Brought to you by:
Topical Review

Spin–orbit effects in pentavalent iridates: models and materials

and

Published 27 August 2021 © 2021 IOP Publishing Ltd
, , Citation Sayantika Bhowal and Indra Dasgupta 2021 J. Phys.: Condens. Matter 33 453001 DOI 10.1088/1361-648X/ac1aed

0953-8984/33/45/453001

Abstract

Spin–orbit effects in heavy 5d transition metal oxides, in particular, iridates, have received enormous current interest due to the prediction as well as the realization of a plethora of exotic and unconventional magnetic properties. While a bulk of these works are based on tetravalent iridates (d5), where the counter-intuitive insulating state of the rather extended 5d orbitals are explained by invoking strong spin–orbit coupling, the recent quest in iridate research has shifted to the other valencies of Ir, of which pentavalent iridates constitute a notable representative. In contrast to the tetravalent iridates, spin–orbit entangled electrons in d4 systems are expected to be confined to the J = 0 singlet state without any resultant moment or magnetic response. However, it has been recently predicted that, magnetism in d4 systems may occur via magnetic condensation of excitations across spin–orbit-coupled states. In reality, the magnetism in Ir5+ systems are often quite debatable both from theoretical as well as experimental point of view. Here we provide a comprehensive overview of the spin–orbit coupled d4 model systems and its implications in the studied pentavalent iridates. In particular, we review here the current experimental and theoretical understanding of the double perovskite (A2BYIrO6, A = Sr, Ba, B = Y, Sc, Gd), 6H-perovskite (Ba3MIr2O9, M = Zn, Mg, Sr, Ca), post-perovskite (NaIrO3), and hexagonal (Sr3MIrO6) iridates, along with a number of open questions that require future investigation.

Export citation and abstract BibTeX RIS

1. Introduction

Spin–orbit coupling (SOC) is a relativistic effect, which may be thought of as an interaction between the intrinsic spin moment of an electron and the magnetic field, generated in the rest frame of the electron due to its orbital motion around the positively charged nucleus. This interaction results in a coupling between the spin ($ \overrightarrow {S}$) and the orbital ($ \overrightarrow {L}$) angular momenta of an electron, lifting the (2S + 1)(2L + 1)-fold degeneracy of an atomic energy level, and hence the name 'SOC'. Mathematically, the coupling emerges naturally from the Dirac Hamiltonian and may be written as ${\mathcal{H}}_{\text{SOC}}=\frac{{\hslash }^{2}}{2{m}_{e}^{2}{c}^{2}}\frac{1}{r}\frac{\mathrm{d}V}{\mathrm{d}r}( \overrightarrow {L}\cdot \overrightarrow {S})=\lambda \overrightarrow {L}\cdot \overrightarrow {S}$, where the potential $V(r)=-\frac{Z{e}^{2}}{4\pi {\varepsilon }_{0}r}$ due to the nuclear charge Ze determines the coupling constant λ.

The strong dependence of λ on the atomic number Z [13], renders the inevitable spin–orbit effects in the heavy elements of the periodic table, leading to a wide variety of fascinating phenomena, which has continued to be the center of attention in the condensed matter research over a long period of time. This includes, for example, various topological effects [48], that stem from the non-trivial geometry of the electronic Bloch bands, providing not only the platform to realize the well known particles in the standard model of relativistic high-energy physics in the low energy excitations of the condensed matter systems, but also leading to the emergence of the field of quantum computation. In addition, there are Rashba and Dresselhaus spin–orbit interactions [911] due to a gradient of electrostatic potential in non-centrosymmetric systems, which constitute the heart of the present day hot topic of spin-orbitronics research [1220]. The present review, however, focuses on a different aspect of the SOC effects, namely, the interplay between SOC and Coulomb interaction in a crystalline solid, which often leads to unconventional phases of matter including the novel spin–orbit Mott insulating state.

While the importance of SOC in correlated systems had already come into notice [2123], the breakthrough came with the pioneering work of Kim et al [24] in the year 2008, where the authors have shown that the SOC is absolutely essential to describe the insulating state of Sr2IrO4, as the rather moderate Coulomb interaction alone cannot explain the opening of an energy gap in the extended Ir-5d states. As predicted by Jackeli and Khaliullin [25], this unique combination of SOC and Coulomb interaction can offer the long-awaited realization of the Kitaev model, that hosts a number of fascinating properties including anyonic excitations, topological degeneracy, etc, in real condensed matter systems, replacing the usual Heisenberg spin model. The idea of spin–orbit assisted Mott insulating state led to extensive studies over the past decade, in the search for new spin–orbit coupled correlated materials, their characterizations, as well as reinvestigation of previously synthesized materials in the light of the new insights on SOC.

Among these various spin–orbit coupled systems, the family of iridates have received considerable attention, primarily because of the availability of various crystal geometries as well as different Ir valencies within the same family. A large section of the studied iridates share the same Ir4+:d5 electronic configuration as the prototypical spin–orbit Mott insulator Sr2IrO4, where the different connectivities of the IrO6 structural unit resulting from diverse crystal geometries, give rise to a plethora of exotic electronic and magnetic properties of matter such as novel spin–orbit entangled metallic state, [2630], magnetic Kitaev interactions [3144], various topological states [28, 29, 4553], e.g., Dirac and Weyl semimetal, nodal line semi-metals, axion insulators, quantum spin liquid states [5456], hidden order multipolar state [43, 5760], superconductivity [61, 62], etc. A number of excellent reviews [6369] already exist in the literature, addressing many of these fascinating properties.

The present review, however, focusses on another charge state of iridium, the 5+ charge state, Ir5+ with d4 electronic configuration, known as pentavalent iridates. Similar to the tetravalent iridates, described above, pentavalent iridates also offer a wide range of structural variety. The magnetic properties of these iridates are, however, strikingly different owing to the dramatic effect of spin–orbit interaction on a local ${t}_{2g}^{4}$ electronic configuration. While in absence of SOC, degenerate ${t}_{2g}^{4}$ state results in a magnetic S = 1 state, in presence of strong SOC orbital and spin angular momenta are coupled to form a non-magnetic J = 0 state within an atomic limit. Locally this singlet J = 0 ground state is separated from the higher lying triplet J = 1 state by an energy scale that is determined by the strength of the SOC, λ. However, similar to the quantum dimer model [70], the superexchange interaction between the neighboring multiplets may cause a dispersion of the triplet state, facilitating the possibility of a magnetic excitation from J = 0 to J = 1 state, depending on the relative values of the energy scales involved. Such a magnetic excitation maybe viewed as annihilation of singlet bosons and the simultaneous creation of triplet bosons, known as triplon excitons. The prediction [71] of such excitonic magnetism in the nonmagnetic J = 0 ground state of d4 spin–orbit coupled systems motivated tremendous interest in the pentavalent iridates.

While the first d4 Ir-based material was synthesized about sixty years ago, around 1960s [72, 73], the interest in the d4 spin–orbit coupled materials is renewed with the recent advancement in the understanding of the SOC effects. In real solids, the situation is even more complicated in the presence of additional interactions such as non-cubic crystal field (ΔNCF), bandwidth (W), Hund's Coupling (JH), Coulomb interaction (U), etc, which often compete among each other. Indeed, the ground state of d4 iridates is determined by the delicate balance between these various competing interactions, leading to the emergence of a number of unconventional and exotic phases [7478]. The magnetic properties of such subtle ground state in real materials are, however, often quite debatable both from theoretical as well as experimental point of view.

The present review aims at providing an overview of the spin–orbit physics in d4 iridates, taking explicitly a few notable oxide members of this family into account. The idea of this review is to focus on some of the key issues in this field as well as put forward some open questions, inspiring future work along this direction. We first discuss the theoretical models, developed for d4 spin–orbit coupled systems, and, then, explicitly examine their implications by reviewing some of the reported d4 iridates, synthesized in the form of double perovskite, post perovskite, 6H perovskite, and hexagonal structures. Interestingly, the various crystal geometries also facilitate different connectivities of the IrO6 octahedra, e.g., double pervoskite and hexagonal iridates contain isolated IrO6 octahedron, while each IrO6 unit in post pervoskite structure shares its edge as well corner with the neighboring octahedron. On the other hand, IrO6 octahedra in 6H iridates are face-shared. Such diverse connectivities, in turn, provide the opportunity to investigate and compare the interplay between the different competing energy scales, governed by the specific structural geometries.

2. Background of the J = 0 state and iridates

In this section, we provide the background of the J = 0 state. In particular, we discuss the necessary criteria for the formation of the J = 0 state, explain how these requirements, often, can be met in iridates, thereby, making them good candidate materials to study, and briefly mention the possibility of excitonic magnetism, which we, further, discuss in detail in section 3, which has propelled enormous interest in d4 iridate research.

2.1. Formation of J = 0 state

The primary ingredient for the formation of J = 0 state is the strong SOC, typical for elements with large atomic number Z, where the electrons experience a strong nuclear potential, leading to unavoidable relativistic effect. This results in stronger SOC effects towards the end members of the periodic table. For example, in the transition metal (TM) series, the strength of the SOC gradually increases with an added n quantum number as we go down in the periodic table, viz, λ5d > λ4d > λ3d , as also schematically illustrated in figure 1(a).

Figure 1.

Figure 1. The formation of J = 0 state. (a) Schematic illustration of the different energy scales of 3d, 4d and 5d orbitals, where the base of the triangle denotes the band-width W of the respective orbitals situated at an increasing height from the base. The increasing and decreasing values of the different parameters: Coulomb correlation U, Hund's coupling JH, SOC λ, and crystal field splitting Δ along the height of the triangle are indicated by big arrows. (b) Radial distribution ${r}^{2}{R}_{nl}^{2}(r)$ for 3d, 4d and 5d orbitals as a function of distance from the nucleus r, indicating that the 5d orbitals are more spatially extended than the rest, leading to larger W as shown in (a). (c) The evolution of the ground state of a d4 atomic system in an ideal octahedral environment, depending on the relative energy scales of the parameters. Left: the magnetic high spin S = 2 state in the JH > Δ regime, relevant for 3d orbitals. Middle: the magnetic low spin S = 1 state in the Δ > JH regime, typical for 4d orbitals in presence of weak SOC. Right: spin–orbit entangled non-magnetic J = 0 state in presence of large SOC, expected for heavy 5d orbitals. See text for details.

Standard image High-resolution image

Interestingly, going from 3d to 5d TM elements, the d orbitals get spatially more extended (see figure 1(b)). As a result, the wide 5d orbitals hybridize strongly with the neighboring ligand atoms, leading to a larger crystal field splitting in 5d TM compounds compared to the systems with narrow 3d bands. This has some interesting consequences, which play crucial role in the formation of J = 0 state in a d4 electronic configuration. For a simple illustration, we consider TM-O6 octahedral unit. Indeed such units are quite common in various TM oxides, particularly in iridates. We consider an ideal octahedral symmetry, which, then, splits the five-fold d states into a triplet t2g and a higher lying doublet eg orbitals. First, let us consider the case when the TM has 3d orbitals in the valence state. Due to weak crystal field splitting (Δ) in these systems, the Hund's coupling JH that favors parallel spin alignment of the electrons in different d orbitals win, and 3d4 electronic configuration leads to a high spin S = 2 state (${t}_{2g}^{3}{e}_{g}^{1}$), as seen for example in LaMnO3 [79]. The SOC effects being rather weak, the crystal field split states are only very weakly perturbed by SOC. In case of 4d TM elements, however, the 4d orbitals being wider than the 3d orbitals, Δ is relatively larger in magnitude. In this case, Δ wins over the Hund's coupling JH, resulting in a low spin state S = 1, with four electrons in the t2g orbitals (${t}_{2g}^{4}{e}_{g}^{0}$) [80, 81]. The formation of the low spin state is the first step towards the J = 0 state. However, due to its moderate value in 4d TMs, often SOC acts only as a perturbation to the t2g states. It is interesting to point outhere that the recent report of a J = 0 ground state in 4d based Ru compound K2RuCl6within LS coupling scheme points to the importance of SOC (beyond perturbation) also in 4d based systems [82]. Nevertheless, the remaining criteria of strong SOC is, however, satisfied in 5d TMs, where strong SOC forms spin–orbital entangled states, destroying the individual orbital and spin identities of the constituents.

We pause here and discuss the formation of these new spin–orbit entangled states within a single particle formalism. Assuming the t2g states are completely decoupled from the higher lying eg orbitals, the spin–orbit entangled states can be immediately visualized by diagonalizing the SOC Hamiltonian in the t2g basis, viz,

Equation (1)

where ${\mathcal{H}}_{+}$ (${\mathcal{H}}_{-}$) in the basis {|yz↓⟩, |zx↓⟩, |xy↑⟩} ({|yz↑⟩, |zx↑⟩, |xy↓⟩}) has the following form

Equation (2)

We have assumed, here, degenerate t2g states in an ideal octahedral environment. The diagonalization of ${\mathcal{H}}_{\text{SOC}}$ yields four-fold degenerate Γ8 (−$\frac{\lambda }{2}$) and two-fold degenerate Γ7 (λ) states, where each of the eigenstates of ${\mathcal{H}}_{+}$ has its time-reversal counterpart in the corresponding eigenstate of ${\mathcal{H}}_{-}$, forming the Kramer's doublet. The explicit form of these eigenstates are given by:

and,

Note that the eigenstates above do not have pure spin or orbital characters any more, as they couple to form entangled states. In the literature, these quartet and doublet are commonly referred to as jeff = 3/2 and jeff = 1/2 states respectively. It is important to point out that these spin–orbit entangled effective states are obtained within a single particle formalism and they are different from the many-body states J = 1/2 or 3/2. To make this distinction clear, now on we represent the single particle states by j, while the many-body states are indicated as J.

If we now go ahead and fill up these spin–orbit coupled states with four electrons, present in the 5d shell, we can see that the lower lying jeff = 3/2 state gets completely occupied while the jeff = 1/2 state remains empty. As there is no unpaired electron left in the system, the system becomes non-magnetic. The effective total angular momentum, jeff for this state may be determined from the effective orbital angular momentum leff = −1 (see appendix for the illustration of this point) and the spin angular momentum s = 1, which again give a non-magnetic jeff = leff + s = 0 state. The evolution of the ground state of a d4 system in the different regimes of the competing energy scales is schematically shown in figure 1(c). Note that the formation of J = 0 state within the many-body framework can similarly be described assuming t2g only model, which we discuss in detail later in the next section.

As an aside, it is interesting to point out that for five electrons in the 5d shell, the jeff = 1/2 state gets partially occupied in addition to the completely filled jeff = 3/2 state. This half-filled jeff = 1/2 state is further split into empty upper and filled lower Hubbard bands in presence of Coulomb interaction, leading to the well known jeff = 1/2 Mott insulating state [24]. In contrast to the case of four electrons, discussed above, here the system remains magnetically active. The role of SOC, here, is to just convert the exchange interactions between spin and orbital angular momentum into an effective total angular momentum. These interactions between the total angular momenta, however, may have interesting consequences, for example they could be bond dependent, inherited from the corresponding interactions between orbital angular momentum for the t2g orbitals [83].

2.2. Origin of excitonic magnetism in spin orbit coupled d4 systems

The non-magnetic J = 0 ground state, imposed by SOC, is very typical for Van-Vleck type paramagentism, where the admixture of the ground state |ψ0⟩ with the magnetic excited states |ψn ⟩ gives rise to a magnetic response in presence of a magnetic field ($ \overrightarrow {B}$). In the language of perturbation theory, this can be expressed as $\vert \psi \rangle =\vert {\psi }_{0}\rangle +{\sum }_{n}\;\frac{\langle {\psi }_{n}\vert ( \overrightarrow {B}\cdot \overrightarrow {J})\vert {\psi }_{0}\rangle }{{E}_{n}-{E}_{0}}$ [84]. The essence of the behavior of a singlet–triplet system in presence of a magnetic field is schematically illustrated in figure 2.

Figure 2.

Figure 2. The simplified picture of the magnetic-response of a singlet–triplet system in presence of a magnetic field $ \overrightarrow {B}$. In absence of a magnetic field, $ \overrightarrow {B}=0$, the singlet |J = 0⟩ and the triplet |J = 1⟩ states are separated by an energy gap of Δst, while in presence of a magnetic field the triplet state |J = 1⟩ undergoes a Zeeman splitting, leading to splitting of the |Jz = ±1⟩ states, with energies E = Δ ∓ gJ μB B. When the strength of the applied magnetic field B exceeds a critical value Bc = Δ/gJ μB , the |Jz = +1⟩ state lies below the J = 0 state, leading to a sharp magnetic transition. In reality, the behavior of the singlet state depends on the matrix elements of the magnetic dipoles between the non-magnetic ground state and the magnetic excited state.

Standard image High-resolution image

2.2.1. Why iridates?

The 5d TM oxides, of which iridates are notable members, facilitate the basic requirements for the formation of a non-magnetic J = 0 state. In such a non-magnetic state, however, magnetism may appear by a mechanism very similar to Van-Vleck paramagentism via excitation to the higher lying triplet J = 1 state, as suggested by Khaliullin [71]. The possibility of such excitation is governed by the relative energy scales of the singlet–triplet splitting (Δst), determined by the SOC constant λ, and the superexchange energy, which is ∼4t2/U. When these two energy scales are comparable, singlet–triplet Van Vleck transitions may occur, resulting in a magnetism in the apparently non-magnetic J = 0 state.

2.2.2. Singlet-triplet transition: excitonic magnetism

Following Khaliullin's idea [71] of singlet–triplet transition, here, we briefly illustrate this process considering 180°d–p–d bond geometry, which is relevant for the corner-shared octahedra. In the 180°d–p–d bond geometry, as shown in figure 3, only two of the three t2g orbitals are active along a certain bond, while the third orbital, say c, remains inactive. For simplicity, we can call this bond c. Therefore, the nearest neighbor hopping matrix along this c bond is given by $t({a}_{i}^{{\dagger}}{a}_{j}+{b}_{i}^{{\dagger}}{b}_{j}+\mathrm{H}.\mathrm{c}.)$, indicating that the hopping is diagonal in the active orbital space. Here ${a}_{i}^{{\dagger}}({a}_{i}),{b}_{i}^{{\dagger}}({b}_{i})$ are the creation (annihilation) operators that create (annihilate) an electron in the respective active t2g orbitals a and b at site i. In the illustration of figure 3, these active orbitals are dxz and dxy orbitals for the bond along x. We now introduce three different operators A, B, and C, which represent the three possible orbital configurations for the ${t}_{2g}^{4}$ shell, viz, A = {a2 bc}, B = {ab2 c}, and C = {abc2}. The one-to-one correspondence between the operators and the orbitals are apparent from this definition. With the help of these operators, we can also rewrite the components of the orbital angular momentum operator, e.g., Lx = −i(B CC B), etc.

Figure 3.

Figure 3. The schematic illustration of the 180°d–p–d bond. The figure shows the two active orbitals, dxz and dxy , along a bond in the x-direction for corner shared octahedral geometry.

Standard image High-resolution image

In the strong coupling limit (Ut, λ, JH), this results in a spin–orbital Hamiltonian, having the form $H=({t}^{2}/U){\sum }_{\langle ij\rangle }[({ \overrightarrow {S}}_{i}\cdot { \overrightarrow {S}}_{j}+1){O}_{ij}^{(\gamma )}+{({L}_{i}^{\gamma })}^{2}+{({L}_{j}^{\gamma })}^{2}]$, where O(γ) is the bond dependent orbital operator which can be written in terms of the A, B, C operators or directly expressed in terms of the orbital angular momentum operators Lx , Ly , and Lz . For example, the operator for the bond c can be written as ${O}^{(c)}={({L}_{i}^{x}{L}_{j}^{x})}^{2}+{({L}_{i}^{y}{L}_{j}^{y})}^{2}+{L}_{i}^{x}{L}_{i}^{y}{L}_{j}^{y}{L}_{j}^{x}+{L}_{i}^{y}{L}_{i}^{x}{L}_{j}^{x}{L}_{j}^{y}$. Similarly, the operators O(a), O(b) for other bonds a, and b can be obtained from cyclic permutations of Lx , Ly , and Lz . Clearly, the Hamiltonian H operates in the (ML , MS ) basis. We now want to project it onto the low-energy subspace of the singlet J = 0 ground state, $\vert 0\rangle =\frac{1}{\sqrt{3}}(\vert 1,-1\rangle -\vert 0,0\rangle +\vert -1,1\rangle )$ and the first excited state J = 1 triplet at an energy Δst determined by λ, $\vert {T}_{0}\rangle =\frac{1}{\sqrt{2}}(\vert 1,-1\rangle -\vert -1,1\rangle ),\;\vert {T}_{{\pm}1}\rangle ={\pm}\frac{1}{\sqrt{2}}(\vert {\pm}1,0\rangle -\vert 0,{\pm}1\rangle )$. Calculating the matrix elements of the orbital and spin angular momentum operators in this Hilbert space, we can write them in terms of the 'triplon' $ \overrightarrow {T}$ with the Cartesian components ${T}_{x}=\frac{1}{i\sqrt{2}}({T}_{1}-{T}_{-1}),\enspace {T}_{y}=\frac{1}{\sqrt{2}}({T}_{1}+{T}_{-1})$, and Tz = iT0, and the 'spin' $ \overrightarrow {J}=-i({ \overrightarrow {T}}^{{\dagger}}{\times} \overrightarrow {T})$. For example, $ \overrightarrow {S}=-i\sqrt{\frac{2}{3}}( \overrightarrow {T}-{ \overrightarrow {T}}^{{\dagger}})+\frac{1}{2} \overrightarrow {J}$, and $ \overrightarrow {L}=i\sqrt{\frac{2}{3}}( \overrightarrow {T}-{ \overrightarrow {T}}^{{\dagger}})+\frac{1}{2} \overrightarrow {J}$. Using these relations we can immediately write down the magnetic moment for a ${t}_{2g}^{4}$ configuration in terms of $ \overrightarrow {T}$ and $ \overrightarrow {J}$, viz, $ \overrightarrow {M}=2 \overrightarrow {S}- \overrightarrow {L}=-i\sqrt{6}( \overrightarrow {T}-{ \overrightarrow {T}}^{{\dagger}})+{g}_{J} \overrightarrow {J}$, where ${g}_{J}=\frac{1}{2}$. Clearly, the magnetism in a J = 0 state is driven solely by the $ \overrightarrow {T}$ exciton as the second term in $ \overrightarrow {M}$ does not contribute for a singlet state, J = 0, and hence, the magnetism is referred to as 'excitonic magnetism'.

The transformation $H(L,S)\to H( \overrightarrow {T}, \overrightarrow {J})$, as described above, gives us the desired effective singlet–triplet model [71],

Equation (3)

Here κ denotes the exchange interaction 4t2/U and terms up to quadratic in $ \overrightarrow {T}$ bosons are considered. The three and four-boson interaction terms, that include the quadrupole operators and bi-quadratic Heisenberg exchange couplings respectively, have only small values near the phase transition for 180° bond geometry and are, therefore, not important for the qualitative understanding of magnetic phase transition driven by the condensation of $ \overrightarrow {T}$ bosons. The diagonalization of this effective Hamiltonian gives the energy dispersion for the triplet excitation, which is useful to understand the condensation of the triplet state. For example, when κ < κc, where κc is some critical value of κ, the magnetic excitation has a finite gap $\lambda \sqrt{1-(\kappa /{\kappa }_{\mathrm{c}})}$, indicating a paramagnetic phase. At κ = κc this gap closes and a magnetic phase transition takes place due to condensation of the $ \overrightarrow {T}$ bosons.

Interestingly, soon after this proposal of excitonic magnetism, Cao et al [74] experimentally showed the existence of magnetism in the d4 double perovskite iridate Sr2YIrO6. Indeed, in most of the pentavalent iridates, the Ir atom has a non-zero effective magnetic moment (see, for example, the list of pentavalent iridates with their magnetic and trasport properties in table 1), where the moment values are often attributed to the non-cubic crystal field, bandwidth effect, etc in the solid causing the deviation from the atomic like picture, described above for the formation of J = 0 state. Furthermore, since the strength of the SOC is renormalized in a solid as λ/2S [84, 85], where λ is the atomic SOC, and S is the total spin, one might anticipate the SOC strength would be weaker in d4 systems with S = 1 as compared to tetra-valent iridates with S = 1/2. Such weak SOC tends to favor excitonic magnetism in d4 iridates as the relative strength of the SOC constant and the superexchange energy scale determines the occurrence of excitonic magnetism. Interestingly, a recent theoretical study [86] based on density matrix renormalization also suggests spin–orbit induced transition to an excitonic anti-ferromagnetic condensate even at an intermediate Coulomb interaction regime.

Table 1. Structural, transport and magnetic properties of pentavalent iridates. The notations FM, AFM and SOL refer to ferromagnetic, anti-ferromagnetic and spin–orbital liquid states respectively. The connectivity of IrO6 unit in each structural family is given within the parenthesis.

Crystal structureCompoundSpace groupTransport propertyMagnetic ordering μeff/Ir (μB )Reference
Double perovskite (isolated)Sr2YIrO6 P21/n InsulatorAFM0.91[74, 75]
  $Fm\overline{3}m$ InsulatorNone0.21[94]
Ba2YIrO6 $Fm\overline{3}m$ InsulatorNone0.3[75, 77]
  InsulatorNone0.44[8792]
  P21/n [95]
  InsulatorAFM1.44[93]
Sr2GdIrO6 Pn-3InsulatorFM[74, 75]
Sr2ScIrO6 P21/n InsulatorNone0.16[9698]
Sr2LuIrO6 P21/n None[95]
Ba2ScIrO6 $Fm\overline{3}m$ InsulatorNone0.39[96, 97]
Ba2LuIrO6 P21/n None[95]
Ba2LaIrO6 P21/n None[95]
  $R\overline{3}$ [99]
Ba2NdIrO6 Cubic[100]
Ba2SmIrO6 Cubic[100]
Ba2DyIrO6 Cubic[100]
Sr2InIrO6 P21/n InsulatorNone1.608[101]
Sr2CoIrO6 I2/m InsulatorAFM0.47[102104]
Sr2FeIrO6 I2/m InsulatorAFM[101]
  $I\overline{1}$ InsulatorAFM[105]
Ca2FeIrO6 $I\overline{1}$ InsulatorAFM[105]
La2LiIrO6 Pmm2None1.42[106, 107]
BaLaMgIrO6 PseudocubicNone[106]
Bi2NaIrO6 P21/n InsulatorNone0.19[108]
SrLaMgIrO6 P21/n InsulatorNone0.61[109]
SrLaZnIrO6 P21/n InsulatorNone0.46[109]
SrLaNiIrO6 P21/n InsulatorAFM0.19[109]
Post perovskite (corner & edge shared)NaIrO3 Cmcm InsulatorNM0[75, 110, 111]
Ba3ZnIr2O9 P63/mmc InsulatorSOL0.2[76]
6H-perovskite (face shared)Ba3MgIr2O9 P63/mmc InsulatorNone0.5–0.6[112, 113]
Ba3CaIr2O9 C2/c InsulatorNone[112, 113]
Ba3SrIr2O9 C2/c InsulatorNone[112, 113]
Ba4LiIr3O12 P63/mmc InsulatorMagnetic[114]
Ba4NaIr3O12 P63/mmc Insulator[114]
Ba3CdIr2O9 C2/c InsulatorNone0.3[115]
Hexagonal (isolated)Sr3NaIrO6 $R\overline{3}c$ InsulatorNone[116, 117]
Sr3LiIrO6 $R\overline{3}c$ InsulatorNone[116118]
Pyrochlore (corner shared)Cd2Ir2O7 FCC[119]
Ca2Ir2O7 FCC[119]
OthersKIrO3 Pn31.04[120]
Ba0.5IrO3 I23Metal[119]
Sr0.5IrO3 BCC[119]
Na3Cd2IrO6 C2/m [121]

The magnetism in pentavalent iridates are, however, not beyond doubt. The situation is, in particular, controversial and flooded with conflicting results for the double peroveskite iridate both in theory and experiments [75, 77, 8793]. Nevertheless, the possibility of the breakdown of the anticipated J = 0 state, leading to the unconventional magnetism, have generated enough interest in pentavalent iridates, which are further accelerated by the observation of novel 'spin–orbital liquid' state in the 6H perovskite iridate [76]. The easy availability of the d4 electronic configuration in iridates, compared to other 5d TM oxides, have further boosted the search for the realization of J = 0 state and/or the excitonic magnetism in iridates.

3. Microscopic model and breakdown of J = 0 state

The various comparative energy scales, e.g., Δ, JH and λ as discussed above, in 5d TM oxides enrich the physics of d4 iridates. The presence of magnetism in different iridates out of apparently non-magnetic individual Ir atoms piqued enormous scientific curiosity in the d4 systems. While the possibility of excitonic magnetism has already been proposed, the understanding of the role of different competing parameters can provide the answer to the two important questions, that are of primary interest, namely, which parameter (or parameters) might drive the unconventional magnetism in d4 spin–orbit coupled systems at the first place. Secondly what would be the nature of this emergent magnetism. Generic theoretical models, both atomic as well as beyond atomic limit, developed in this context, are proven to be useful to gain valuable insight into the physics of the d4 spin–orbit coupled systems. Here we give a brief overview of these results that essentially provide the ground to analyze the magnetic response of the real materials.

3.1. Atomic model and non-cubic crystal field effect

For simplicity, we first consider a many-body atomic model and, then, eventually, we shall include the band structure effects by allowing hopping between the spin–orbit coupled atoms. Assuming the eg orbitals are completely decoupled from the t2g sectors owing to the strong octahedral splitting Δ, only the t2g orbitals are considered at each atomic site. The corresponding atomic many-body Hamiltonian, spanned in the Hilbert space of dimension 6C4 = 15, includes the electron–electron interactions as well as the SOC interaction, viz,

Equation (4)

Here the Kanamori Hamiltonian ${\mathcal{H}}_{\text{int}}$ contains the intra-orbital and the inter-orbital Coulomb interactions U and U', which are related to each other via the Hund's coupling JH, viz, U = U' + 2JH. The schematic illustration of the various interaction terms are given in figure 4. The explicit form of the Hamiltonian at an atomic site i is given by [98, 123125]

and the spin–orbit interaction is

Equation (5)

Here, i-l-σ denotes the site-orbital-spin index and di, (${d}_{i,l\sigma }^{{\dagger}}$) is the annihilation (creation) operator.

Figure 4.

Figure 4. Schematic illustration of the Kanamori interaction terms, intra-orbital Coulomb interaction U (left), Hund's Coupling JH (middle), and inter-orbital Coulomb interaction U' (right). For simplicity two orbitals and two electrons are considered, the different arrangements of which leads to different energies E depending on the various interactions.

Standard image High-resolution image

The exact diagonalization of the atomic model, equation (4), gives the spin–orbit coupled eigenstates that are characterized by the total angular momentum quantum number J: J = 0 (1), J = 1 (3), J = 2 (5), J = 2 (5), and J = 0 (1) with increasing order in energy (see figure 5(a)). Here, the number within the parenthesis denotes the degeneracy of the states. As anticipated, J = 0 singlet forms the ground state (similar to the single particle picture, described in the previous section) which is separated from the excited triplet J = 1 state. The energies of these states are dictated by the interaction parameters such as Hund's coupling JH, Coulomb interaction U and SOC constant λ, while their energy difference is independent of U. The exact energy eigenvalues of the atomic states J = 0 and J = 2 depend on the relative strength of JH and λ (In the limit JHλ, EJ=0 = E0 − 4U + 7JHλ and EJ=2 = E0 − 4U + 7JH + λ/2; for JHλ, EJ=0 = E0 − 4U + 7JH − 2λ and EJ=2 = E0 − 4U + 7JHλ/2) but their energy difference is always 3λ/2. The energy eigen value of the J = 1 state, on the other hand, always have the same analytical form EJ=1 = E0 − 4U + 7JHλ/2 [126].

Figure 5.

Figure 5. Emergence of magnetism in a spin–orbit coupled atomic d4 model. (a) The energy evolution of the many-body states of the atomic model, equation (4), as a function of SOC for the d4 electronic configuration. The ground state of the atomic model is a non-magnetic J = 0 state. (b) The emergence of magnetism in the phase diagram of an atomic model in presence of non-cubic crystal field ΔNCF, showing the variation of the ground state expectation value of J2 operator in the parameter space of SOC λ and ΔNCF for a fixed value of JH/U = 0.1. ΔNCF in (b) represents the energy separation between the subsequent t2g levels, the degeneracy of which is broken completely and are energetically equispaced. The figure is reprinted from reference [98], ©2020, with permission from Elsevier.

Standard image High-resolution image

It is noteworthy that in the simple atomic model in equation (4), we have assumed an ideal cubic crystal field with degenerate t2g states. In most of the real materials, however, the TM-O6 octahedral unit undergoes distortions, tilt, and rotation, removing the degeneracy of the t2g states. Indeed, the resulting non-cubic crystal field is important to destabilize the atomic J = 0 state, leading to magnetism even within an atomic model. This is straight forward to see by adding a non-cubic crystal field term, ${\mathcal{H}}_{\text{NCF}}^{i}={\sum }_{lm\sigma }{\varepsilon }_{lm}{d}_{i,l\sigma }^{{\dagger}}{d}_{i,m\sigma }$ in the Hamiltonian, equation (4), where the onsite energies of the different orbitals are depicted by ɛlm and the difference in the onsite energies is the measure of non-cubic crystal field ΔNCF. The phase diagram, figure 5(b), for the ground state shows the emergence of magnetism (J ≠ 0) in the d4 system due to non-cubic crystal field. The competition between ΔNCF and SOC λ is evident from this phase diagram. While in the strong SOC regimes (λ ≫ΔNCF) the non-magnetic J = 0 state remains intact, magnetism appears in the energy regime ΔNCFλ (see figure 5(b)).

3.2. Two-site model: band structure effect

The atomic model, described above, misses the important ingredient of a solid which is the hopping between different atoms. The simplest model that incorporates the hopping effect is a two-site model where in addition to the atomic interactions, described in equation (4), hopping between two atomic sites are also allowed. The two-site many body model Hamiltonian in a Hilbert space of 12C8 = 495 basis states reads as

Equation (6)

where the first term represents the inter-site hopping between different orbitals, designated by ${t}_{ij}^{l\sigma ,m{\sigma }^{\prime }}$

Equation (7)

and the second and the third terms in equation (6) represent the atomic interactions. While the second atomic term, spin–orbit interaction, is the same as described before, the first atomic term, the interaction term, may be rewritten in terms of the total occupation number operator ${\hat{N}}_{i}$, total spin momentum operator ${\hat{S}}_{i}$, and total orbital angular momentum operator ${\hat{L}}_{i}$ at a site i as [127]

Equation (8)

A diagonal hopping in the orbital and spin space i.e., ${t}_{ij}^{l\sigma ,m{\sigma }^{\prime }}\to t{\delta }_{lm}{\delta }_{\sigma {\sigma }^{\prime }}$ is assumed, where all three t2g orbitals actively participate in the hopping. The exact diagonalization results of the above Hamiltonian, as discussed in reference [127], are more reliable in the Mott limit (U/t ≫ 1). In this regime, a magnetic phase transition occurs from the non-magnetic J = 0 state to a magnetic J = 1 state. The phase diagram in the parameter space of U, λ and t for a chosen value of JH = 0.2U is shown in figure 6(a). Interestingly, the phase transition J = 0 → J = 1 at a critical value of λ/t is associated with abrupt changes in both the expectation values of individual Ji and total J quantum numbers across the transition. The local Li and Si expectation values, however, remain unchanged throughout (see figure 6(b)). In the magnetic state the local moments are produced by hopping (superexchange) driven mixing of the onsite singlet state with the higher energy triplet states. It is important to note, in contrast to standard 3d Mott insulators, here the local moments are not robust and deviate from their atomic values.

Figure 6.

Figure 6. Schematic illustrations of the emergence of magnetism in two-site d4 spin–orbit coupled model system. (a) The schematic phase diagram in the parameter space of t/λ and t/U, showing the transition from the non-magnetic singlet J = 0 state to magnetic J = 1 and J = 2 states with increasing hopping. (b) The variation of the expectation values of the local ${L}_{i}^{2}$, ${S}_{i}^{2}$ and ${J}_{i}^{2}$ operators and the total J2 operator as functions of λ/t, indicating the discontinuous jump of the total angular momenta ${J}_{i}^{2}$ at each site and J2 for the two-site system across the phase transition, while the local spin and orbital angular momenta ${L}_{i}^{2}$ and ${S}_{i}^{2}$ do not change. (c) The FM to AFM transition with increasing SOC for three sets of (t, JH) as indicated in the figure. The different superexchange paths for the (d) FM and the (e) AFM interactions. Figures (a), (b), (d), and (e) are reprinted with permission from reference [127] ©2015 by the American Physical Society and (c) is reprinted from reference [98], ©2020, with permission from Elsevier.

Standard image High-resolution image

The magnetic transition, discussed above for orbitally symmetric hopping, assumes full rotation symmetry which is usually absent in a real material. For example, in simple cubic systems where the hopping t between TM atoms occur from the superexchange via the oxygen atoms, only two orbitals participate in the hopping (see figure 3), so that ${t}_{ij}^{l\sigma ,m{\sigma }^{\prime }}\to t{\delta }_{lm}(1-{\delta }_{km}){\delta }_{\sigma {\sigma }^{\prime }}$. Here the direction of the line, connecting the two sites i and j, determines the blocked orbital k. On the other hand, for the hopping between nearest neighbor atoms on a face-centred cubic lattice may be approximated by only one active orbital k, viz, ${t}_{ij}^{l\sigma ,m{\sigma }^{\prime }}\to t{\delta }_{lm}{\delta }_{km}{\delta }_{\sigma {\sigma }^{\prime }}$. Here, the active orbital is determined by the common plane, shared by the two sites. Interestingly, the magnetic phase transition occurs for both these cases, emphasizing the crucial role of hopping in generating the magnetism in d4 spin–orbit coupled systems [128]. A next question that naturally occurs is the nature of the magnetism. Meetei et al [127] showed that the generated local moments favor ferromagnetism and a phase transition from the AFM superexchange to the FM superexchange takes place at a critical value of JH/U. An intuitive picture for the Hund's coupling driven ferromagnetism, may be followed from reference [127], based on the simple perturbative analysis. The atomic ground state in the d4-d4 configuration with Li = 1 and Si = 1 has a total energy, E0 = 12U − 26JH, computed using equation (8). The hopping in equation (7) acts as a perturbation and the ground state gains energy via virtual hopping. For FM superexchange, the intermediate d3d5 state (see figure 6(d)) with S1 = 3/2, L1 = 0, S2 = 1/2, and L2 = 1 has the total energy EFM = 13U − 29JH. This results in an energy gain by the FM state ${\Delta}{E}_{\text{FM}}=-\frac{2{t}^{2}}{U-3{J}_{\mathrm{H}}}$. Similarly, the intermediate paths for the AFM superexchange, as schematically illustrated in figure 6(e), gains an energy ${\Delta}{E}_{\text{AFM}}=-\frac{2}{3}\frac{{t}^{2}}{U-3{J}_{\mathrm{H}}}-\frac{14}{6}\frac{{t}^{2}}{U}-\frac{{t}^{2}}{U+2{J}_{\mathrm{H}}}$. Comparing the two energy scales ΔEFM and ΔEAFM, it is easy to argue that the superexchange gradually changes from the AFM to FM state with increasing value of JH/U. It is to be noted that in the above analysis, SOC is ignored and inclusion of SOC tends to compete against the ferromagnetism.

An independent exact diagonalization study [98] with the similar two-site model including also the effect of non-cubic crystal field, also suggests the competition between ferromagnetism and AFM with SOC in d4 systems. While ferromagnetism is stabilized in absence of SOC (λ = 0), with increase in λ ferromagnetic interaction become weaker, and eventually at a critical value λc, an AFM interaction is favored. The critical value of λc depends on the strength of the other parameters of the Hamiltonian as shown in figure 6(c). In particular, a larger Hund's coupling tends to increase the λc, thereby favoring the ferromagnetism. Dynamical mean-field theory (DMFT) study combined with the continuous-time quantum Monte Carlo [129] also supports these findings.

Inspired by the results of the two site exact diagonalization, the lattice problem of magnetism in d4 systems has also been analysed by constructing an effective spin–orbital lattice model [127, 128, 130].

The effective model was employed by Svoboda et al [128] for the analysis of magnetism for the three-different cases of active t2g orbitals, mentioned before, viz, three active orbitals (Norb = 3), two active orbitals (Norb = 2) and single active orbital (Norb = 1). While the detail discussion of the effective magnetic models, developed from the full Hamiltonian in equation (6), can be found in reference [128] for each of the three mentioned cases, here we focus on the key finding of this analysis. According to this study, while ferromagnetism is stabilized at large JH/U limit for Norb = 2 and 3 cases, for the single active orbital case, ferromagnetic interaction is not supported for any reasonable value of JH/U. The authors argue that the absence of ferromagnetism in Norb = 1 is due to the fewer available exchange paths. As illustrated in figure 6, the number of superexchange paths supporting a FM state is less than the corresponding number of such paths available for the AFM state. However, there are more possible active paths in Norb = 2 and 3 to reduce the energy of the FM state so that Hund's coupling takes over, favoring ferromagnetism. This is, however, not the case for Norb = 1, where the energy of the FM state is lowered by only a single factor of −t2/U, failing to favor ferromagnetism.

In addition, reference [128] also predicts an intriguing orbital frustration for the cases of Norb = 1 and 2 with broken rotational symmetry in contrast to the orbital symmetric Norb = 3 case. The orbital frustration is demonstrated on a non-frustrated crystal geometry, as illustrated in figure 7, considering the case of Norb = 2. In a d4 configuration, each site has at least one doubly occupied t2g orbital. For an AFM interaction between two sites, each with two active orbitals (Norb = 2), the doubly occupied orbitals on both sites remain inactive (see figure 6(e)). As a result, inter-site AFM interaction in the Norb = 2 model favors the orbital, perpendicular to the line joining these two sites, to be doubly occupied, e.g., dxz orbital in figure 7(a) is doubly occupied for the bond 1, along y-direction. These doubly occupied orbitals are, however, not inactive for bond 2, say along z-direction, simply because dxz orbital is not perpendicular to this second bond along z. As a result, AFM interaction is no longer favorable and, therefore, the next energetically favorable FM interaction occurs along bond 2. This FM interaction, in turn, leads to a different doubly occupied orbital (dxy orbital in figure 7) for the bond 3 along y-direction, which neither favors energetically lowest FM interaction nor the energetically lowest AFM interaction, stabilizing rather an energetically higher AFM interaction along bond 3. Similarly, starting from the FM interaction between two sites along a particular bond (see bond 1 in figure 7(b)), it is possible to show that a higher energy configuration is taken by one of the four bonds on a plaquette. This leads to a frustration due to orbital degrees of freedom even on a non-frustrated lattice, which may result in an orbital liquid state in the regime λzJSE (z being the coordination number) with strong orbital effects in absence of strong octahedral distortions.

Figure 7.

Figure 7. Graphical illustration of the orbital frustration for Norb = 2 model. The orbital frustration may be described by fixing the type of interaction along bond 1 to either (a) AFM or (b) FM. Once the interaction along bond 1 is fixed, it can be seen from (a) and (b) that the interaction along bond 3 corresponds to a higher energy configuration, leading to orbital frustration. The double occupied orbitals at each site of the square lattice are indicated in the figure. The schematic illustrations in (a) and (b) are reprinted with permission from reference [128] ©2017 by the American Physical Society.

Standard image High-resolution image

3.3. Metal–insulator transition

We now turn to the transport properties of d4 spin–orbit coupled systems. The transport properties in d4 systems, as studied in reference [131] using the rotational invariant Gutzwiller approximation method as well as the DMFT. The calculations reveal metal–insulator transition as a function of SOC constant λ and Coulomb interaction U. A Hubbard model for the t2g orbitals with full Hund's rule coupling and spin–orbit interaction, similar to the many-body model, discussed above, is considered in this work. The computed phase-diagram in the parameter space of λ and U (see figure 8) depicts the three different states, viz, metallic state, Mott insulating state, and band insulating state. The two insulating states can be visualized by considering the two limiting cases. First, let us consider the strong SOC limit (U = 0). In this limit, the degenerate t2g bands split into jeff = 3/2 and jeff = 1/2 bands, separated from each other with an energy gap of ≈3λ/2. With increase in λ, the higher lying jeff = 1/2 bands tend to get empty as the electrons move to the jeff = 3/2 bands. Consequently, at a critical value of λ/D, D being the half-bandwidth, the jeff = 3/2 bands are completely occupied, resulting in a transition to the band insulating state from a metallic state. Moving to the other limit where λ = 0, the degenerate t2g orbitals have four electrons at each site. With increasing U, the system undergoes an interaction driven Mott transition at some critical value of U/D with each band having 4/3 electrons.

Figure 8.

Figure 8. Schematic illustration of the phase diagram, adapted with permission from reference [131] ©2013 Springer Nature, showing the metal to insulator transition in the parameter space of U/D and λ/D, where D is the half-bandwidth.

Standard image High-resolution image

Coming back to the discussion of the phase diagram in figure 8, we can see that the phase diagram may be subdivided into three different regions: (I) the regime 0 < λ/D < 0.24, where increase in U leads to a transition from metal to Mott insulator. The decrease in the critical value of U for this metal–insulator transition with increasing SOC, suggests the enhancement of the correlation effects by the SOC. Intuitively, this may be attributed to the narrow bands, resulting from the SOC driven band splitting. It is interesting to point out that similar cooperation between SOC and correlation has been observed for d5 filling as well [23]. In this Mott phase, the local moment at each site vanishes, forming a spin–orbital singlet (SOS) state and the ground state is simply the product state of the local singlets at each site. (II) The next regime is 0.24 < λ/D < 1.33. Two back to back transitions occur in this regime as we increase U. The first transition is from metal to band insulator, followed by the second transition from band to Mott insulator. For the intermediate U, the effective bandwidth of the system is reduced by the correlation effect first, which upon relatively small band splitting due to intermediate λ results in a band insulating state. Increasing U further pushes the system into Mott regime. (III) Finally, in the regime λ/D > 1.33, strong SOC leads to completely empty jeff = 1/2 bands even at U = 0, suggesting a band insulating state in the non-interacting case. Increasing U to a large value induces again a transition from band insulating state to Mott insulating state.

4. Materials

The various theoretical predictions [71, 127] described in the previous section, for the d4 spin–orbit coupled systems generated enormous interest in pentavalent iridates for the realization of excitonic magnetism. The search is further, triggered by the observation of novel magnetism in d4 double perovskite system Sr2YIrO6 [74], followed by a series of other d4 iridates. The observed magnetism in iridates is, however, quite enigmatic. The ground state properties of these materials are full of controversial and conflicting results. While a large number of materials have been studied and characterized in past years, here we focus on a few particular materials that have received considerable attention in the d4 iridate research.

The pentavalent iridates, that we discuss in this review, can be broadly categorized into four classes according to their crystal geometry: (a) double perovskite (b) post perovskite (c) 6H perovskite and (d) hexagonal iridates. The crystal structure plays a key role in deciding magnetism in iridates, so the ground state properties of the iridates, belonging to the same geometry class, are expected to be quite similar. Interestingly, the ground state of d4 iridates is so delicate that even small changes in the fine details of the structure can lead to dramatic changes in the ground state properties, as we now proceed to discuss below.

4.1. Double perovskite iridates

The general formula for the double perovskite iridates, that we discuss here, is A2 BIrO6, where the A site is occupied by the divalent cations, e.g., Sr, Ba. In this case, the trivalent ions such as Y3+, Sc3+, Lu3+, La3+, etc, at the B site renders the desired 5+ charge state at the Ir site. The basic structure of the double perovskite iridates consists of corner shared BO6 and IrO6 octahedra that alternate along each of the crystallographic axes. These octahedra are either ideal or undergo specific types of distortions, depending on the particular crystal symmetry of the structure. In general, the crystal structure accompanies strong geometric frustration due to face-centered cubic lattice formed by the neighboring Ir atoms (see figure 9).

Figure 9.

Figure 9. Schematic illustration of the two different scenarios, proposed for double perovskite iridate Ba2YIrO6. (a) Crystal structure of a double perovskite iridate (top) and the geometric frustration of the face-centered lattice formed by the Ir ions (bottom). (b) The summary of the two conflicting scenarios, discussed in theory and experiment, regarding the magnetism in Ba2YIrO6. While the scenario (I) supports the breakdown of the J = 0 state, scenario (II) suggests the presence of J = 0 state. According to (I), the experimentally observed moment is intrinsic and it appears due to band structure effect, illustrated in (c), where the double perovskite iridate lies in the intermediate regime of λW. The scenario (II), on the contrary, suggests absence of any moment at the Ir site and the observed moment in the experiment is ascribed to the extrinsic effects, such as anti-site disorder, impurity, magnetic defect etc. The figure (d) in this scenario shows the weak dispersion of the triplet excitation spectrum, failing to overcome singlet–triplet energy gap. The figures for the first scenario is taken from reference [75] and the second scenario is reprinted with permission from reference [90] ©2017 by the American Physical Society.

Standard image High-resolution image

Owing to the compositional and structural flexibility, double perovskite iridates have particularly received enormous attention. The interest in the d4 double perovskite iridates began with the observation of the onset of magnetism in double perovskite iridate Sr2YIrO6 in the year 2014, although, the magnetism in iridates with d4 electronic configuration was reported even earlier, e.g., in another double perovskite system Sr2CoIrO6 [102] (this is even before the theoretical prediction by Khaliullin [71]). In this work Narayanan et al [102] studied the structural, electronic and magnetic properties of La2−x Srx CoIrO6 for 0 ⩽ x ⩽ 2, the end member of which is Sr2CoIrO6 with high spin Co3+ and low spin Ir5+ ions. Sr2CoIrO6 undergoes successive structural transitions starting from a two-phase region upon heating, viz, $P{2}_{1}/n+I2/m\to I2/m\to I4/m\to Fm\overline{3}m$, and exhibits AFM ordering as revealed from magnetization measurements. Interestingly, the estimated effective magnetic moment for the system μeff = 5.1(1)μB /f.u. suggests for magnetic contribution from the Ir5+ ion, as the contribution (μeff = 4.9μB /Co) for the Co3+ ion alone cannot describe the estimated moment value. This small moment, however, remains undetected in the neutron powder diffraction measurements. Nevertheless, the density functional calculations, presented in the same work, also suggest significant spin and orbital moments at the Ir site. However, the material is reported to have significant anti-site disorder and, also, the presence of strongly magnetic Co3+ ions brings up the question if the Co ions have any role in the magnetism at the Ir site. Later, it was also pointed out that in addition to Ir5+ ion, the system also has Ir6+ ions, which have rather strong magnetic signal [104]. In view of the above, the double perovskite iridates where the magnetism can appear only via Ir (B site is non-magnetic) are the best candidates for the investigation of magnetism in the Ir5+ ion.

A series of such double perovskite iridates A2 R3+Ir5+O6 (A = Sr, Ba and R = La, Y, Sc, Lu) were synthesized long ago in 1999 by Wakeshima et al [95] and all of these compounds were reported to have paramagnetic behavior down to 4.5 K. With the new insight into the SOC effect in iridates, later, Cao et al [74] revisited the compound Sr2YIrO6 and showed that the system undergoes a novel magnetic transition below T = 1.3 K. The observed magnetism was surprising as both the presence of Ir5+ ion as well as strong geometric frustration in the double perovskite structure in general should forbid any long range magnetic order. The reported moment μeff at the Ir site is 0.91μB and the Curie–Weiss temperature θCW is as large as −229 K, indicating the presence of AFM interaction as well as strong frustration in the system. Sr2YIrO6 crystallizes in monoclinic P21/n symmetry, where the IrO6 octahedron undergo distortion. The authors argued the non-cubic crystal field, resulting from this distortion, as the possible origin of deviation from the singlet J = 0 state, leading to magnetism in Sr2YIrO6.

Interestingly, the parallel study of another d4 double perovskite iridate Sr2GdIrO6 in the same work [74] showed that unlike the Y-compound, the Gd-compound orders ferro-magnetically at Tc = 2.3 K with a comparable Curie Weiss temperature θCW = 1.2 K. The saturation moment was found to be Ms ∼ 7μB /fu consistent with the Gd3+ (4f7) charge state, present in the system. The authors proposed Ir atoms in Sr2GdIrO6 to be non-magnetic, however the presence of a strong magnetic signal from Gd3+ ion makes the confirmation difficult. According to this study, the latter iridate crystallizes in cubic space group Pn-3 and, therefore, is free from any non-cubic distortion. In absence of any non-cubic crystal field, SOC dominates, resulting in a J = 0 state at the Ir site in Sr2GdIrO6. Both double perovskite iridates were reported to be insulating with an energy gap of 490 meV and 272 meV respectively.

The magnetic properties of these two double-perovskite iridates were later studied within the framework of density functional theory [75]. The electronic structure calculations presented in this work reveal that in addition to non-cubic crystal field the band structure effects are, more crucial for the breakdown of the anticipated J = 0 state leading to the formation of moments. It is interesting to point out that the crucial role of the band structure effects has also been discussed in the context of tetravalent iridates [132]. The conclusions of reference [75] is based on the estimated non-cubic crystal field and the magnetic moments at the Ir site in the two double perovskite iridates, mentioned above, along with another double perovskite, Ba analogue of Sr2YIrO6, Ba2YIrO6. The authors showed that even though Sr2GdIrO6 crystallizes in the cubic structure, the Pn-3 symmetry allows for the rotation of the IrO6 octahedra, that leads to a strong trigonal splitting of the Ir-t2g orbitals, which is about an order of magnitude larger than that in Sr2YIrO6. Indeed, the electronic structure calculation showed the presence of finite moment at the Ir site in both these systems. The situation became even more interesting with the cubic double perovskite iridate Ba2YIrO6, where the $Fm\overline{3}m$ symmetry does not allow any tilt or rotation of the IrO6 octahedra, resulting in a triply degenerate t2g orbitals. Even in this case, finite moment at the Ir site persists. This emphasizes the band structure effects to be crucial in driving magnetism in these series of double perovskite iridates (see scenario I in figure 9(b)). Interestingly, despite the different non-cubic crystal field, the bandwidths in this series of double perovskites are comparable and large in magnitude (∼1.2 eV) that leads to significant mixing between jeff = 1/2 and jeff = 3/2 states (see inset of figure 10(b)), resulting in deviation from the ideal atomic limit. The details of the evolution of the electronic structure of Ba2YIrO6, obtained in this work, is summarized in figure 10.

Figure 10.

Figure 10. The densities of states (DOS) for Ba2YIrO6, computed within (a) GGA, (b) GGA + SOC, and (c) GGA + SOC + U, considering ferromagnetic spin configuration. The dominant orbital characters are indicated in the figure. The GGA spin polarized calculation gives a half-metallic solution in (a). Inclusion of SOC leads to changes in the DOS, as indicated in (b). The inset in (b) shows the mixing between jeff = 1/2 (shaded region) and jeff = 3/2 (solid line) states, obtained from a non-magnetic Ir-t2g only tight-binding model in presence of SOC, indicating the origin of finite moment at the Ir-site. Application of Coulomb interaction U fails to open up a gap within the ferromagnetic spin configuration, while the energy gap is present in the computed energetically lowest type-I anti-ferromagnetic configuration as shown in the inset of (c).

Standard image High-resolution image

In complete contrast to references [74, 75], DMFT calculations [88] suggest absence of any long range magnetic order in both the compounds Sr2YIrO6 and Ba2YIrO6. In the above work, taking density functional theory as a starting point, within the constrained random phase approximation, a multi band Hubbard model is constructed which is further analyzed within static and dynamic mean field theories and strong coupling expansion. The authors showed that a large U (U = 4 eV) is required to get an insulating solution, reducing significantly the mixing between jeff = 1/2 and jeff = 3/2 states that was present within GGA + SOC. Using effective strong coupling model, the authors concluded that the reduced magnitude of hopping due to large Ir–Ir distance in the double perovskite structure is too small to overcome the singlet–triplet energy gap, which was predicted to be in the range of 200 meV. This results in a non-magnetic J = 0 singlet ground state in both the iridates.

Around the same time, Dey et al [87] revisited and studied in detail the structural, magnetic, and thermodynamic properties of Ba2YIrO6, which was originally synthesized more than 45 years ago [100]. This study confirmed the ideal cubic structure of Ba2YIrO6, in contrast to the monoclinic symmetry reported earlier [95], and showed that the system does not order down to 0.4 K. However, the Curie–Weiss fit yields a Curie Weiss temperature θCW ∼ −8.9 K and an effective moment of 0.44μB /Ir, unexpected for a J = 0 state. The authors exclude the possibility of any anti-site disorder due to different sizes of Ir5+ and Y3+ ions or oxygen vacancy from their magnetization measurements on the oxygen-annealed crystals. On the contrary, the DFT calculations presented in the same work show a non-magnetic ground state in Ba2YIrO6 with significant mixing between j3/2 and j5/2 states (the effective jeff = 1/2 state is part of this full j5/2 manifold [58]). While in absence of Hubbard U, the system remains metallic, the insulating state is obtained with a rather small value of U, viz, U = 1.4 eV and JH = 0.5 eV in contrast to reference [88]. The origin of the moment at the Ir site remains elusive from this work.

In the meantime, independent experiments [133, 134] probed the effect of structural distortions on the magnetism of the series of double perovskites Ba2−x Srx YIrO6 (0 ⩽ x ⩽ 1). Both studies show no evidence for magnetism down to 2 K. The effective moment, reported in reference [133], is rather small, 0.16μB /Ir and was speculated as a result of local disorder in the structure which was not captured in the diffraction measurements. These results were, further, disputed by Terzic et al [93], emphasizing that the realization of the magnetic ground state requires investigation below 2 K. They confirmed the presence of magnetic order in the single-crystal, double-perovskite Ba2YIrO6 and Sr-doped Ba2YIrO6 below 1.7 K using dc magnetization, ac magnetic susceptibility, as well as heat-capacity measurements. From a fitting over a wide temperature range 50 < T < 300 K in this work, the authors estimated an effective moment of 1.44μB /Ir in Ba2YIrO6 and a large Curie–Weiss temperature of −149 K. The authors questioned the large singlet–triplet gap predicted earlier [88], claiming that such a large gap does not explain the significant temperature dependence of the magnetic susceptibility at low temperature.

The debate continued as, in the same year, Corredor et al [94] even questioned the previously reported monoclinic symmetry of Sr2YIrO6 [74, 133] and proposed instead the same cubic structure as its Ba analogue. According to this work Sr2YIrO6 does not undergo any magnetic transition down to 430 mK. They pointed out the low-temperature anomaly in the specific heat as a Schottky anomaly due to paramagnetic impurities, and is, therefore, not related to the onset of magnetic order. The small effective magnetic moment μeff = 0.21μB /Ir and the small correlation were attributed to the J = 1/2 impurities in the sample. Along the same line, the observed moment in Ba2YIrO6 was ascribed to extrinsic origin, such as paramagnetic impurity [89], anti-site disorder [90], Ir4+/Ir6+ magnetic defects [91], excluding the possibility of any excitonic magnetism (see scenario II in figure 9(b)).

On the contrary, Nag et al [77] pointed out that the temperature dependence of the magnetic response below 10 K for Ba2YIrO6 as revealed from their μSR studies is rather unusual. In particular, the temperature dependence of the stretched exponent parameter β of the fitted muon polarization P(t) = $\mathrm{exp}-{\left(\alpha t\right)}^{\beta }$ cannot be corroborated with a J = 0 state, for which no such temperature dependence of β should exist. The observed behavior is explained as follows: while a small iridium moment exists ubiquitously in Ba2YIrO6, at higher temperature the system behaves as a paramagnet. When the temperature is lowered below 10 K, most of these Ir moments start to interact with each other anti-ferromagnetically and pair up to form SOSs, leaving behind a few Ir ions due to Y/Ir disorder and with a very low residual interaction between them. This is in sharp contrast with the previous proposal where the impurity in the J = 0 background has been suggested for Ba2YIrO6. The proposed scenario of Nag et al is quite similar to its 4d analogue Ba2YMoO6 [135], supporting the scenario I in figure 9(b).

Clearly the ground states of Sr2YIrO6 and Ba2YIrO6 are quite controversial. The situation becomes even more puzzling with the report of resonant inelastic x-ray scattering (RIXS) measurements in these double perovskite systems [77, 136] that facilitate a quantitative estimation of the parameters λ and JH in these systems. We shall discuss the RIXS technique in some detail later in the context of 6H perovskite iridates. Interestingly, regardless of the presence of the non-cubic crystal field (Sr vs Ba compound) or its strength (Y vs Gd compound), the RIXS spectra in all these double perovskite iridates are quite similar and can be described by an atomic model with λ ≈ 0.39–0.42 and JH ≈ 0.24–0.26 eV [77, 136]. While theoretical studies [90, 92, 137] suggest excitonic magnetism is unlikely in Ba2YIrO6 due to large estimated singlet–triplet gap in RIXS measurement, reference [93] indicates the comparable energy scales of Hund's coupling and SOC, as a possible factor to drive magnetism in these double perovskite iridates. Furthermore, recent DFT calculations [138] show the presence of type-I AFM state in Ba2YIrO6, similar to the initial DFT results [75], confirming the possibility of interesting physics beyond the atomic limit. The insulating state in reference [138], however, occurs at a higher U than in reference [75] (U = 4 vs U = 2 eV). The magnetism in reference [138] is attributed to the 'intracluster excitons' that arise from the Ir–O interaction within an IrO6 cluster.

The formation of moments at the iridium site was also discussed in the Sc counterparts of Sr2YIrO6 and Ba2YIrO6, viz, Sr2ScIrO6 and Ba2ScIrO6 [9698]. Similar to the Y compounds, Ba–Sc compound crystallizes in the cubic $Fm\overline{3}m$ structure while Sr–Sc compound undergoes monoclinic distortion of the P21/n space group. Experimentally, the systems do not exhibit long range magnetic order while an effective moment of 0.16μB /Ir for the Sr compound and even larger moment of 0.39μB /Ir for the Ba compound was estimated. The generated magnetic moment was attributed to the hopping induced de-localization of holes in these systems [97, 98].

The compositional flexibility within the general formula of A2 BIrO6 leads to a large number of pentavalent double perovskite iridates, that includes but not limited to the listed materials in table 1. Further flexibility is achieved by doping at the A-site or/and using a magnetic atom at the B site. While doping offers the possibility to explore the effect of the resulting structural distortion on the magnetic properties of the system, magnetic B-site offers the direct study of 3d–5d interaction, combining the strong correlation of 3d TM ion to the strong SOC of 5d TM ion. While this offers a broad area of research spanned by some excellent works [101, 103105], we do not discuss them in the present review.

4.2. Post-perovskite iridate

The first post-perovskite oxide with pentavalent atom NaIrO3, containing Ir in the d4 configuration, was synthesized using high-pressure solid state method [110]. The post perovskite NaIrO3 crystallizes in the layered orthorhombic Cmcm space group. In contrast to the corner shared IrO6 octahedra in the double perovskite iridates, discussed above, here the structure consists of distorted IrO6 octahedra which not only share their corners along the c-axis but also share the edges along the a-axis and, thereby, form IrO3 sheets. As shown in figures 11(a) and (b), these sheets are, further, stacked along the b direction, separated by the Na atoms.

Figure 11.

Figure 11. The S = 0, insulating state of NaIrO3. (a) The crystal structure of NaIrO3 in the orthorhombic Cmcm space group. (b) The distorted IrO6 octahedra and their connection in the ac plane. Different bond lengths and bond angles are also indicated in the figure. (c) The resulting non-degenerate Ir-t2g states, t1, t2, and t3, and the non-cubic crystal field splitting between the states. The large ΔNCF leads to S = 0 state, rather than J = 0 state. (d) The schematic illustration of the phase diagram, adapted with permission from reference [111], in the parameter space of U and λ relevant for NaIrO3. The realistic parameters for NaIrO3 (U = 2, λ = 0.33 eV, and JH = U/4) correspond to a band insulating state, indicated by the black triangle in the figure. (e) The computed density of states for NaIrO3, adopted from reference [75], considering the experimentally reported crystal structure [110] (bottom) and the optimized structure as in Table. 2 (top), indicating correlation effects may not be important for the experimentally observed insulating behavior.

Standard image High-resolution image

Interestingly, the magnetic measurement as well as DFT calculations [110] confirm the nonmagnetic state of Ir in NaIrO3. This was, thus, considered an ideal example for the realization of the J = 0 nonmagnetic state [110], in marked contrast with the series of compounds, discussed in the previous section. The electrical resistivity in this work showed that the system is non-metallic, however the DFT calculation fails to describe the insulating state observed in experiments, leaving the origin of the resistivity unclear. A variable range hopping model was used to describe the low temperature resistivity of NaIrO3. This put forward two important questions for this post-perovskite iridate: firstly, whether NaIrO3 is a true realization of the J = 0 ground state? Second, whether the material is closer to a 'renormalized band insulator', where the correlation reduces the effective bandwidth which upon SOC becomes gapped out, or is it closer to 'Mott insulator', driven by strong Coulomb interaction. The DFT calculations in references [75, 111] addressed these issues, providing insights to the answer to these questions, which we now proceed to discuss.

4.2.1.  S = 0 instead of J = 0 state in NaIrO3

According to the theory that supports the magnetism at the Ir site in the pentavalent double perovskite iridates, as discussed in section 4.1, emphasizes the crucial role of band structure effects in preventing the realization of the desired J = 0 state. The band structure effects become particularly important due to the three-dimensional connectivity of the Ir atoms. Therefore, the reduced dimensionality of the connecting Ir atoms, resulting from the two-dimensional Ir sheets in the post perovskite structure (see figure 11) raises the hope to find the J = 0 state in NaIrO3.

However, in addition to the desired layered structure, the IrO6 octahedra in NaIrO3 also suffer from strong distortions (see figure 11(b)). The ∠O–Ir–O bond angle deviates significantly from the ideal 90° angle to about ≈73.2°, while the edge and corner sharing ∠Ir–O–Ir angles are respectively 106.8° and 140.4°. The Ir–O bond lengths also differ by about 0.02 Å. Such a high distortion leads to strong non-cubic crystal field, that removes the degeneracy of the t2g orbitals completely, splitting it into three non-degenerate energy levels. The energy separation between these levels were estimated to be ΔNCF ≈ 0.57–0.59 eV [75], which is comparable to the typical energy scale of the Hund's coupling JH ≈ 0.4–0.6 eV in 5d TM oxides. As a result, ΔNCF dominates over JH, stabilizing an electronic configuration (↑↓, ↑↓, 0) (see figure 11(c)) that leads to a non-magnetic state.

Clearly, the non-magnetic state in NaIrO3 is distortion driven and has nothing to do with the SOC. Thus, it is better described by an S = 0 state, rather than a J = 0 state. The stabilization of the non-magnetic state even without SOC in the DFT calculations [75, 110] further supports this finding.

4.2.2. Insulating state

In contrast to the observed insulating behavior in the experiments, DFT calculations show a metallic state in the non-magnetic ground state [75, 110]. Application of SOC and Coulomb correlation U do not affect this scenario. This leads to the puzzling insulating behavior in NaIrO3.

Failing to describe the insulating behavior within DFT + SOC + U approach, reference [111] made an attempt to investigate the insulating state with the LDA + Gutzwiller method. Starting from a low energy tight-binding model, derived from the LDA calculation, with atomic SOC in it, they succeeded in obtaining an insulating solution for NaIrO3 for reasonable values of SOC (λ ≈ 0.33 eV) and U = 2 eV, relevant for NaIrO3. Furthermore, they identified the insulating state of NaIrO3 as a 'band insulator' rather than a Mott insulator (see figure 11(d)).

Reference [75] further emphasized the influence of the structural geometry in driving the insulating state in NaIrO3. Structural optimization, keeping the experimental symmetry and the lattice parameters intact, showed that the position of the light oxygen atoms were changed significantly (see table 2), leading to an almost insulating state, as depicted in figure 11(e). This further supports earlier finding of band insulating state in NaIrO3, indicating correlation effects may not be significant in NaIrO3. With these new theoretical insights, it might be worthwhile to revisit the transport properties of NaIrO3, possibly on single crystal samples, motivating future experiments in this direction.

Table 2. Comparison of the Wyckoff positions for the experimental and the optimized crystal structure of the primitive unit cell of NaIrO3, as reported in reference [75].

AtomsExperimental structureOptimized structure
x y z x y z
Na0.00.25070.250.00.25130.25
Ir0.00.00.00.00.00.0
O1 0.50.437 660.250.50.426 210.25
O2 0.50.101 590.056 810.50.116 520.058 52

4.3. 6H perovskite iridates

We now discuss the d4 6H-perovskite iridates with the general formula Ba3 MIr2O9. The divalent cation M, e.g., Ca, Mg, Zn, Sr renders 5+ charge state to Ir. These 6-H perovskite iridates were synthesized by Sakamoto et al [113] back in 2006 and, then, further revisited with an idea to understand the SOC effects [76, 78, 112, 115]. It is interesting to point out here that similar to the double perovskite structure, 6H pervoskite structure also offers the composition flexibility, viz, the varying charge state of the M cation results in different valency of the Ir ion, which includes the rather unusual fractional charge states in addition to the common 4+ charge state of Ir atom [100, 113, 139142]. For example, Li ion at the M site or trivalent M ions such as, Y, Sc, In, give rise to fractional charge states 5.5+(d3.5), and 4.5+(d4.5) at the Ir site respectively, while tetravalent M ions, e.g., Ti and Zr result in a 4+ charge state of Ir ion (d5). Here we address the two key issues, viz, the impact of the local distortion, assisted by the M cation, on the magnetic properties of the 6H perovskite iridates and, also, discuss the hopping induced formation of the magnetic moment at the Ir site.

4.3.1. Structural distortion and its impact on the magnetic properties of Ba3 MIr2O9

The basic crystal structure of Ba3 MIr2O9 is quite intriguing. In contrast to the previously discussed structures, here the IrO6 octahedra share their faces with each other forming thereby Ir2O9 bi-octahedra, which in turn leads to the formation of structural Ir–Ir dimer. Such dimers are again arranged on a geometrically frustrated triangular network, as illustrated in figure 12. Furthermore, the size of the cation M, has a significant impact on the crystal geometry, which, in turn, affect the magnetic properties in different members within the same 6H perovskite family. For example, the Rietveld analyses of the x-ray diffraction data [76, 112, 113] show that while Ba3ZnIr2O9 and Ba3MgIr2O9 with the smaller cations Zn and Mg crystallize in the hexagonal P63/mmc space group, having less distorted IrO6 octahedra, rest of the compounds, where the cation M is larger in size, adopt the crystal structure with monoclinically distorted C2/c space group, having significantly distorted IrO6 octahedra. While in the Zn and Mg compounds, the trigonal distortion of the face-shared IrO6 octahedra in the hexagonal symmetry splits the Ir-t2g levels into a low-lying doublet (${e}_{g}^{\pi }$) and a singlet (a1g ) with a non-cubic crystal field splitting ΔNCF, the monoclinic symmetry of the Ca and Sr compounds leads to complete removal of the degeneracy of the t2g orbitals, denoted by t1, t2, and t3 in figure 12(d), with the corresponding energy differences Δ1 and Δ2 respectively (see figure 12(d) for pictorial representation). The stronger distortion in the latter compounds also leads to stronger non-cubic crystal field splitting in these systems. The comparison of the crystal geometry and the resulting non-cubic crystal field splitting (ΔNCF or Δ1, Δ2) are listed in table 3.

Figure 12.

Figure 12. Structural distortion and magnetic properties of Ba3 MIr2O9. (a) Crystal structure of Ba3 MIr2O9. The structure consists of face-shared Ir2O9 units, which are arranged on a triangular network, shown in (b). The intra- (dintra) and inter-dimer (dinter,ab and dinter,c ) distances are indicated in (a) and (b). (c) The Ir2O9 unit, that undergoes different types of distortion for M = Zn, Mg, Ca, and Sr (see table 3). This results in a trigonal crystal field splitting ΔNCF in Zn and Mg compounds (left), while the t2g degeneracy is completely removed for Sr and Ca compounds (right), shown schematically in (d). The non-degenerate t2g states for Sr and Ca are denoted by t1, t2, and t3 and the energy separations between the consecutive levels are denoted by Δ1 and Δ2 respectively. The cartoons, adopted from reference [76], showing (e) the FM and (f) the AFM interactions between Ir5+ ions in absence and presence of SOC (see text for details). (g) and (h) show the symmetry allowed magnetic structures corresponding to the magnetic space groups C2'/c' and C2/c' which are found to be energetically lowest magnetic configurations for Sr and Ca compounds respectively. Figures (g) and (h) are adopted from reference [112].

Standard image High-resolution image

Table 3. The comparison of the structural parameters and the resulting non-cubic crystal field splitting in Ba3 MIr2O9, M = Zn, Mg, Ca and Sr.

 Ba3ZnIr2O9 Ba3MgIr2O9 Ba3CaIr2O9 Ba3SrIr2O9
a (Å)5.770 345.77055.90915.9979
b (Å)5.770 345.770510.237010.3278
c (Å) 14.344114.321614.762415.1621
β (°)  91.1492.78
Ir–O1 (Å)2.0212.0582.0102.041
Ir–O2 (Å)1.9481.9322.0792.026, 2.181
Ir–O3 (Å)  1.8491.991
Ir–O4 (Å)  1.8651.839
Ir–O5 (Å)  2.0922.079
${d}_{\text{Ir}}^{\text{intra}}$ (Å)2.752.762.7352.69
${d}_{\text{Ir}}^{\mathrm{i}\mathrm{n}\mathrm{t}\mathrm{e}\mathrm{r},c}$ (Å)5.545.525.52, 5.78, 5.865.86, 5.99, 6.11
${d}_{\text{Ir}}^{\mathrm{i}\mathrm{n}\mathrm{t}\mathrm{e}\mathrm{r},ab}$ (Å)I5.785.775.909, 5.91, 5.915.97, 5.97, 5.99
∠Ir–O1–Ir (°)85.9285.9685.7582.55
∠Ir–O2–Ir (°)  80.0679.50
ΔNCF (eV)0.040.060.07, 0.240.09, 0.16

The impact of the crystal geometry on the magnetic properties have been studied both theoretically and experimentally. While Sakamoto et al showed that all the systems behave paramagnetically with no long range magnetic order down to 1.8 K and described the magnetic properties of these systems by Kotani's theory [143], later a combined experimental and theoretical analysis [76] of Ba3ZnIr2O9 showed that the compound exhibits a fascinating spin–orbital liquid state down to at least 100 mK. In contrast to the favorable FM interaction in absence of SOC between two Ir ions with d4 electronic configuration, the spin–orbit entangled Γ7 and Γ8 states at the neighboring Ir sites, originating from the Ir-t2g orbitals in presence of SOC, interact with each other anti-ferromagnetically (see figures 12(e) and (f)). The key to the AFM interaction in presence of SOC is the unusual hopping between pseudo-spins. Unlike the hopping between real spin that are always spin-conserved, the spin–orbital entanglement also allows for certain hopping between two states with different pseudo-spins, e.g., Γ8 electrons can virtually hop from |1α⟩ (|1β⟩) to |3β⟩ (|3α⟩) state. These electrons, which are promoted to the Γ7 state further can gain energy by hopping to the other Γ7 state as illustrated in figure 12(f). This later hopping is, however, forbidden if the interaction is ferromagnetic, thereby stabilizing an AFM interaction between the closely spaced Ir atoms forming dimers. Such AFM interaction manifests in the negative Curie–Weiss temperature θCW ∼ −30 K with an effective moment ∼0.2μB /Ir in the susceptibility measurements. The magnetic moment at the Ir site represents the hopping assisted excitation from J = 0 state to J = 1 state. We come to the detail discussion of this point in the latter part of this section.

The AFM interaction within the Ir–Ir dimer leads to the formation of SOS state. Each of these SOSs, further, anti-ferromagnetically interact with six other neighboring SOSs located on a frustrated triangular network within the ab plane and three out of the plane neighbors, corresponding to the Ir–Ir distances ${d}_{\text{Ir}}^{\mathrm{i}\mathrm{n}\mathrm{t}\mathrm{e}\mathrm{r},ab}$ and ${d}_{\text{Ir}}^{\mathrm{i}\mathrm{n}\mathrm{t}\mathrm{e}\mathrm{r},c}$ in table 3 respectively. The energy scales of these intra-dimer ${J}_{d}^{\text{intra}}$ (−14.6 meV, with neighboring single Ir ion) and inter-dimer interactions ${J}_{d}^{\text{inter}}$ (−1.5 meV for six neighbors within the ab plane and −1.6 meV with three neighbors perpendicular to the ab plane) being comparable, a strong magnetic frustration is developed in the system, giving rise to a spin–orbital liquid state, as evident from the muon-spin rotation, magnetization and heat capacity measurements [76].

The magnetic properties of the Mg compound is very similar to the Zn compound, showing noticeable AFM interaction from magnetic susceptibility measurement, although the heat capacity and μSR measurements suggest a weaker frustration in this case. The weaker frustration in the Mg compound results from the stronger ${J}_{d}^{\text{intra}}$ (−21.3 meV) compared to ${J}_{d}^{\text{inter}}$ (−0.9 meV within the ab plane and −1.4 meV out-of-plane interaction) in the system. On the other hand, the strong distortion of the IrO6 octahedra in the Sr and the Ca compounds break the local inversion symmetry, giving rise to uncompensated nearest neighbor Dzyaloshinskii–Moriya (DM) interactions. These DM interactions are expected to cause the canting of the spins, resulting in non-collinear magnetic structures. Indeed, explicit DFT calculations [112] with the symmetry allowed magnetic structures corresponding to an assumed propagation vector $ \overrightarrow {k}=(0,0,0)$ shows that the lowest energy magnetic configuration corresponds to the magnetic space group C2/c', in which the Ir ions forming the dimer have anti-parallel spin components along the x, and z directions while due to canting have a parallel y component. This net y component of spin moments from a pair of neighboring Ir ions, however, is anti-parallel to the net y moment of another pair of Ir ions, situated along the c direction, leading to a net zero moment in the system. The weak dimeric feature observed in the magnetic susceptibility of Ba3CaIr2O9 may be an outcome of the antiparallel inter-dimer arrangement of moments in the C2/c' magnetic structure as shown in figure 12(g). On the contrary, similar analysis [112] showed that another symmetry allowed magnetic configuration C2'/c' is energetically favorable for the Sr compound. Interestingly, this magnetic configuration allows for a net FM moment in the system, possibly leading to a ferro-magnetic like feature in magnetic susceptibility under low magnetic fields. Thus, the family Ba3 MIr2O9 with M = Zn, Mg, Ca, Sr constitutes an intriguing platform to investigate the effect of local structural distortion in the magnetic properties of spin–orbit entangled d4 state.

4.3.2. Hopping induced magnetism

As already discussed, a spontaneous magnetic moment, due to the large Ir–Ir hopping, is developed at the Ir site in Ba3 MIr2O9 (M = Zn, Mg, Ca, Sr). In contrast to the double perovskite iridates, where the Ir atoms are separated apart, in the 6H perovskite structure the Ir atoms form dimer, which constitute the basic building block of the 6H structure. This strong Ir–Ir dimer interaction manifests itself in the RIXS spectrum, that probes the low energy excitations of these iridates. In contrast to the double perovskite iridates, the RIXS spectrum in Ba3 MIr2O9 has distinct features, as shown in figure 13(b), which cannot be captured within an atomic model. A meaningful description of such RIXS spectrum requires at least a two-site model with finite inter-site hopping, emphasizing the importance of the band structure effect on the ground state properties of these series of 6H iridates. Here, we provide an overview of the band structure effect and its implications to the RIXS spectra observed for Ba3 MIr2O9 systems.

Figure 13.

Figure 13. The role of hopping in the RIXS spectrum of the 6H iridates. (a) Schematic illustration of the RIXS process in general. (b) The low-energy RIXS spectra from the SOC states in Ba3ZnIr2O9 (BZIO). In marked contrast to the corresponding spectra in Ba2YIrO6 (BYIO), the spectra in 6H perovskite depicts the presence of additional inelastic peaks and spectral-weight transfer to lower energies. (c) The energy states obtained from an atomic model and a two-site model. The higher energy states with low intensity are not shown. The feature of the observed spectra for 6H iridates in (b), although cannot be explained within an atomic model but is a characteristic of a two-site model indicating the crucial role of hopping in the present iridates. (d) The deviation of the ground state ψg of a two-site model, relevant for 6H iridates, from the J = 0 state. (e) The same can be concluded from the phase diagram, showing the variation of the expectation value of J2 operator in the parameter space of t/U and λ/U. The black triangle in the phase diagram corresponds to the parameters t = 0.1 eV, ΔNCF = 0.02 eV, JH = 0.5 eV, λ = 0.25 eV and U = 2.0 eV, relevant for the iridates.

Standard image High-resolution image

We begin with a brief description of the RIXS process. RIXS is a second order 'photon-in photon-out' spectroscopic process, as illustrated schematically in figure 13(a). In case of d4 iridates, the incoming photon with frequency ω1 excites the Ir-2p core electron to the Ir-d, t2g unoccupied states (corresponding to the L3 absorption edge) forming an intermediate state $\vert n\rangle =\vert 2{p}_{\frac{3}{2}}^{\enspace \text{hole}}{t}_{2g}^{5}\rangle $ with core-hole. To fill up the hole in the core level, a different t2g electron corresponding to the initial ${t}_{2g}^{4}$ configuration may come back to the 2p state resulting in an emission of a photon of frequency ω2 and a final state $\vert f\rangle =\vert {t}_{2g}^{4}\rangle $. Note that the final state |f⟩ can be any kind of arrangement of four electrons within the three t2g orbitals and the energy difference ω1ω2 corresponds to these dd excitation energies. The estimation of these energies are, therefore, beyond the scope of single-particle description and it is important to consider a many-body Hamiltonian with proper interactions.

The typical RIXS spectrum for Ba3 MIr2O9 is shown in figure 13(b). Such a spectrum, in particular the additional peaks as shown in the figure, cannot be described by the triplet (J = 0 → J = 1) and quintet (J = 0 → J = 2) spin–orbit excitons of an atomic model for spin–orbit coupled d4 systems. Indeed, the two-peak feature is the characteristic of a two-site model with a finite inter-site hopping (see figure 13(c)), indicating the necessity to go beyond the atomic limit to incorporate the non-local effects, such as non-cubic crystal field and the hopping, present within the solid.

The spin–orbit exciton energies in the RIXS spectra may be mapped into the corresponding differences in the energy eigenvalues computed using exact diagonalization of the two-site model (similar to the Hamiltonian discussed in section 3.2, equation (6) with an additional non-cubic crystal field term). In the two-site model, the atomic states interact with each other in presence of significant hopping giving rise to a new set of spin–orbit coupled states, e.g., atomic states J = 1 (J1) at each sites interact with each other to give rise to a new set of nine states, which we denote here by J1J1 for simplicity. The resulting new sets of spin–orbit coupled states are: J0J0 (1), J0J1 (6), J1J1 (9), J0J2 (10), J1J2 (30), J2J2 (25), etc, where the number within the parenthesis indicates the number of states arising due to interaction between the two-atomic states. The excitation from the J0J0 state to the states J1J1 and J0J2 states describe the broad peak around ∼0.4 eV in the RIXS spectrum, observed in the experiments.

An estimation of the various parameters, starting from a realistic set of values extracted from the ab initio calculations, may be obtained when the theoretically computed excitation energies coincide with the extracted energy loss values within the uncertainty of the measurement (full width at half maximum of the Lorentzian function). Note that unlike the atomic model, the energies of the spin–orbit excitons obtained from the two-site model depend on the Coulomb repulsion U possibly due to the superexchange energy scale ∼t2/U involved in the latter case. The estimated Hund's coupling JH is about ∼0.45 eV, which is comparable to the other 5d5 materials but higher than the d4 double perovskite iridates [136, 144]. More interestingly, the SOC λ is highly suppressed to a value of ∼0.26 eV, highlighting the decisive role of hopping in these 6H iridates. The large intra-dimer hopping and the strongly suppressed SOC leads to deviation of the two-site ground state from the ideal J = 0 state by about 10%, resulting in a non-zero moment as also evident from the ground state expectation value of ⟨J2⟩ (see figures 13(d) and (e)). In a nut shell, Ba3 MIr2O9 is an exemplary d4 system, where the hopping dramatically affects the spin–orbit coupled ground state, leading to formation of moment at the 'atomically non-magnetic' Ir site.

4.4. Hexagonal iridates Sr3MIrO6

The pentavalent iridate Sr3 MIrO6 crystallizes in the hexagonal structure with the $R\overline{3}c$ space group symmetry, which is isostructural to the parent compound Sr4IrO6. The doping with alkaline metals M, e.g., Na, Li, K at the Sr site results in a reduced charge state of Ir, viz, the pentavalent iridate, the subject of the present review. To the best of our knowledge these pentavalent iridates have been synthesized as early as 1996 [117, 118, 122]. The crystal structure of Sr3 MIrO6 is quite interesting in the sense that the Ir atoms are far apart from each other, constituting isolated IrO6 octahedral unit. These octahedra share their face with the neighboring MO6 trigonal prism, forming thereby an alternating chain of IrO6–SrO6 along the crystallographic c-axis. The crystal structure of Sr3 MIrO6 consists of such well separated alternating chains, as shown in figure 14. The large Ir–Ir distance and the lower connectivity of Ir atoms results in a narrow Ir-t2g bands, which, in turn, gives rise to the hope of realizing the J = 0 state. Interestingly, existing experiments [117, 118] indicate a temperature-independent paramagnetic behavior for Sr3 MIrO6, with M = Na, Li, which further points towards a non-magnetic J = 0 state.

Figure 14.

Figure 14. Crystal structure of Sr3 MIrO6, showing the alternating chain of IrO6–SrO6 running along the crystallographic c-axis. The face-shared IrO6 and MO6 unit is shown separately. The Ir atoms, both within the chain (dintra), and between the chains (dinter), are quite well separated in the structure, which leads to narrow Ir-t2g bands.

Standard image High-resolution image

With this viewpoint, recently DFT calculations [116] have been performed on both Na and Li compounds. These calculations find a non-magnetic state for these two systems. With the projection on to the jeff states, the authors show that the jeff = 3/2 and jeff = 1/2 states are very well separated in energy without any mixing between them. Interestingly, the J = 0 states, proposed for this system, are insulating in nature even without Coulomb correlation. This is in sharp contrast to the other d4 iridates, discussed previously in this review, where finite Coulomb correlation is always required to open up an energy gap. The insulating property of these systems are, therefore, better described as SOC driven band insulating state, rather than the Mott insulating state. A simple explanation to this behavior lies within the reduced dimensionality of the structure, resulted from the isolated IrO6 octahedra, as discussed before. As result of this structural low dimensionality, the bandwidth also reduces which, in turn, favors an insulating state. Notably, the bandwidth of these materials are almost half of the Ir-t2g bandwidths found in the double perovskite and 6H perovskite structures which may be conducive for the realization of the J = 0 state.

Interestingly, recent experiments [145] however point towards magnetism at the Ir site for both Na and Li compounds, with even larger effective magnetic moment. Notably, the IrO6 octahedra in these systems are not ideal and undergo distortions. It might be, therefore, interesting to revisit these materials both experimentally and theoretically with the new insights at hand to confirm or refute the presence of J = 0 non-magnetic state in Sr3 MIrO6. It is important to point out here that the isostructural d5 sister compound Sr3NiIrO6 exhibits a strong single-ion magnetic anisotropy [146]. In presence of any magnetism in the d4 counterparts, it would be furthermore interesting to see whether such single-ion magnetic anisotropy is also present in pentavalent iridates.

5. Outlook

Understanding electronic and magnetic properties of iridates is a rapidly growing field of research, a part of which is pentavalent iridate. While most of these pentavalent iridates have been synthesized long ago, e.g., the first d4 Ir-based material was synthesized as early as 1960s [72, 73], understanding and revealing many of the interesting properties of these materials remain elusive until recently. The surge in the interest of discovering intriguing properties of d4 iridates are primarily driven by the theoretical proposal of excitonic magnetism in these materials. While theoretical and experimental understanding of the ground state of real pentavalent iridate systems are not always conclusive, the enormous curiosity to understand the electronic and magnetic properties of these materials over the past decade builds up the solid ground to further tune and manipulate the states to get desired properties of matter. While the review of each of the members of this giant family of materials is beyond the scope of the present article, the present review provides an overall view of this family, highlighting the important issues that have been of interest to the scientific community and focusing on a few specific materials that have received considerable attention over the time.

It is hard to draw any conclusion to the iridates research as the field is still at its peak with more than 300 publications over the past five years (according to the 'iridates' keyword search in the web of science) and many more avenues to be explored in future. We, therefore, end this review by drawing the attention of the readers to a few interesting open questions or directions that still need to be answered or ventured.

The ground state properties of d4 double perovskite iridates have been at the centre of debate and controversies. The Ir atoms do not order magnetically in most of the d4 iridates, including the double perovskite systems. A more clear and better understanding of the ground states of these systems may be obtained by studying the magnetic properties of these systems under pressure or strain. Intuitively speaking, the tensile strain would separate the Ir ions far apart, leading to more atomic like scenario, where the spin–orbit effects prevail. On the other hand, the compressive strain would place the Ir ions close to each other, and if the strain is strong enough a competition between SOC and covalency may result in. This, in turn, may lead to the onset of long range order. It might be interesting to see the effect of strain on the spin–orbit excitons as well, which may be probed in RIXS experiments. The spin–orbit excitons have received a great deal of attention in recent times, particularly driven by the fact that its condensates near phase transitions exhibit dispersive Higgs mode, a long-sought-after physical phenomenon [147]. External perturbations such as pressure or strain may trigger this critical transition points, that results from the competition between various non-local effects and spin–orbit effect, which may be difficult to attain intrinsically in the bulk iridates.

The strain effects might also shed light into the insulating nature of the iridates. While most of the iridates are found to be insulating, their origin of the insulating behavior, as discussed in the literature, is actually quite different. For example, while NaIrO3 and Sr3 MIrO6 are close to band insulating state [111, 116], double perovskite iridate Ba2YIrO6 is described as a Mott insulator [87]. It is noteworthy that reference [138] also indicated possible connection between the AFM configuration and the insulating state. Structural modifications with pressure or strain might be useful to understand and analyze the transport properties of the different d4 iridates, revealing also the connection between magnetism and insulating state, if any. It is interesting to point out here that the tetravalent perovskite iridate exhibits dramatical changes in the electrical conducting properties in the two-dimensional limit [148, 149]. The corresponding behavior of the pentavalent counterparts may be worth investigating.

While the magnetic interactions in the d4 iridates are either very weak or completely absent, it would also be interesting to look for higher order 'hidden' parameter, magnetic and/or non-magnetic, which may act as primary order parameters, governing the properties of the materials. Since SOC, in general, is often found to promote such hidden orders, d4 iridates are excellent materials to investigate them. Interestingly, such hidden orders have been predicted and also experimentally verified for 5d2 double perovskites [150, 151]. Since the d4 electronic configuration is particle-hole analogue of d2 configuration (two electrons in t2g orbitals are equivalent to four holes in the t2g orbitals), it would be interesting to explore the studied pentavalent iridates from this point of view.

Undoubtedly there are still plenty of avenues to be explored both theoretically and experimentally, motivating many more future researches on the pentavalent iridates in the coming years.

Acknowledgments

We take this opportunity to deeply acknowledge Avinash V Mahajan, Sugata Ray, Tanusri Saha-Dasgupta, Santu Baidya, Atasi Chakraborty, Shreemoyee Ganguly, Abhishek Nag, and Swarup K Panda, who provided us with new insights and inspiring ideas and fruitful collaborations. We thank all our collaborators in the iridate research over the last few years: Fabrice Bert, Pabitra K Biswas, I Carlomagno, Anna Delin, Anna Efimenko, Olle Eriksson, Paul G Freeman, Adrian D Hillier, Mitsuru Itoh, Som Datta Kaushik, Ying Li, Martin Mansson, Roland Mathieu, Philippe Mendels, Carlo Meneghini, Srimanta Middey, Lars Nordström, Jean-Christophe Orain, Henrik M Ronnow, Marco M Sala, Tapati Sarkar, Dipankar Das Sarma, Denis Sheptyakov, Vasudeva Siruguri, Mark T F Telling, Roser Valentí. We also thank Bernhard Keimer, Giniyat Khaliullin, Daniel I Khomskii, Arun Paramekanti, Sashi Satpathy, and Nandini Trivedi for many useful discussions. ID thanks Science and Engineering Research Board (SERB), India (Project No. EMR/2016/005925) and Department of Science and Technology-Technical Research Centre (DST-TRC) for financial support.

Data availability statement

The data that support the findings of this study are available upon reasonable request from the authors.

Appendix.: l eff = −1 state

In presence of strong octahedral crystal field, atomic Ir-d levels are split into three-fold degenerate t2g and an eg doublet. Now by writing ${L}_{x}=\frac{1}{2}({L}_{+}+{L}_{-})$ and ${L}_{y}=\frac{1}{2i}({L}_{+}-{L}_{-})$ in terms of ladder operators and following the orbital momentum algebra,

Equation (9)

we can calculate the matrix elements of ⟨t2g |L|t2g ⟩,

Equation (10)

where t2g ≡ {dyz , dxz , dxy }. Note that here we have dropped out the ⟨t2g |L|eg ⟩ terms, assuming that the t2g states are completely decoupled from the eg states, due to large separation in the energy scale, determined by Δ. In reality, however, the t2g eg mixing may not be ignored. As discussed in reference [152], the mixing may be crucial in interpreting the branching ratio, measured in experiments as well as in theory to construct effective models for the purpose of studying the magnetic phases and magnetic excitations. Nevertheless, if such mixing be ignored, the projection of the L operator for the d-electrons on the t2g manifold, as shown in equation (10) may be identified as L(t2g ) = −L(p), where L(p) is the matrix representation of the L operator in the basis of p ≡ {px , py , pz }, indicating that the t2g orbitals effectively behave like p-orbitals having angular momentum leff = −1, but with an extra negative sign.

Please wait… references are loading.