to tetragonal or even monoclinic symmetry as a result of the tilting of the

    octahedra, however the crystal structure tends to evolve from cubic to hexagonal symmetry

    when the tolerance factor is larger than unity. Because when t > 1, the B-O bonds are under

    tensile stress and A-O bonds are under compressive stress, which cannot be relieved by tilting

    of BO6 octahedra; instead, large A cations can be accommodated in the space formed by

    face-shared BO6 octahedra columns. If all BO6 octahedra share faces to form one-dimensional

    columns along the c axis, then a 2H polytype is formed such as BaNiO3 and CsNiF3 [7,8]. In

    between the cubic (referred as 3C) and 2H, several polytypes of mixed cubic (c) and

    hexagonal (h) layers are identified, such as 9R (chhchhchh), 4H (chch), and 6H (cchcch) [9].

    The hexagonal-perovskite polytypes can be transformed gradually in the sequence of

    2H-9R-4H-6H-3C under high pressure as the high pressure stabilizes the high density phases.

    For example, BaRuO3, which crystallizes in 9R structure at ambient pressure, was

    transformed to 4H phase at 3GPa, the 6H phase at 5GPa, and the 3C phase at 18 GPa [10].

    For double-perovskite oxides A2BB′O6, when the tolerance factor is greater than unity, the

    crystal structure also tends to evolve from cubic to hexagonal symmetry at ambient pressure.

  • 5

    Unlike the perovskite that has various hexagonal-perovskite polytypes, the double perovskite

    oxides are found in 4H, 6H and 8H structures [11,12,13], particularly 6H structure is

    crystallized by most compounds. Figure 1.4 shows the 6H double perovskite structure, which

    consists of dimer units of face-shared octahedra (1/4 BO6 and

    3/4 B′O6) connected through

    common corners by a single layer of BO6 ocatahedra [12]. Double perovskites with

    hexagonal structure may transform to rock-salt type ordering double perovskite structures

    under high pressure [14].

    Figure 1.4. The double perovskite Ba2EuIrO6 with 6H structure [12].

    1.2. Electronic structure

    1.2.1. Crystal field considerations

    Crystal field theory is used to describe the electronic structures of metal ions in

    crystals, where they are surrounded by oxide ions or other anions, which create an

    electrostatic field with symmetry dependent on the crystal structure [15].

  • 6

    When the d orbitals of a transition-metal ion in a crystal with perovskite structure, it

    is surrounded by six oxygen ions, O2-

    , which give rise to the crystal field potential and hinder

    the free rotation of the electrons and quenches the orbital angular momentum by introducing

    the crystal field splitting of the d orbitals. The 𝑑𝑥2−𝑦2 and 𝑑𝑧2 orbitals, which are directed

    to the surrounding oxygen ions, are raised in energy. The dzx, dyz, and dxy orbitals, which are

    directed between the surrounding oxygen ions, are relatively unaffected by the field. The

    former 𝑑𝑥2−𝑦2 and 𝑑𝑧2 orbitals are called eg orbitals, whereas the latter, dzx, dyz, and dxy

    orbitals are called t2g orbitals as shown in Figure 1.5. The resulting energy difference is

    identified as ∆O (O for octahedral) [15,16].

    Figure 1.5. The d orbital splitting in octahedral field conditions [15,16].

    When the ligands orbitals strongly interact with the metal orbitals, the splitting

    between the t2g and eg orbitals is large (∆O is large), and the corresponding ligands are called

    strong-field ligands. Ligands with weak interactions with the metal orbitals are called

    weak-field ligands, and the split between the t2g and eg orbitals is small (∆O is small) [15].

    The orbital configuration of the electron is determined by the relationship among the ∆O,

    coulombic energy (∏𝑐), the exchange energy (∏𝑒). For each of d0 to d

    3 and d

    8 to d

    10 ions,

    only one electron configuration is possible, but for d4 to d

    7 ions, strong ligand fields lead to

  • 7

    low-spin state and weak ligand fields lead to high-spin state. Generally, the first transition

    metal ions, except Co3+

    , are in high spin state in their oxide compounds [16-18]. However,

    some of 3d transition metal oxides showed pressure-induced spin-state transitions , such as

    BiNiO3, BiMnO3, BiFeO3, and Fe3O4 [19-23]. By studying of LaCoO3, it is found that the

    spin state of Co3+

    is in low-spin 𝑡2g6 𝑒g

    0 (S = 0) state at lowest temperature, changes to

    intermediate-spin 𝑡2g5 𝑒g

    1 (S = 1) state in the interval 35 K < T < 100 K, changes to a mixture

    of intermediate spin and high-spin 𝑡2g4 𝑒g

    2 (S = 2) state in the range 300 K < T < 600 K, and

    then undergoing a transition from localized electrons to itinerant electrons [24].

    Compared with first transition metal series, one important characteristic of the

    second and the third transition metal series is that they tend to form low-spin compounds

    [25-28]. There are three main reasons for this intrinsically greater tendency to spin-pairing.

    First, the 4d and 5d orbitals are spatially larger than 3d orbitals so that double occupation of

    an orbital produces significantly less inter-electronic repulsion as compared to L-S coupling.

    Second, a given set of ligands atoms produces larger splitting of 5d than that of 4d and in

    both cases larger splitting than that for 3d orbitals [29]. Third, the spin-orbit coupling

    constant of 4d and 5d metal ions is much larger than 3d metal ions [30].

    1.2.2. The Jahn-Teller interaction

    The Jahn-Teller interaction is a theorem, which proves that any nonlinear molecular

    system in a degenerate electronic state would be unstable and existing some

    symmetry-breaking interactions. The symmetry-breaking associated with molecular distortion

    would remove the electronic degeneracy. In fact, the Jahn-Teller interaction is an example of

    electron-phonon coupling that there must be a multiplicity of electronic states interacting with

    one or more normal modes of vibration [31]. Examples of significant Jahn-Teller effects are

    found in complexes of Cr2+

    (d4), high-spin Mn

    3+ (d

    4), and Cu

    2+ (d

    9) [32-35]. For example,

  • 8

    octahedral Cu2+

    , a d9 ion, would have three electrons in the two eg levels without the

    Jahn-Teller effect, as shown in left of Figure 1.6a. The Jahn-Teller effect elongates the CuO6

    octahedra along z-axis. The effect of elongation on d orbital energies is shown in right of

    Figure 1.6a, where the 𝑑𝑧2 orbital is completely filled and the 𝑑𝑥2−𝑦2 orbital is half filled

    [15]. The half filled 𝑑𝑥2−𝑦2 orbital of Cu2+

    carries a spin S = 1/2 aligning within the

    ab-planes, which may result in two-dimensional antiferromagnetic behavior as shown in

    Figure 1.6b [36].

    Figure 1.6. (a) Diagram of electronic structures of Cu2+

    with Jahn-Teller interaction [15]; (b)

    Magnetic structure of Ba2CuWO6 [36]

    1.3. Magnetism

    1.3.1. Origin of paramagnetic moments

    Paramagnetism is generally caused by unpaired electrons in the substance of ions,

    atoms or molecules. Electrons determine the magnetic properties of matter in two ways.

    Firstly, the spinning of electron on its axis produces a magnetic moment. Secondly, an

    electron traveling in a closed path around a nucleus will also produce a magnetic moment

  • 9

    according to the pre-wave-mechanical picture of an atom. The magnetic properties of any

    individual atom or ion will result from some combination of these two contributions, which

    are the inherent spin moments of the electrons and the orbital moments resulting from the

    motion of the electrons around the nucleus [29].

    The magnetic moment due to the electron spins alone, according to wave mechanics,

    by the equation: μs = g√𝑆(𝑆 + 1) (unit is μB), in which S is the sum of the spin quantum

    numbers and g is the gyromagnetic ratio, also known as the “g factor”. For the free electron,

    g has the value 2.00023. The calculated magnetic moment μs due to the electron spins alone

    is called “spin-only” moment. In general, however, most of the 3d transition metal ions do

    possess orbital angular momentum. Wave mechanics show that for such cases, if the orbital

    motion makes its full contribution to the magnetic moments, they will be given by equation:

    μS+L = √4𝑆(𝑆 + 1) + 𝐿(𝐿 + 1), in which L represents the angular momentum quantum

    number for the ion [19,29].

    Table 1.1 listed magnetic moments experimentally observed for the common ions of

    the first transition series together with the calculated values of μS and μS+L. Generally, the

    observed values of μeff are close to or larger than μS, but smaller than μS+L. This is due to the

    Table 1.1. Effective moments and spin-orbit coupling constant associated with first series

    transition-metal ions [37]

    Ti3+

    V3+

    Cr3+

    Mn3+

    Mn2+

    , Fe3+

    Fe2+

    Co2+

    Ni2+

    Cu2+

    μS 1.73 2.83 3.87 4.0 5.92 4.90 3.87 2.83 1.73

    μeff (exp.) ~1.80 ~2.80 ~3.80 ~4.9 ~5.90 ~5.40 ~4.80 ~3.20 ~1.90

    μS+L 3.00 4.47 5.20 5.48 5.92 5.48 5.20 4.47 3.00

    λ [cm-1

    ] 154 105 91 88 - - - - -

  • 10

    electric fields surrounding the metal ions restrict the orbital motion of the electrons, so that

    the orbital angular momentums are wholly or partially “quenched” [29].

    1.3.2. Spin-orbit coupling

    For the first transition series ions, a simple interpretation of magnetic susceptibilities

    of the compounds usually gives the number of unpaired electrons as shown in Table 1.1. But,

    for the second and third transition series ions, the susceptibility data are often less easily

    interpreted with unpaired electrons. For instance, for OsIV

    complexes, which also have 𝑡2g4

    configuration, commonly have moments of the order of 1.2 μB or less [28, 29]; such a

    moment certainly does not give any simple indication of the presence of two unpaired

    electrons. This is caused by the interaction of the spin magnetic moment and the magnetic

    field arising from the orbital angular momentum, which is known as spin-orbit coupling. The

    strength of the spin-orbit coupling depends on the nuclear charge; the greater the nuclear

    charge the stronger the spin-orbit coupling. The coupling increases sharply with atomic

    number (as Z4) [38]. So there is stronger spin-orbit coupling in the second and third transition

    series ions than the first transition series ions. The spin-orbit coupling constant values are

    about 100~200 cm-1

    for first series transition metal ions as listed in Table 1.1, about

    1000~2500 cm-1

    for the second series transition metal ions and 3500~5000 cm-1

    for the third

    series transition metal ions as listed in Table 1.2.

    Table 1.2. Spin-orbit coupling constant (λ) for 4d and 5d shell ions [39, 40]

    4d shell ions 5d shell ions

    ions Ru3+

    Rh2+

    Rh3+

    Pd2+

    Ag2+

    Ag3+

    Cd3+

    Os4+

    Os5+

    Os6+

    λ [cm-1

    ] 1180 1220 1400 1600 1840 1930 2325 3600 4500 5000

  • 11

    Figure 1.7 shows how the effective magnetic moment of a 𝑡2g4 configuration

    depends on the ratio of the thermal energy kT to the spin-orbit coupling constant λ. For Mn3+

    and Cr2+

    , λ (as listed in Table 1.1) is sufficiently small such that at room temperature (kT

    ≈200 cm-1

    ) both these ions fall on the plateau part of the curve, where their behaviors are of

    the familiar sort. Os4+

    , however, has a spin-orbit coupling constant (~3600cm-1

    ) that is an

    order of magnitude higher, and at room temperature kT/λ is quite small. Thus at ordinary

    temperatures octahedral Os4+

    compounds should have low, strongly temperature-dependent

    magnetic moments [29].

    Figure 1.7. Curve showing the dependence on temperature and on the spin-orbit coupling

    constant λ of the effective magnetic moment of a d4 ion in octahedral coordination [29].

    1.3.3. Magnetic susceptibility and Currie-Weiss law

    When a material is placed in magnetic field H, a magnetization M is induced in the

    material which is related to H by M=χH. The χ is known as magnetic susceptibility of the

    material. Materials that have no unpaired electron orbital or spin angular momentum

    generally have negative values of χ and are called diamagnetic. Materials with unpaired

  • 12

    electrons, which are termed paramagnetic, have positive χ and show strong temperature

    dependence [41]. In classic studies, Pierre Curie showed that paramagnetic susceptibilities

    depend inversely on the temperature by the simple equation χ=C/T, where T represents the

    absolute temperature, and C is a constant that is characteristic of the substance and known as

    Curie constant. This equation is known as Curie’s law [42]. On theoretical ground, the

    magnetic field tends to align the moments of the paramagnetic atoms or ions, but at the same

    time thermal agitation tends to randomize the orientations of these individual moments [29].

    Materials in which there is no interaction between neighboring magnetic moments

    should obey Curie’s law. A more general Curie-Weiss law, χ = 𝐶 𝑇 − 𝜃⁄ , is an adapted

    version of Curie's law. In this equation, Weiss temperature (θ) can either be positive, negative

    or zero. When θ = 0 then the Curie-Weiss law equates to Curie’s law. If θ is positive then

    there is ferromagnetic interaction; if θ is negative then there is antiferromagnetic interaction.

    The Schematic diagrams of Currie’s law and Curie-Weiss law are shown in Figure 1.8. With

    Currie’s law or Curie-Weiss law, the effective moment can compute from Currie constant by

    using the equation μeff = 2.84√𝜒(𝑇 − 𝜃) = 2.84√𝐶 [42].

    Figure 1.8. Schematic diagrams of Curie’s law and Curie-Weiss law [42].

  • 13

    1.3.4. Ferromagnetism and ferrimagnetism

    In diamagnetic and paramagnetic materials, there is no magnetic order, whereas

    there are magnetic orders at low temperatures in ferromagnetic, antiferromagnetic and

    ferrimagenitc materials as shown in Figure 1.9.

    Figure 1.9. Schematic diagrams of paramgentic, ferromagnetic, simple antiferromagnetic and

    ferrimagnetic spin arrangement.

    The characteristic feature of ferromagnetic order is spontaneous magnetization due

    to spontaneous alignment of atomic magnetic moment. The critical temperature below which

    the spontaneous ordering occurs is called the Curie temperature (TC). When the temperature

    is higher than the Curie point, the substance become paramagnetic and can be described quite

    well by using Curie-Weiss law. According to the simple molecular-field treatment, the TC

    should be equal with the Weiss temperature obtained from the Curie-Weiss law. However, the

    experimental TC are frequently found to be somewhat smaller than Weiss temperature cause

    of that the effects of short-range order above TC are neglected in the simple molecular-field

    treatment [43].

    Ferrimagnetic materials also have non-zero magnetization below the critical

    temperature, which is also known as Curie temperature. The macroscopic magnetic

  • 14

    characteristics of ferrimagnet and ferromagnet are similar. However ferrimagnetic order is a

    special case of antiferromagnetic order with unequal moment in antiparallel arrangement as

    shown in Figure 1.9. So, the distinction between ferrimagnet and ferromagnet lies in the net

    magnetic moment [44].

    1.3.5. Antiferromagnetism

    In an antiferromagnet, the spins of magnetic electrons align in an antiparallel

    arrangement with zero net moment at temperature below the ordering. The critical

    temperature below which the antiferromagnetic order occurs is called the Neel temperature

    (TN). Above the Neel temperature, the material is paramagnetic. Below the Neel temperature

    the susceptibility generally decreases with decreasing temperature [44]. By studying the

    magnetic structures of manganite perovskites, it was found mainly three antiferromagnetic

    structure types, named A-Type, C-Type and G-Type as shown in Figure 1.10. Moreover, a

    complex spin and charge ordered antiferromagnetic phase, which is known as CE-Type, is

    also possible [45].

    Figure 1.10. Three antiferromagnetic structure types in perovskite

  • 15

    1.4. Theory of magnetic coupling

    1.4.1. Superexchange

    Superexchange is an indirect interaction between two magnetic cations (M1 and M2)

    via an intervening anion often O2-

    (2p6). The superexchange, was first proposed by Kramers

    [46], systematized by Anderson [47] and then refined by Goodenough [48] and Kanamori

    [49]. The 180º M1-O-M2 superexchange couples two cations on opposite sides of an anion.

    There is little direct overlap of the orbitals on the two cations, so the p-electrons of the O2-

    must be involved in the superexchange. The p-orbitals are described by p(x), p(y) and p(z),

    depending on the axis of rotation. These orbitals are classed into two types: (I) the pσ orbitals

    (p-orbital whose axis points to one of the cations as shown in Figure 1.11a) and (II) pπ

    orbitals (p-orbital whose axis is perpendicular to the line connecting the anion and cations as

    shown in Figure 1.11b). It is obvious that the pσ orbitals is orthogonal to the t2g orbitals, but

    not the eg orbital of principal overlap, and that the pπ orbitals are orthogonal to eg orbitals, but

    not the t2g orbital of principal overlap. Therefore electron transfer, or partial covalence, can

    only take place between pσ orbital and the eg orbital of principal overlap, or between pπ

    orbital and t2g orbital of principal overlap [43].

    Figure 1.11. Schematic diagrams of (a) pσ orbital and (b) pπ orbital [43].

  • 16

    There are three principal contributions to the superexchange: a correlation effect, a

    delocalization effect, and a polarization effect. The correlation mechanism, which takes into

    account the simultaneously partial-covalent bonds formation on each side of the anion, and

    delocalization mechanism, which assumes the electron drift from one cation to the other,

    contributes the majority interaction of superexchange [43]. According to the superexchange

    theory, the magnetic interaction between two magnetic ions mediated with superexchange

    interaction can be qualitatively predicated [50], whether the two spins are ferromagnetic

    coupling or antiferromagnetic coupling. As shown in Table 1.3, the superexchange

    interaction between d5 ions via oxygen was expected to be antiferromagnetic, which was

    confirmed by perovskite LaFeO3, which is an antiferromagnetic insulator with Neel

    temperature 740 K [51]. The superexchange interaction between d3 ions via oxygen was

    expected to be antiferromagnetic, which was confirmed by perovskite LaCrO3 with Neel

    temperature in the range 295-320 K [52,53]. Moreover, the Goodenough-Kanamori rules

    predicated that 180º superexchanges between electronic configurations d3-d

    5 bridged via

    oxygen is ferromagnetic. A possible combination is Fe3+

    and Cr3+

    introduced alternately on

    the B-sites of the perovskite (ABO3). Attempts to synthesize a ferromagnet with composition

    LaCr0.5Fe0.5O3 failed because the perovskite LaFe0.5Cr0.5O3 was found to be an

    antiferromagnetic material with TN about 265 K [54]. The disordered Sr2FeRuO6 exhibited

    spin-glass behavior below the freezing temperature of 60 K [55]. However, with new thin

    film fabrication techniques, it can deposit alternating monolayers of LaCrO3 and LaFeO3, and

    the resulting superlattice is indeed a ferromagnet with a Curie temperature of 375 K [50].

    Cause that the perovskite (AMO3) or double perovskite oxides structures always

    deviate from the ideal cubic structure to lower symmetry structures due to the equilibrium

    A-O and M-O bond lengths, so the M1-O-M2 bond angles deviate from 180º to 180º -Φ [56].

    Consequently, the linear M1-O-M2 superexchange changes to nonlinear superexchange

  • 17

    interactions. The bending of superexchange path would affect superexchange interactions.

    For instance, superexchange theory predicated that the d3-O-d

    5 superexchange interaction

    results in ferromagnetic coupling when the superexchange path is linear, and result in

    antiferromagnetic coupling when the bong angle deviates to range of 125º to 150º [43].

    Table 1.3. Sign of 180º M1-O-M2 superexchange between octahedral-site cations [43]

    Electronic

    configuration

    (Illustrative Cations)

    𝑡2g2 𝑒g

    0 𝑡2g3 𝑒g

    0 𝑡2g4 𝑒g

    0 𝑡2g3 𝑒g

    2 𝑡2g5 𝑒g

    0 𝑡2g4 𝑒g

    2 𝑡2g5 𝑒g

    2

    Ti2+

    ,V3+

    ,Ru6+

    Cr3+

    ,

    Mn4+

    ,

    Fe4+

    ,

    Ru4+

    ,

    Mn2+

    ,

    Fe3+

    Co4+

    ,

    Ru3+

    ,

    Fe2+

    ,

    Co3+

    Co2+

    ,

    Ni3+

    𝑡2g1 𝑒g

    0 Ti3+

    , Re6+

    ↑↓ ↑↓ ↑↓ ↑↑ ↑↑ ↑↑

    𝑡2g2 𝑒g

    0 Ti2+

    ,V3+

    ↑↓ ↑↓ ↑↓ ↑↑ ↑↓ ↑↑ ↑↑

    𝑡2g3 𝑒g

    0 Cr3+

    ,Mn4+

    ↑↓ ↑↓ ↑↑ ↑↓ ↑↑ ↑↑

    𝑡2g4 𝑒g

    0 Fe4+

    , Ru4+

    , ↑↓ ↑↑ ↑↓ ↑↑ ↑↑

    𝑡2g3 𝑒g

    2 Mn2+

    , Fe3+

    ↑↓ ↑↑ ↑↓ ↑↓

    According to superexchange theory, if a pair of cations is separated by two anions,

    both the delocalization and correlation superexchange are still possible but the magnitude is

    reduced, probably by an order of magnitude compared with M1-O-M2 superexchange

    interaction [43].

    1.4.2. Double exchange

    The double exchange interaction is a type of magnetic exchange that was proposed

    by Zener to explain the ferromagnetism in the mixed-valences manganite perovskite oxides,

    such as LaxCa1−xMnxIIIMn1−x

    IV O3 [57,58]. Figure 1.12 shows the schematic diagram of

    double exchange in mixed-valences manganite. If oxygen gives one of its spin-up electron to

  • 18

    Mn4+

    , its vacant orbital can then be filled by an electron from Mn3+

    . At the end of this process,

    an electron has transferred between the neighboring metal ions [59]. The double exchange

    interaction predicted that the extra electron on the Mn3+

    can travel back and forth between the

    two Mn ions only if the spins of the ions are parallel [56]. Besides the ferromagnetism in the

    mixed-valences manganite [59-63], a generalized double exchange mechanism can also be

    used to explain the high temperature half-metallic ferrimagnetism in Sr2FeMoO6 [64].

    Figure 1.12. Schematic diagram of double exchange in mixed-valences manganite [59]

    1.5. Magnetism in double perovskite oxides

    1.5.1. Overview

    Transition-metal perovskite oxides have been extensively studied since the 1940s.

    The discovery of high temperature superconductivity in copper oxides and magnetoresistance

    in manganites stimulated great interesting in investigations on transition metal oxides with

    perovskite and related structures [65-67].

  • 19

    Double perovskite oxides have a general formula A2BB′O6 and have been studied

    intensively because the crystal structure is robust in the wide composition range, such as

    Ba2LnMoO6 [68], Ba2LnOsO6 [69], Bi2NiMnO6 [70], A2MnWO6 (A = Ba, Pb) [71,72],

    Sr2BUO6 (B = Mn, Fe, Ni, Zn) [73], Sr2MnTeO6 [74], Sr2MnMO6 (M = Mo, W) [75],

    La2MTiO6 (M = Co, Ni) [76], A2CoTeO6 (A = Ca, Sr) [77], Ba2YMoO6[78] La2MIrO6 (M =

    Mg, Co, Ni, and Zn) [79], Sr2MReO6 (M = Ni, Co, Zn) [80], A2FeReO6 (A = Ca, Sr, Ba, Pb)

    [81-83], Sr2FeIrO6 [84], A2NiOsO6 (A = Ca, Sr) [85], Sr2MOsO6 (M = Cr, Co, Cu) [86-88],

    La2LiReO6, Ba2YReO6 [89], Ba2MOsO6 (M = Li, Na) [90,91], La2NaOsO6 [92], Sr2InReO6

    [93], and Sr2MReO6 (M = Mg, Ca) [94,95]. The range of properties varied from metals to

    insulators, ranging from ferromagnets, ferrimagnets, antiferromagnets, and ferroelectrics to

    spin liquids [96-99]. Previous studies indicated that double perovskite oxides A2BB′O6 exhibit

    extraordinary magnetic properties when B and B′ are occupied by 3d and 4d/5d ions,

    respectively. Particularly, the discovery of high TC ferrimagnetic transitions in Sr2FeMoO6

    and Sr2FeReO6 [97,100], which may be used as spintronics materials, stimulated the great

    interesting in studies on double perovskite oxides.

    1.5.2. High TC ferrimagnetic half-metal

    Perovskite oxides with high TC (around room temperature) ferromagnetic transition

    was firstly discovered in doped manganites in 1950 by Jonker and Van santen [101]. In these

    manganite compounds, mixed valence of Mn (Mn3+

    and Mn4+

    ) appears. The ferromagnetism

    in these manganites was explained via double exchange mechanism [57,58]. These

    discoveries encouraged further studies on perovskite oxide materials that could show high

    temperature ferromagnetism.

  • 20

    Figure 1.13. (a) The density of states (DOS) of Sr2FeMoO6 [97]. (b) The corresponding

    energy levels schematic diagram of Sr2FeMoO6. The Fermi level lies at the band formed

    exclusively by the Fe(t2g↓)-O(2p)-Mo(t2g↓) sub-band [96].

    In 1960s, high TC ferrimagnetic behavior was found in 3d-4d/5d hybrid double

    perovskite oxides A2FeMoO6 and A2FeReO6 (A = Ca, Sr, Ba) [102,103]. Unexpectedly, the

    A2FeMoO6 and A2FeReO6 compounds were found highly conductive [104,105]. Since the

    discovery of half-metallic ferrimagnetism in Sr2FeMoO6 with a relative high Curie

    temperature (~420 K), double perovskite oxides are receiving a great deal of renewed

    attentions [97]. In the HM state, 3d-t2g and 5d(4d)-t2g electrons play a pivotal role in the

    highly spin-polarized conduction [96,97]. In Figure 1.13, we note the half-metallic nature of

    this compound: the density of states (DOS) for the up-spin band forms a gap at the Fermi

    level, whereas the DOS for the down-spin band is at the Fermi level [96,97]. The down-spin

    band around the Fermi level is mainly occupied by the Mo 4d-t2g and Fe 3d-t2g electrons,

    which are strongly hybridized with oxygen 2p states [97]. The mechanism for high TC

    ferrimagnetic transition in 3d-4d/5d hybrid double perovskite oxides A2FeMoO6 and

    A2FeReO6 have been well studied [106-109]. The Alonso’s double exchange-like model

    succeeded in the stabilization of ferromagnetic state at high temperatures and half-metallicity,

  • 21

    but failed in the description of that TC is dependent on the band filling [96,108]. By deriving

    and validating a new effective spin Hamiltonian for these materials, Erten et al., presented a

    comprehensive theory in which a generalized double exchange mechanism can explain

    ferromagnetism with the scale of TC set by the kinetic energy of itinerant electrons [110].

    1.5.3. High TC ferrimagnetic insulator

    Figure 1.14. The calculated band-structure for Sr2CrOsO6 [86].

    In 2007, Krockenberger et al. reported a new 3d-5d hybrid double perovskite

    Sr2CrOsO6 which showed highest TC (around 725 K) among all perovskite oxides [86,111].

    Compared with high TC ferrimagentic transitions in half-metallic A2FeMoO6 and A2FeReO6

    (A = Ca, Sr, Ba), the observation of a higher TC in an insulator Sr2CrOsO6 is puzzling.

    Because Sr2CrOsO6 is a multi-orbital Mott insulator, the generalized double exchange

    mechanism was unable to account for the ferrimagnetic transition at TC of ~725 K in double

    perovskite oxide Sr2CrOsO6 [86,112]. Figure 1.14 shows a calculated band structure of

  • 22

    Sr2CrOsO6, indicating a small charge gap. There have been several density functional theory

    (DFT) calculations of Sr2CrOsO6 to study the mechanism of high TC ferrimagnetic transition

    [112-115]. However, the mechanism for high TC ferrimagentic transition in Sr2CrOsO6 is still

    unsure.

    1.5.4. Other novel magnetic behavior

    Besides the high TC ferrimagnetism, 3d-4d/5d hybrid double perovskites also show

    other novel magnetic behaviors, such as independent ordering of two interpenetrating

    magnetic sublattices in Sr2CoOsO6 [87]. In Sr2CoOsO6, the Os spins ordered

    antiferromagnetically below 108 K, while the Co spins ordered antiferromagnetically below

    70 K. Magnetic structure of Sr2CoOsO6 is shown in Figure 1.15. The DFT calculations

    indicated that the long-range Os-O-Co-O-Os and Co-O-Os-O-Co extended-superexchange

    interactions are stronger than the nearest-neighbor Os-O-Co superexchange interactions,

    which is contradicted with superexchange theory [43]. The discovery in Sr2CoOsO6 has

    broad implications for the magnetism in mixed 3d-5d transition-metal oxides [87].

    Figure 1.15. Magnetic structure of Sr2CoOsO6 [87].

  • 23

    Double perovskite oxides A2BB′O6 (B is nonmagnetic ion, B′ is 4d/5d transition

    metal ion) have attracted considerable attention recently due to the observation of a series of

    exotic magnetic states [92]. In double perovskites A2BB′O6, B′ ions can form a face centered

    cubic (fcc) lattice and it may form geometric frustration when magnetic interaction between

    B′ ions is dominated by nearest neighbor (NN) antiferromagnetic interaction (B′-O-O-B′)

    [116]. For instance, a valence bond glass (VBG) was observed in Ba2YMoO6 [78]; a

    collective singlet state in La2LiReO6; spin freezing in Ba2YReO6 [89], Sr2MgReO6 [94], and

    Sr2CaReO6 [95]; ferromagnetic Mott insulating state in Ba2NaOsO6 [90].

    1.6. Objectives of this thesis

    The aim of this thesis is to synthesize new osmium containing double perovskite

    oxides, which are useful to elucidate the mechanism of the high-TC FIM transition, as well as

    to explore notable magnetic behavior.

    (i) Crystal structure and magnetic properties of new 3d-5d hybrid double

    perovskite oxides.

    The high TC ferrimagnetism was observed in 3d-4d/5d hybrid double perovskite

    oxides such as A2FeMoO6, A2FeReO6 (A =Ca, Sr, Fe) and Sr2CrOsO6 [86,102,103]. To study

    the mechanism of the high-TC FIM transition, as well as to explore notable magnetic behavior,

    I have synthesize new 3d-5d hybrid double perovskite oxides Ca2FeOsO6 and Ba2CuOsO6.

    The novel 3d-5d hybrid double perovskite oxide Ca2FeOsO6 presents high temperature

    ferrimagnetic transition and is not a band insulator, but is electrically insulating like the

    recently discovered Sr2CrOsO6 (TC ~725 K). The electronic state of Ca2FeOsO6 is adjacent to

    half-metallic state as well as that of Sr2CrOsO6. The detailed crystal structure and magnetic

    properties of Ca2FeOsO6, and doping studies of Ca2-xSrxFeOsO6 and Ca2-xCdxFeOsO6 will be

  • 24

    discussed in Chapter 3. A compositionally new 3d-5d hybrid double perovskite oxide

    Ba2CuOsO6 was synthesized by using high pressure method. Because the existence of

    Jahn-Teller distortion in Cu2+

    , so it is expected two dimensional antiferromagnetic behaviors

    in Ba2CuOsO6 as reported in other Cu containing double-perovskite oxide [36]. The detailed

    crystal structure and magnetic properties of Ba2CuOsO6 will be discussed in Chapter 4.

    (ii) Crystal structure and magnetic properties of osmium containing double

    perovskite oxides A2BOsO6, where B is nonmagnetic ion.

    Double perovskite oxides A2BB′O6 (B is nonmagnetic ion, B′ is 4d/5d transition

    metal ion) have attracted considerable attention recently due to the observation of a series of

    exotic magnetic states [92]. To explore notable magnetic behavior in osmium double

    perovskite oxides, a series of osmium containing double-perovskite oxides Ca2InOsO6 (Os5+

    ,

    S = 3/2), Ca3OsO6 (Os6+

    , S = 1), and Sr2LiOsO6 (Os7+

    , S = 1/2) with different osmium

    valence states were synthesized by using high pressure method. The detailed crystal structure

    and magnetic properties of these oxides will be discussed in Chapter 5.

    References in chapter 1

    [1] M. T. Anderson, K. B. Greenwood, G. A. Taylor, K. R. Poeppelemier, Prog. Solid

    State Chem. 22, 197-233 (1993).

    [2] M. A. Pena, J. L. G. Fierro, Chem. Rev. 101, 1981-2017 (2001).

    [3] P. K. Davies, H. Wu, A. Y. Borisevich, I. E. Molodetsky, L. Farber, Annu. Rev. Mater.

    Res. 38, 369-401 (2008).

    [4] C. J. Howard, B. J. Kennedy, P. M. Woodward, Acta Cryst. B59, 463-471 (2003).

    [5] M. T. Anderson, K. R. Poeppelmerier, Chem. Mater. 3, 476-482 (1991).

    [6] A. M. Glazer, Acta Cryst. B28, 3384-3392 (1972).

    [7] Y. Takeda, F. Kanamaru, M. Shimada, M. Koizumi, Acta Cryst. B32, 2464-2466

  • 25

    (1976).

    [8] D. Babel, Structure and bonding, 3, 1-87 (1967) Berlin.

    [9] J.-G. Cheng, J. A. Alonso, E. Suard, J.-S. Zhou, J. B. Goodenough, J. Am. Chem. Soc.

    131, 7461-7469 (2009).

    [10] C.-Q. Jin, J.-S. Zhou, J. B. Goodenough, Q. Q. Liu, J. G. Zhao, L. X. Yang, Y. Yu, R.

    C. Yu, T. Katsura, A. Shatskiy, E. Ito, Proc. Natl. Acad. Sci. U.S.A. 105, 7115 (2008).

    [11] P. Kohl, D. Reinen, Z. Anorg. Allg. Chemie. 409, 257 (1974).

    [12] C. Lang, H. Muller-Buschbaum, J. Less Common Met. 161, 1-6 (1990).

    [13] J. Darriet, R. Bontchev, C. Dussarrat, F. Weill, B. Darriet, Eur. J. Solid State Inorg.

    Chem. 30, 287-296 (1993).

    [14] D. Iwanaga, Y. Inaguma, M. Itoh, J. Solid State Chem. 147, 291-295 (1999).

    [15] G. L. Miessler, D. A. Tarr, Inorganic Chemistry, Pearson Education (2004).

    [16] Y. Tokura, N. Nagaosa, Science 288, 462-468 (2000).

    [17] E. Rodriguez, M. L. Lopez, J. Campo, M. L. Veiga, C. Pico, J. Mater. Chem. 12,

    2798-2802 (2002).

    [18] R. M. Pinacca, M. C. Viola, J. C. Pedregosa, M. J. Martinez-Lope, R. E. Carbonio, J.

    A. Alonso, J. Solid State Chem. 180, 1582-1589 (2007).

    [19] M. Azuma, S. Carlsson, J. Rodgers, M. G. Tucker, M. Tsujimoto, S. Ishiwata, S. Isoda,

    Y. Shimakawa, M. Takano, J. P. Attfield, J. Am. Chem. Soc. 129, 14433-14436 (2007).

    [20] A. A. Belik, H. Yusa, N. Hirao, Y. Ohishi, E. Takayama-Muromachi, Inorg. Chem. 48,

    1000-1004 (2009).

    [21] A. A. Belik, H. Yusa, N. Hirao, Y. Ohishi, E. Takayama-Muromachi, Chem. Mater. 21,

    3400-3405 (2009).

    [22] Y. Ding, D. Haskel, S. G. Ovchinnikov, Y.-C. Tseng, Y. S. Orlov, J. C. Lang, H.-K.

    Mao, Phys. Rev. Lett. 100, 045508 (2008).

    [23] S. Ju, T.-Y. Cai, H.-S. Lu, C -D. Gong, J. Am. Chem. Soc. 134, 13780-13786 (2012).

  • 26

    [24] J.-Q. Yan, J.-S. Zhou, J. B. Goodenough, Phys. Rev. B 70, 014402 (2004).

    [25] M. Bremholm, S. E. Dutton, P. W. Stephens, R. J. Cava, J. Solid State Chem. 184,

    601-607 (2011).

    [26] Y. Singh, P. Gegenwart, Phys. Rev. B 82, 064412 (2010).

    [27] K. Yamaura, Y. Shirako, H. Kojitani, M. Arai, D. P. Young, M. Akaogi, M. Nakashima,

    T, Katsumata, Y. Inaguma, E. Takayama-Muromachi, J. Am. Chem. Soc. 131,

    2722-2726 (2009).

    [28] Y. Shi, Y. Guo, Y. Shirako, W. Yi, X. Wang, A. A. Belik, Y. Matsushita, H. L. Feng, Y.

    Tsujimoto, M. Arai, N. Wang, M. Akaogi, K. Yamaura, J. Am. Chem. Soc. 135,

    16507-16517 (2013)

    [29] F. A. Cotton, G. Wilkinson, Advanced Inorganic Chemistry, John Wiley&Sons (1980).

    [30] J. H. Choy, D. K. Kim, J. Y. Kim, Solid State Ionics 108, 159-163 (1998).

    [31] M. C. M. O’Brien, C. C. Chancey, Am. J. Phys. 61, 688-697 (1993)

    [32] J. T. Vallin, G. D. Watkins, Solid State Commun. 9, 953-956 (1971).

    [33] J. A. Alonso, M. J. Martinez -Lope, M. T. Casais, M. T. Fernandez -Diaz, Inorg. Chem.

    39, 917-923 (2000).

    [34] M. W. Lufaso, W. R. Gemmill, S. J. Mugavero III, S.-J. Kim, Y. Lee, T. Vogt, H.-C.

    Zur loye, J. Solid State Chem. 181, 623-627 (2008).

    [35] M. P. Attfield, P. D. Battle, S. K. Bollen, S. H. Kim, A.V. Powell, M. Workman, J.

    Solid State Chem. 96, 344-359 (1992).

    [36] Y. Todate, J. Phys. Soc. Jpn. 70, 337-340 (2001).

    [37] J. S. Griffith: The theory of transition-metal ions, Cambridge university press,

    Cambridge (1961).

    [38] P. Atkins, J. de Paula, Physical Chemistry (9th edition), W. H. Freeman and Company,

    New York (2010).

    [39] K. M. Mogare, W. Klein, H. Schilder, H. Lueken, M. Jansen, Z. Anorg. Allg. Chem.

  • 27

    632, 2389-2394 (2006).

    [40] M. Blume, A. J. Freeman, R. E. Watson, Phys. Rev. 134, A320-A327 (1964).

    [41] Landolt-Bornstein, Numerical Data and Functional Relationships in Science and

    Technology, New Series, Springer-Verlag, Heidelberg (1986).

    [42] C. Kittle, Introduction to solid state physics, 8th edition, Wiley (2004).

    [43] J. B. Goodenough, Magnetism and the chemical bond, Wiley: Cambridge, MA (1963).

    [44] Ferromagnetism and antiferromagnetism, www.phy.ilstu.edu.

    [45] W. Koshibae, Y. Kawamura, S. Ishihara, S. Okamoto, J.-I, Inoue, S. Maekawa, J. Phys.

    Soc. Jpn. 66, 957-960 (1997).

    [46] H. A. Kramer, Physica. 1,182-186 (1934).

    [47] P. W. Anderson, Phys. Rev. 79, 350-357(1950).

    [48] J. B. Goodenough, Phys. Rev. 115, 564-567 (1955).

    [49] J. Kanamori, J. Phys. Chem. Solids. 10, 87-98 (1959).

    [50] K. Ueda, H. Tabata, T. Kawai, Science, 280, 1064-1066 (1998).

    [51] G.R. Hearne, M. P. Pasternak, R.D. Taylor, P. Lacorre, Phys. Rev. B 51, 11495-11500

    (1995).

    [52] J.-S. Zhou, J. A. Alonso, A. Muonz, M. T. Fernandez-Diaz, J. B. Goodenough, Phys.

    Rev. Lett. 106, 057201 (2011).

    [53] I. Weinberg, P. Larssen, Nature, 4, 445-446 (1961)

    [54] A. K. Azad, A. Mellergard, S.-G. Eriksson, S. A. Ivanov, S. M. Yunus, F. Lindberg, G.

    Svensson, R. Mathieu, Mater. Res. Bull. 40, 1633-1644 (2005).

    [55] P. D. Battle, T. C. Gibb, C. W. Jones, J. Solid State Chem. 78, 281-293 (1989).

    [56] A. S. Bhalla, R. Guo, R. Roy, Mat. Res. Innovat, 4, 3-26 (2000).

    [57] C. Zener, Phys. Rev. 82, 403 (1951).

    [58] P. W. Anderson, H. Hasegawa, Phys. Rev. 100, 675-681 (1955).

    [59] J. M. B. Lopes dos Santos, V. M. Pereira, E. V. Castro, A. H. Neto, Advances in

  • 28

    Physical Science: meeting in honor of professor Antonio Leite Videira, (2005).

    [60] T. Taniguchi, S. Mizusaki, N. Okada, Y. Nagata, S. H. Lai, M. D. Lan, N. Hiraoka, M.

    Itou, Y. Sakarai, T. C. Ozawa, Y. Noro, H. Samata, Phys. Rev. B 77, 014406 (2008).

    [61] X. J. Fan, H. Koinuma, T. Hasegawa, Phys. B 329- 333, 723-724 (2003).

    [62] Y. Zhou, I. Matsubara, R. Funahashi, G. Xu, M. Shikano, Mater. Res. Bull. 38,

    341-346 (2003).

    [63] B. Raveau, Y. M. Zhao, C. Martin, M. Hervieu, A. Maignan, J. Solid State Chem. 149,

    203-237 (2000)

    [64] O. Erten, O. Nganba Meetei, A. Mukherjee, M. Randeria, N. Trivedi, P. Woodward,

    Phys. Rev. Lett. 107, 257201 (2011).

    [65] J. G. Bednorz and K. A. Müller, Z. Phys. B 64, 189-193 (1986).

    [66] R. Von Helmolt, J. Wecker, B. Holzapfel, L. Schultz, K. Samwer, Phys. Rev. Lett. 71,

    2331-2333 (1993).

    [67] S. Jin, T. H. Tiefel, M. McCormark, R. A. Fastnacht, R. Ramesh, L. H. Chen, Science

    264, 413-415 (1994).

    [68] E. J. Cussen, D. R. Lynham, J. Rogers, Chem. Mater. 18, 2855-2866 (2006).

    [69] Y. Hinatsu, Y. Doi, M. Wakeshima, J. Solid State Chem. 206, 300-307 (2013).

    [70] M. Azuma, K. Takata, T. Saito, S. Ishiwata, Y. Shimakawa, M. Takano, J. Am. Chem.

    Soc. 127, 8889-8892 (2005).

    [71] C. P. Khattak, D. E. Cox, F. F. Y. Wang, J. Solid State Chem. 17, 323-325 (1976).

    [72] J. Blasco, R. I. Merino, J. Garcia, M. C. Sanchez, J. Pys.: Condens. Matter. 18,

    2261-2271 (2006).

    [73] R. M. Pinacca, M. C. Viola, J. C. Pedregosa, M. J. Matinez-Lope, R. E. Carbonio, J. A.

    Alonso, J. Solid State Chem. 180, 1582-1589 (2007).

    [74] L. Ortega-San Martin, J. P. Chapman, L. Lezama, J. S. Marcos, J.

    Rodriguez-Fernandez, M. I. Arriortua, T. Rojo, Eur. J. Inorg. Chem. 7, 1362-1370

  • 29

    (2006).

    [75] A. Munoz, J. A. Alonso, M. T. Casais, M. T. Martinez-Lope, M. T. Fernandez-Diaz, J.

    Pys.: Condens. Matter. 14, 8817-8830 (2002).

    [76] E. Rodriguez, M. L. Lopez, J. Campo, M. L. Veiga, C. Pico, J. Mater. Chem. 12,

    2798-2802 (2002).

    [77] M. S. Augsburger, M. C. Viola, J. C. Pedregosa, A. Munoz, J. A. Alonso, R. E.

    Carbonio, J. Mater. Chem. 15, 993-1001 (2005).

    [78] M. A. de Vries, A. C. McLaughlin, J.-W. G. Bos, Phys. Rev. Lett. 104, 177202 (2010).

    [79] R. C. Currie, J. F. Vente, E. Frikkee, D. J. IJdo, J. Solid State Chem. 116, 199-204

    (1995).

    [80] M. Retuerto, M. J. Martinez-Lope, M. Garcia-Hernandez, M. T. Fernandez-Diaz, J. A.

    Alonso, Eur. J. Inorg. Chem. 4, 588-595 (2008).

    [81] N. Auth, G. Jakob, W. Westerburg, C. Ritter, I. Bonn, C. Felser, W. Tremel, J. Magn.

    Magn. Mater. 272-276, (2004).

    [82] W. Westerburg, O. Lang, C. Ritter, C. Felser, W. Tremel, G. Jakob, Solid State

    Commun. 122, 201-206 (2002).

    [83] K. Nishimura, M. Azuma, S. Hirai, M. Takano, Y. Shimakawa, Inorg. Chem. 48,

    5962-5966 (2009).

    [84] P. D. Battle, G. R. Blake, T. C. Gibb, J. F. Vente, J. Solid State Chem. 145, 541-548

    (1999).

    [85] R. Macquart, S. -J. Kim, W. R. Gemmill, J. K. Stalick, Y. Lee, T. Vogt, H.-C. Zur Loye,

    Inorg. Chem. 44, 9676-9683 (2005).

    [86] Y. Krockenberger, K. Mogare, M. Reehuis, M. Tovar, M. Jansen, G. Vaitheeswaran, V.

    Kanchana, F. Bultmark, A. Delin, F. Wilhelm, A. Rogalev, A. Winkler, L. Alff, Phys.

    Rev. B 75, 020404 (2007).

    [87] R. Morrow, R. Mishra, O. D. Restrepo, M. R. Ball, W. Windl, S. Wurmehl, U.

  • 30

    Stockert, B. Buchner, P. W. Woodward, J. Am. Chem. Soc. 135, 18824-18830 (2013).

    [88] M. W. Lufaso, W. R. Gemmill, S. J. Mugavero III, S.-J. Kim, Y. Lee, T. Vogt, H.-C.

    zur Loye, J. Sold State Chem. 181, 623-627 (2008).

    [89] T. Aharen, J. E. Greedan, C. A. Bridges, A. A. Aczel, J. Rodriguez, G. MacDougall, G.

    M. Luke, V. K. Michaelis, S. Kroeker, C. R. Wiebe, H. Zhou, L. M. D. Cranswick, Pys.

    Rev. B 81, 064436 (2010).

    [90] K. E. Stitzer, M. D. Smith, H.-C. zur Loye, Solid State Science, 4, 311-316 (2002).

    [91] A. J. Steele, P. J. Baker, T. Lancaster, F. L. Pratt, I. Franke, S. Ghannadzadeh, P. A.

    Goddard, W. Hayes, D. Prabhakaran, S. J. Blundell, Phys. Rev. B 84, 144416 (2011).

    [92] A. A. Aczel, D. E. Bugaris, L. Li, J.-Q. Yan, C. de la Cruz, H.-C. Zur Loye, S. E.

    Nagler, Phys. Rev. B 87, 014435 (2013).

    [93] H. Gao, A. Llobet, J. Barth, J. Winterlik, C. Felser, M. Panthofer, W. Tremel, Phys.

    Rev. B 83, 134406 (2011).

    [94] C. R. Wiebe, J. E. Greedan, P. P. Kyriakou, G. M. Luke, J. S. Gardner, A. Fukaya, I. M.

    Gat-Malureanu, P. L. Russo, A. T. Savici, Y. J. Uemura, Phys. Rev. B 68, 134410

    (2003).

    [95] C. R. Wiebe, J. E. Greedan, G. M. Luke, J. S. Gardner, Phys. Rev. B 65, 14413

    (2002).

    [96] D. Serrate, J. M. De Teresa, M. R. Ibarra, J. Phys. Condens. Matt. 19, 023201 (2007).

    [97] K. I. Kobayashi, T. Kimura, H. Sawada, K. Terakura, Y. Tokura, Nature 395, 677 -680

    (1998).

    [98] G. Chen, R. Pereira, L. Balents, Phys. Rev. B 82, 174440 (2010).

    [99] G. Chen, L. Balents, Phys. Rev. B 84, 094420 (2011).

    [100] L. Balcells, J. Navarro, M. Bibes, A. Roig, B. Martinez, J. Fontcuberta, Appl. Phys.

    Lett. 78, 781-783 (2001).

    [101] G. H. Jonker, J. H. Van Santen, Physica 16, 337-349 (1950).

  • 31

    [102] J. Longo, R. Ward, J. Am. Chem. Soc. 83, 2816-2818 (1961).

    [103] F. K. Patterson, C. W. Moeller, R. Ward, Inorg. Chem. 2, 196-198 (1963).

    [104] A. W. Sleight, J. F. Weiher, J. Phys. Chem. Solids 33, 679-687 (1972).

    [105] T. Nakagawa, J. Phys. Soc. Jpn. 24, 806-811 (1968).

    [106] E. Carvajal, O. Navarro, R. Allub, M. Avignon, B. Alascio, Eur. J. B 48, 179-187

    (2005).

    [107] A. Chattopadhyay, A. J. Millis, Pys. Rev. B 64, 024424 (2001).

    [108] J. L. Alonso, L. A. Fernandez, F. Guinea, F. Lesmes, V. Martion-Mayor, Phys. Rev. B

    67, 214423 (2003).

    [109] L. Brey, M. J. Calderon, S. Das Sarma, F. Guinea, Phys. Rev. B 74, 094429 (2006).

    [110] O. Erten, O. N. Meetei, A. Mukherjee, M. Randeria, N. Trivedi, P. Woodward, Phys.

    Rev. Lett. 107, 257201 (2011).

    [111] Y. Krockenberger, M. Reehuis, M. Tovar, K. Mogare, M. Jansen, L. Alff, J. Magn.

    Magn. Mater. 310, 1854-1856 (2007).

    [112] O. Nganba Meetei, O. Erten, M. Randeria, N. Trivedi, P. Woodward, Phys. Rev. Lett.

    110, 087203 (2013).

    [113] H. Das, P. Sanyal, T. Saha-Dasgupta, D. D. Sarma, Phys. Rev. B 83, 104418 (2011).

    [114] T. K. Mandal, C. Felser, M. Greenblatt, J. Kubler, Phys. Rev. B 78, 134431 (2008).

    [115] K.-W. Lee, W. E. Pickett, Phys. Rev. B 77, 115101 (2008).

    [116] P. D. Battle, T. C. Gibb, C. W. Jones, J. Solid State Chem. 78, 281-293 (1989).

  • 32

    Chapter 2. Experimental methods

    Double perovskite and perovskite structures are densely packed crystal structures, and are

    commonly found as high pressure phases [1,2]. So the high pressure synthesis is very

    effective to synthesize both polycrystalline samples and single crystal samples with double

    perovskite structures. In this thesis all the samples were synthesized by using high pressure

    and high temperature method. Then the crystal structures, transport properties, and magnetic

    properties of the as prepared samples were characterized. The various characterization

    methods and systems used in this thesis are presented in this chapter, such as X-ray

    diffraction (XRD), thermo-gravimetric analysis (TGA), magnetic property measurement

    system (MPMS), and physical property measurement system (PPMS).

    2.1. Sample synthesis: high-pressure method

    In this thesis, a belt-type high-pressure apparatus (Kobe Steel, Ltd.), which can press

    to 6 GPa and heat to 2000°C, was used to synthesize samples. Figure 2.1 shows the picture

    of the high-pressure apparatus and the schematic representation of the capsule and the sample

    container. To obtain quasi-hydrostatic conditions, we used pyrophyllite cell [3]. The samples

    can be heated by means of an internal graphite furnace. The Pt capsule is in the inside of

    graphite furnace, and is electrically insulated with graphite furnace by using NaCl sleeve and

    two cylindrical pieces.

    To prepare samples, stoichiometric amounts of starting materials were mixed

    thoroughly in an Ar-filled glove box. The mixtures were sealed into a Pt capsule by using

    hand press and the sealed Pt capsules were put into the sample cell as shown in Figure 2.1.

    The sample cell then was compressed to 6GPa by using the high-pressure apparatus. After the

  • 33

    pressure became stable, samples were being heated at various temperatures for 1 hour. Then

    the samples were subsequently quenched to ambient temperature before releasing the

    pressure.

    Figure 2.1. Image of high pressure apparatus set in National Institute for Materials Science

    (NIMS), and the schematic diagram of the capsule and sample container.

    2.2. X-ray diffraction

    2.2.1. Powder X-ray diffraction

    Powder X-ray diffraction (XRD), which is based on the Bragg’s law: nλ = 2dsinθ, is

    a powerful technique to determine the crystal structure [4]. In this thesis, powder XRD

    facility: Rigaku RINT 2200 (Cu Kα radiation) was used to identify the phase structure and

    phase quality.

    2.2.2. Single crystal X-ray diffraction

    The crystal structure of Ca3OsO6 was determined by using single crystal X-ray

    diffraction. A selected single crystal of Ca3OsO6, which was glued with epoxy onto the tip of

  • 34

    a thin glass fiber, was subjected to single crystal XRD analysis using a SMART APEX

    (Bruker) diffractometer (Mo Kα radiation, λ = 0.71073 Å). The XRD intensity data were

    collected at ambient temperature and pressure. SMART, SAINT+, and SADABS were used

    as the software packages for data acquisition, extraction/reduction, and empirical absorption

    correction, respectively [5]. Structure refinement was conducted on the intensity data by a

    full-matrix least-squares method using SHELXL-97 software [6].

    2.2.3. Synchrotron X-ray diffraction

    In Synchrotron X-ray Diffraction (SXRD), X-rays are generated by a synchrotron

    facility. The intensities of synchrotron radiation X-rays are several orders of magnitude

    higher than that of the best X-ray laboratory source. Synchrotron radiation was seen for the

    first time in 1947 in a particle accelerator (synchrotron). In the mid-1970s, scientists tried to

    produce extremely bright X-rays using synchrotrons. In 1990 European Synchrotron

    Radiation Facility (ESRF) was constructed, thereafter the Advanced Photon Source in the

    United States and SPring-8 in Japan were established. At synchrotron facility, electrons are

    usually accelerated by a synchrotron, and then injected into a storage ring, in which they

    circulate, producing synchrotron radiation. In this thesis, the synchrotron X-ray diffraction

    (SXRD) measurements were done in the BL15XU beam line (λ = 0.65297 Å) of the Spring-8

    synchrotron radiation facility in Japan. The BL15XU synchrotron-based X-ray diffraction

    facility uses the large Debye-Scherrer camera [7]. The obtained SXRD data were analyzed

    via the Rietveld method using the RIETAN-FP computer program [8,9].

    2.3. Thermo-gravimetric analysis

    Thermo-gravimetric analysis (TGA) is commonly used to determine the

    characteristics of selected materials that exhibit either mass loss or mass gain due to

  • 35

    decomposition or oxidation [10]. In this thesis, the TGA was used to check the oxygen

    contents of as prepared samples by measuring the mass loss. For example, when the sample

    Ca3OsO6 was heating up to 500 °C under hydrogen gas, a decomposition reaction

    Ca3OsO6+3H2→3CaO+Os+3H2O↑ was expected, and the oxygen contents of this sample

    could be estimated from the mass loss. A Perkin-Elmer Thermo-gravimetric analyzer was

    used to determine the oxygen content of the samples. During the measurement, a gas mixture

    of 5% hydrogen/argon was used. The samples were heating up to 500 °C with ramping rate of

    2.0 °C/min, and keeping 500 °C for 20-30 hours. The oxygen content was determined from

    the weight loss of the samples.

    2.4. Magnetic properties measurement

    Figure 2.2. Image of the MPMS-7T in NIMS Namiki-site

    The magnetic properties of all the samples were measured by using MPMS-7T

    (Quantum Design) as shown in Figure 2.2. The temperature dependent magnetic

    susceptibilities (χ) and field dependent isothermal magnetizations of all the as prepared

    samples were measured. The temperature dependent magnetic susceptibilities (χ) were

  • 36

    measured under both field cooling (FC) and zero field cooling (ZFC) conditions in a

    temperature range 2-400 K under an applied magnetic field of 10 kOe. The field dependence

    isothermal magnetizations were measured between +70 kOe and ‒70 kOe.

    2.5. Electrical properties measurement

    The electrical resistivity (ρ) of all the samples was measured by using PPMS-9T

    (Quantum Design) as shown in Figure 2.3. The measurements were performed by the usual

    four-terminal method to minimize the effects of contacting electrical resistivity. Electrical

    contacts on the four terminals were prepared by gold/Pt wires and silver paste. The

    temperature dependence resistivity was measured in a range of 2-300 K.

    Figure 2.3. Image of the PPMS-9T in NIMS Namiki-site. And a schematic diagram of the

    DC resistivity measurement Puck.

    2.6. Heat capacity

    Heat capacity is a measurable physical quantity of heat energy required to change the

    temperature of a material by a given amount. The specific heat capacity is the heat capacity

    per unit mass of a pure material and often simply called specific heat. In this thesis, the

  • 37

    temperature dependent heat capacities (Cp) of as prepared samples were measured with the

    PPMS-9T (as shown in Figure 2.3) using the relaxation method. The polycrystalline pellet,

    attached to the heat capacity platform with Apiezon N grease, was used to measure heat

    capacity. The sample chamber was pumped down to 0.01 mTorr to minimize the thermal

    contact with the environment.

    References in chapter 2

    [1] J. B. Goodenough, J. M. longo, Crystallographic and magnetic properties of perovskite

    and perovskite related compounds. Landolt-Bornstein Numerical Data and Functional

    Relationships in Science and Technology, Springer, Berlin (1970).

    [2] J. B. Goodenough, J. A. Kafalas, J. M. Longo, High-pressure Synthesis Preparation

    Methods in Solid State Chemistry, Academic Press, Inc., New York and London (1972).

    [3] J. Fernandez-Sanjulian, E. Moran, M. A. Alario-Franco, High Press. Res. 30, 159-166

    (2010).

    [4] J. M. Cowley, Diffraction physics, North-Holland, Amsterdam (1975).

    [5] SMART, SAINT+, and SADABS pacakages, Bruker Analytical X-ray Systems Inc.,

    Madison, WI, (2002).

    [6] G. M. Sheldrick, SHELXL97 Program for the Solution and Refinement of Crystal

    Structures, University of Gottingen, Germany (1997).

    [7] M. Tanaka, Y. Katsuya, and A. Yamamoto, Rev. Sci. Instrum. 79, 075106 (2008).

    [8] F. Izumi and K. Momma, Solid State Phenom. 130, 15-20 (2007).

    [9] F. Izumi, H. Asano, H. Murata, N. Watanabe, J. Appl. Cryst. 20, 411-418 (1987).

    [10] A. W. Coats, J. P. Redfern, Analyst, 88, 906-924 (1963).

    http://dx.doi.org/10.4028/www.scientific.net/SSP.130.15

  • 38

    Chapter 3. High-temperature ferrimagnetism of Ca2FeOsO6 and

    doping studies of Ca2-xSrxFeOsO6 and Ca2-xCdxFeOsO6

    Recently, 5d(4d)-3d hybrid double perovskite oxides (A2BB′O6) have attracted great attention

    due to the half-metallic (HM) nature found in oxides Sr2FeMoO6 [1-3] and Sr2FeReO6 [4-6]

    which exhibit spin-polarized transitions at temperatures greater than 400 K. In the HM state,

    3d-t2g and 5d(4d)-t2g electrons in the B and B′ sites, respectively, play a pivotal role in the

    highly spin-polarized conduction, which was argued to result from a generalized double

    exchange mechanism [7,8]. The high transition temperatures imply possible developments of

    novel spintronic applications, which may work without cooling. Furthermore, additional

    double perovskite oxides have been suggested to be HM at room temperature [4,9-20].

    Therefore, prospects for spintronics applications are very high and enormous efforts have

    been made not only to develop HM properties, but also to explore novel 5d(4d)-3d hybrid

    properties.

    The double perovskite oxide Sr2CrOsO6, which was synthesized in 2007, showed a

    ferrimagnetic (FIM) transition at TC of ~725 K [21], the highest reported TC of the perovskite

    and double perovskite oxides [13]. However, the generalized double exchange mechanism

    was unable to account for the 725 K transition because Sr2CrOsO6 is a multi-orbital Mott

    insulator and the FIM state is thus distinguishable from the HM state [22]. The high-TC FIM

    transition was novel, therefore, an additional 5d-3d hybrid high-TC FIM double perovskite

    oxide was highly desired to elucidate the mechanism of the high-TC transition.

    Here we report a novel double perovskite oxide, Ca2FeOsO6, synthesized using a

    high-pressure and high-temperature method. Ca2FeOsO6 crystallizes into a monoclinic double

    perovskite structure like many double perovskite oxides. Notably, it shows an FIM transition

    at a TC of 320 K. It is not a band insulator, but is highly insulating electrically. Thus, the

  • 39

    explanation for the high-TC FIM transition of Ca2FeOsO6 is likely similar to that of

    Sr2CrOsO6.

    3.1. High-temperature ferrimagnetism driven by lattice distortion in double

    perovskite Ca2FeOsO6

    3.1.1. Experimental details

    Polycrystalline sample of Ca2FeOsO6 were synthesized by solid state reaction under

    high pressure conditions. Powders of CaO2 (Lab made), Os (99.95%, Heraeus), Fe2O3

    (99.998%, Alfa Aesar), and KClO4 (99.5%, Kishida Chemical Co. Ltd.) were mixed in

    stoichiometric amounts using an agate mortar and a pestle in an Ar-filled dry box. This

    mixture was then placed in a one-side-sealed Pt capsule, and the Pt capsule sealed completely

    with a Pt cap using a hand press. The mixture was sealed in a Pt capsule and statically and

    isotropically compressed in a belt-type high-pressure apparatus (Kobe Steel, Ltd.) at a

    pressure of 6 GPa. The Pt capsule was heated at 1500 °C for 1 h, while maintaining the

    high-pressure conditions. The samples were subsequently quenched at ambient temperature

    before releasing the pressure.

    A sample of the as-prepared polycrystalline Ca2FeOsO6 was finely ground and

    characterized via synchrotron-based X-ray diffraction (SXRD) analysis using the large

    Debye-Scherrer camera at the BL15XU beam line of the SPring-8 synchrotron radiation

    facility (λ = 0.65297 Å) in Japan [23]. The SXRD data were analyzed via the Rietveld

    method using the RIETAN-FP computer program [24]. The crystal structure of Ca2FeOsO6

    was determined using the RIETAN-VENUS computer program [24].

    The magnetic susceptibility (χ) of the polycrystalline Ca2FeOsO6 was measured

    using the MPMS (Quantum Design) under field cooling (FC) and zero field cooling (ZFC)

    conditions in a temperature range 2-395 K at an applied magnetic field of 10 kOe. The field

    dependence of the isothermal magnetization was measured between +50 kOe and ‒50 kOe at

  • 40

    2 K and 300 K. Using a piece of the pellet, the electrical resistivity (ρ) was measured with a

    DC gauge current of 0.1 mA by a four-point method in a PPMS (Quantum Design, Inc.).

    Electrical contacts were prepared in the longitudinal direction by Pt wires and Ag paste.

    The electronic structure of Ca2FeOsO6 was studied by a generalized gradient

    approximation method of the density functional theory. We used the WIEN2k package, which

    was based on a full-potential augmented plane-wave method. Experimental lattice parameters

    and atomic coordinates were used for the calculation.

    3.1.2. Results and discussion

    Figure 3.1. Rietveld refined SXRD profiles of Ca2FeOsO6.

    Polycrystalline Ca2FeOsO6 was synthesized by solid-state reaction under

    high-pressure, and the final product was characterized using synchrotron X-ray diffraction.

    The SXRD study determined the crystal structure of Ca2FeOsO6 to be monoclinic with a

    space group of P21/n, as was reported for many B-site ordered double perovskites [25,26].

    The refined SXRD patterns are shown in Figure 3.1. Generally in a double perovskite

    6x106

    5

    4

    3

    2

    1

    0

    Inte

    nsi

    ty

    3632282420161282 /

  • 41

    structure, the degree of the B-site ordering depends on the differences of charge and Ionic

    radius of the B site ions, and the degree of the B-site ordering affects the magnetic properties.

    Thus, during the refinements, Fe and Os ions were assumed to be randomly mixed at the B

    site first. However, we were unable to reach a reasonable structure solution using these

    conditions. In subsequent refinements, the displacement parameters of Fe and Os were

    temporarily fixed at 0.5 Å2, the occupancies of Fe and Os (with Fe/Os mole ratio is 1:1) in B

    sites (2c and 2d site) were refined. Finally, the analysis indicated that Fe and Os atoms were

    ordered approximately 95% in the B-site. The small imperfection may be related to the

    proximity of the ionic radii of Fe3+

    (0.645 Å) and Os5+

    (0.575 Å) [27,28]. The detailed cell

    parameters and structure parameters of Ca2FeOsO6 are summarized in Tables 3.1 and Table

    3.2.

    Table 3.1. Structural parameters of Ca2FeOsO6 at room temperature.

    Atoms sites Occupancy x y z B(Å2)

    Ca 4e 1 0.9907(12) 0.0477(3) 0.2506(16) 0.77(5)

    Fe1/Os1 2c 0.95/0.05 0.5 0.0 0.0 0.78(4)

    Os2/Fe2 2d 0.95/0.05 0.5 0.0 0.5 0.22(2)

    O1 4e 1 0.0868(8) 0.4753(13) 0.2507(13) 0.60(2)

    O2 4e 1 0.7150(11) 0.2946(10) 0.042(3) 0.52(3)

    O3 4e 1 0.2021(10) 0.2149(10) 0.9525(19) 0.29(3)

    Space group: P21/n; Cell: a = 5.3931(6) Å, b = 5.5084(3) Å, c = 7.6791(3) Å, β = 90.021 (5)°,

    Z = 2; R indexes: Rp = 1.81, Rwp = 2.95.

    The average bond length of Os5+

    -O was ~1.956 Å, comparable to those of Os5+

    –O in

    other perovskite oxides (1.946 Å for NaOsO3 [30] and 1.960 Å for La2NaOsO6 [31]. The

    bond valance sum, calculated from the bond distances, were 3.00 for Fe and 4.74 for Os,

  • 42

    indicating 3+ and 5+ valence states of the ions, respectively [32,33]. Bond valance

    parameters RFe-O = 1.751, ROs-O = 1.868, and B = 0.37 were used in the estimation [33].

    Table 3.2. Selected interatomic distances and angles of Ca2FeOsO6

    Bond Bond distance (Å) Bond Bond angle (º)

    Fe1-O1 1.976(10) ×2 Fe1-O1-Os2 151.5(1)

    Fe1-O2 2.020(6) ×2 Fe1-O2-Os2 154.0(1)

    Fe1-O3 2.019(6) ×2 Fe1-O3-Os2 151.7(1)

    Os2-O1 1.986(10) ×2

    Os2-O2 1.936(7) ×2

    Os2-O3 1.946(6) ×2

    Figure 3.2. Crystal structure of Ca2FeOsO6, (a) along the [110] direction, and (b) along the

    [001] direction. The blue and brown octahedral represent FeO6 and OsO6, respectively. The

    solid spheres represent Ca.

    The average bond length of Fe3+

    -O (2.005 Å) was comparable or slightly larger than

    that of Ca2FeMoO6 (2.026 Å) Sr2FeOsO6 (1.938 Å), hence it was reasonable to conclude that

  • 43

    Fe3+

    of Ca2FeOsO6 has the high-spin state as well as Fe3+

    of Ca2FeMoO6 [34] and Sr2FeOsO6

    [35,36]. Note that a high-spin to low-spin state transition via intermediate-spin state of Fe3+

    in

    octahedral environment usually occurs at extremely high-pressure conditions far beyond 6

    GPa [37,38]. Comprehensive high-pressure studies of Fe oxides also imply the high-spin state

    for Fe3+

    of Ca2FeOsO6 [37,38]

    The crystal structure of Ca2FeOsO6 is shown in Figure 3.2, based on the

    experimental solution, showing that Fe and Os ions alternate to occupy the octahedral like a

    checker board. Bond angles of the inter-octahedral Fe-O-Os are 151° and 154°, which are far

    from 180°, indicating significant buckling of the octahedral connection. Figure 3.2 shows

    alternate rotations of the octahedra along and perpendicular to the c axis, respectively. The

    Glazer’s notation is a-a

    -b

    +, where the superscripts indicate that neighbor octahedra rotated in

    the same (+) and opposite (

    -) direction [39]. The degree of distortion is clearly more enhanced

    than that of Sr2FeOsO6, where the inter-octahedral Fe-O-Os are 180° and 165° [35].

    Figure 3.3. Isothermal magnetizations of Ca2FeOsO6 and Sr2FeOsO6 [35].

  • 44

    Isothermal magnetizations of Ca2FeOsO6 at 2 K and 300 K are compared with that

    of Sr2FeOsO6 in Figure 3.3. The spontaneous magnetization of Ca2FeOsO6 was 1.2 μB/f.u. at

    2 K (0.5 μB/f.u. at 300 K). This was in large contrast to the near zero magnetization of

    isoelectronic Sr2FeOsO6 [35]. The observed magnetization was, however, much lower than

    the theoretical sum of the spin only moment 8.0 μB/f.u. of Fe3+

    (t2g3 eg

    2) and Os

    5+ (t2g

    3). It was

    closer to the difference of each spin only moment (2.0 μB/f.u). Temperature dependent

    magnetic susceptibility in an applied field of 10 kOe is shown in Figure 3.4. A remarkable

    increase was observed approximately 320 K, indicating the occurrence of a long-range

    magnetic order. Specific heat measurements revealed that it is a bulk transition (see Figure

    3.5). The measurements therefore suggest that an FIM order is established at 320 K with

    cooling.

    Figure 3.4. Temperature dependence of magnetic susceptibility of Ca2FeOsO6.

  • 45

    A small disagreement between the observed moment (1.2 μB/f.u.) and the theoretical

    FIM spin-only moment (2.0 μB/f.u.) was partly due to the imperfect order of Fe/Os atoms.

    The disorder should decrease the magnetic moment as follows: M = Mexp × (1–2AS), where

    Mexp is the expected moment and AS is the degree of disorder between Fe and Os atoms

    [40-42]. In Ca2FeOsO6, Mexp was 2 μB/f.u. and AS was 0.05 (5% disorder), resulting in 1.8

    μB/f.u., approaching the observed moment of 1.2 μB/f.u. In addition, the lower observed

    magnetization moment may attributed to presence of possible Fe multiple oxidation states as

    observed in A2FeMoO6 based high temperature ferrimagnets [9].

    Figure 3.5. (a) Temperature dependence of the Cp of Ca2FeOsO6. (b) Linear fit to the low

    temperature part of the Cp/T vs. T2 curve; the least-squares analysis using the formula Cp/T

    =βT2 + where β and are a constant and the Sommerfeld coefficient, respectively, yielded β

    = 1.4310-4

    J mol-1

    K-4

    and = 0.14 mJ mol-1

    K-2

    .

  • 46

    The Figure 3.6a shows the temperature dependent electrical resistivity (ρ) of

    Ca2FeOsO6. At room temperature, the ρ was approximately 5 kΩ·cm and continued to

    increase upon cooling beyond the instrumental limit. The observed electrically insulating

    behavior i