Introduction

The 5d elements of the periodic table possess singular properties including a strong coupling of the electronic spin to its orbit as well as spatially extended valence orbitals leading to a reduced electronic repulsion and large effect on the ligand field. These intrinsic characteristics have been experimentally or theoretically demonstrated to be responsible for new exotic states of matter such as spin-orbit (SO) Mott insulators1,2, topological insulators3,4,5, super-conductors6,7,8, spin–liquids and –ices9,10,11 and for quantum metal–insulator transitions12. The common denominator in these promising materials, collectively referred to as ‘iridates’ (herein named oxido-iridates), is the presence of octahedrally coordinated IrIV ions (Fig. 1a,b) featuring a t2g5 electronic configuration. The half-filled jeff=1/2 level (Fig. 1c)13, resulting from the SO coupling, is at the origin of a narrow band gap responsible for the abovementioned phenomena. In all oxido-iridates including Sr2IrO4, the archetypal SO Mott insulator2, corner- or edge-sharing {IrO6}8− distorted octahedra are present (Fig. 1a). While magnetic IrIV–IrIV interactions and their implications for the physical properties have been studied extensively in oxido-iridates14,15, the intrinsic magnetic properties16 of the elemental {IrO6}8− moiety are consistently masked by the same long-range interactions. Birol and Haule17 recently suggested that the design of materials incorporating isolated IrIV octahedra should facilitate smaller bandwidths and promote Mott ground states. If this strategy is developed to its logical end, a discouraging result is obtained as the isolation of the [IrO6]8− ion (Fig. 1b) is chemically impossible. However for the realization of dimensionally reduced oxido-iridates, the fluoride ion appears as an ideal substitute to oxide being isoelectronic with comparable chemical and physical characteristics, but, importantly, with a reduced charge. Along this idea, we present the synthesis of molecular fluorido-iridates incorporating spatially isolated [IrF6]2− units. Their local magnetic properties are probed by X-ray magnetic circular dichroism (XMCD) spectroscopy demonstrating that the [IrF6]2− and {IrO6}8− units possess virtually identical electronic ground states as suggested by theory17. Thus this fluorido-iridate moiety and its intrinsic properties, experimentally determined in this work, can be confidently used to model and emulate the basic {IrO6}8− unit in oxido-iridates.

Figure 1: Oxido-iridates versus fluorido-iridates.
figure 1

Ball and stick representations of (a) the oxido IrIV-based layer in Sr2IrO4, (b) its smallest {IrO6}8− unit and the isoelectronic fluoride counterpart, [IrF6]2−. (c) Energy level diagram for IrIV (5d5; t2g5 electronic configuration). The octahedral ligand field splits the 5d orbitals into eg and t2g levels, and SO coupling further lifts the degeneracy of the t2g levels into filled jeff=3/2 and half-filled jeff=1/2 levels. Note that this one-electron picture is equivalent to the SO splitting of the 2T2g (Oh) term into an Jeff=1/2 ground state and excited Jeff=3/2 state.

Results

Syntheses and structures

Highly water-soluble Na2[IrF6] was obtained by direct F2 fluorination of Na2[IrCl6]·6H2O at elevated temperature (200 °C) and characterized by powder X-ray diffraction18,19,20. Addition of a solution of PPh4Cl (PPh4+=tetraphenylphosphonium(V) cation) to an aqueous [IrF6]2− solution yielded (PPh4)2[IrF6]·2H2O (1, Fig. 2a and Supplementary Fig. 1). Electrospray mass spectrometry on 1 indicated complete absence of any hydrolysis products such as [IrF5(OH)]2− (Supplementary Fig. 2). 1 was subsequently used to synthesize Zn(viz)4[IrF6] (2, Fig. 2b and Supplementary Fig. 3; viz=1-vinylimidazole). These compounds incorporate a low symmetry, but close-to-octahedral [IrF6]2− unit in 1 (P space group) and a tetragonally distorted octahedral [IrF6]2− unit in 2 (P42/n space group). The axial elongation in both 1 and 2 is 1%, slightly smaller than, for example, the 3% found in Sr2IrO4, and the bond angles are all within 1.5 and 0.1% of 90°, respectively. To elucidate any differences in the IrIV electronic structure when modifying the ligand field, we included in this study the related chloride complex, (PPh4)2[IrCl6] (3, Fig. 2c and Supplementary Fig. 4), also featuring an approximately octahedral, but axially compressed (0.7%) [IrCl6]2− ion. The nearest neighbour Ir–Ir distances (Å) are 10.0 (1), 8.1 (2) and 10.1 (3), which are much longer than in the oxido-iridates (for example, 3.9 Å in Sr2IrO4). Thus the IrIV–IrIV interactions are negligibly small and the [IrX6]2− (X=F, Cl) unit can be considered as magnetically isolated in 1, 2 and 3.

Figure 2: X-ray structure view of the molecular iridates.
figure 2

Thermal ellipsoid plots of 1 (a), 2 (b) and 3 (c) are shown at 80% probability level. The counterions for 1 and 3 and the auxiliary parts of the 1-vinylimidazole ligands in 2 are omitted for clarity. Colour code: Ir, yellow; Zn, grey; Cl, dark green; F, pale green; O, red; N, blue; H, white. Selected bond lengths (Å) and angles (°) for 1: Ir–F 1.9339(8)–1.9510(8), F–Ir–Fcis 88.89(4)–91.35(4); for 2: Ir–F 1.9281(1), 1.9449(1) (only two crystallographically different bond lengths), F–Ir–Fcis 90, 90.0112(1); for 3: Ir–Cl 2.3205(3)–2.3370(3), Cl–Ir–Clcis 88.54(1)–91.47(1).

Magnetic properties and X-ray spectroscopy

The susceptibility-temperature product, χT, of 13 (Supplementary Fig. 5) is practically temperature independent (Curie's law) confirming the lack of significant magnetic interactions between IrIV spins and thus the absence of any magnetic order at least down to 1.8 K. The complete set of magnetization data (M versus μ0HT–1; Supplementary Figs 6–8) could be fitted to the Brillouin function for an effective spin-1/2 (with g≈2), which corroborates the Curie susceptibility and demonstrates the presence of an energetically isolated Jeff=1/2 ground state.

Orbital (Morbital) and spin (Mspin) contributions to the total magnetic moments (Mtotal) have been determined experimentally using XMCD, which is defined as the difference between two X-ray absorption spectra (XAS) recorded with either opposite helicity or magnetization direction. XAS spectra were collected on 13 at the iridium L2,3 edges under an external magnetic field of μ0H=±17 T and at low temperatures (T=2.6–2.9 K). Since the measurements were performed on powdered samples, the isotropic L2,3 XAS spectra could be approximated as (σ++σ)/2, where σ+(σ) is the absorption cross-section obtained with helicity and magnetization aligned either parallel (+) or antiparallel (–) (Fig. 3). The XAS spectra are dominated by strong resonance peaks (‘white lines’) at both the L3 (2p3/2→5d3/2,5/2) and L2 (2p1/2→5d3/2) edges with the L3 edge being significantly more intense.

Figure 3: X-ray spectroscopy.
figure 3

X-ray spectra of 13 showing the isotropic XAS (positive values) and XMCD (negative values) obtained in a magnetic field of +17 T (at 2.7, 2.6 and 2.5 K for 1, 2 and 3 respectively). The spectra of 2 and 3 were shifted horizontally (35 and 70 eV, respectively) and vertically for clarity. The filled patterns are the integrals used for the sum rule analysis. Inset: Field dependence of the magnetization, M versus μ0HT−1, of a polycrystalline sample of 2 at T=2.0 K. ‘XMCD’ designates the field dependence of the XMCD maximum signal at the Ir L3 edge at T=2.6 K. The green line is the best fit of the magnetization and scaled XMCD data to the Brillouin function.

The spectra of 1 and 2 are virtually identical, whereas for 3, a slight shift of the white line peak of ca. 2 eV towards lower energies is observed concomitantly with an additional component at ca. 12 eV higher energy (Supplementary Fig. 9). The latter feature is likely the signature of excitations into delocalized states originating from ligand to metal charge transfer that is not observed in 1 and 2 due to the more ionic Ir–F bond. The effect of SO coupling on the 5d states was quantified through the so-called SO sum rule21. It relates the branching ratio (BR) of the white line integrals at the SO split absorption edges to the expectation value of the angular part of the ground state SO operator per hole. For the p→d transitions (L2 and L3 edges), the branching ratio can be expressed as

where, 〈nh〉=〈nhj=3/2〉+〈nhj=5/2〉 is the total number of holes in the 5d levels and 〈i li·si〉/ħ2=–3/2 × 〈nej=3/2〉+〈nej=5/2〉 is the expectation value of the one-electron li·si operator summed over all electrons; 〈ne〉 being the occupation number of the corresponding levels. The white line integrals are larger in 3 than in 1 or 2, thus more holes in the 5d band should be present for 3 compared with 1 or 2. If one assumes that 1 and 2 are completely ionic, 〈nh〉=5, we obtain 〈nh〉=5.26 for 3. The corresponding results are shown in Table 1. The quantitative analysis of the XMCD spectra was performed by means of the magneto-optical sum rules providing direct access to Morbital=–〈LzμB and Mspin,eff=–2 〈SeffμB (the details of this analysis are given in the ‘Methods’ section; Table 1)22,23. The scaling of the field dependence of the XMCD to the bulk magnetization (inset Fig. 3 and Supplementary Fig. 10; see the ‘Methods’ sections) allows the determination of the absolute value of the magnetization at μ0H=+17 T: Mtotal=1.03, 1.0 and 0.96 μB for 1, 2 and 3 respectively. From these values, the magnetic dipole contribution, 〈Tz〉, and Mspin (given by Mspin,eff+7〈Tz〉) can be estimated (Table 1) without the need of any sophisticated theoretical modelling from Mtotal=Mspin+Morbital considering Mspin,eff and Morbital deduced from the sum rules. The magnetic dipole contribution has never been experimentally determined for any iridate systems, but according to these results, 〈Tz〉 cannot be neglected without a significant underestimation of the Morbital/Mspin ratio, as previously anticipated24.

Table 1 Summary of the X-ray spectroscopy results.

Theoretical considerations

The molecular nature of 1, 2 and 3 implies that intuitive localized bonding models should be able to disentangle the magnetic moment contributions. The angular overlap model (AOM) allows for a decomposition of the ligand field potential into σ- and π-bonding parameters25. As a first approximation, a cubic [IrF6]2− model considering only σ-bonding (with ΔO=27,000 cm–1, Racah interelectronic repulsion parameters B=510 cm–1 and C/B=4.9, and the SO coupling constant ζ=3,300 cm–1; Fig. 1c)26 affords 〈i li·si〉/ħ2=–2.65, Morbital=0.74 μB and Mspin=0.36 μB, which are close to the ionic values (Morbital=0.67 μB and Mspin=0.33 μB) for a pure Jeff=1/2 system2. Whereas the agreement with the experimental results of the fluorido-iridates (Table 1) for 〈i li·si〉/ħ2 and Morbital is surprisingly good, Mspin is overestimated in this simplistic approach which neglects the covalency. A more realistic model including π-bonding and scaled AOM parameters to account for the different Ir–F bond lengths (see the ‘Methods’ section) leads to the same conclusion: 〈i li·si〉/ħ2=–2.65, Morbital=0.72 μB and Mspin=0.32 μB. However, when covalency is explicitly accounted for by CASSCF calculations, the spin contribution (Mspin=0.22 μB) is in much better agreement with the experiment, but the orbital contribution becomes slightly underestimated (Morbital=0.53 μB). Indeed, this departure from cubic symmetry has little influence on the energy of the first excited Jeff=3/2 state (see Fig. 1 caption) as evidenced by calculations (Supplementary Table 2) and near-IR absorption spectroscopy revealing absorption bands from 6,000 to 8,500 cm–1 (Supplementary Fig. 11).

Electron paramagnetic resonance spectroscopy

For an ideal Jeff=1/2 state, magnetic anisotropy is absent, but minuscule deviations from cubic symmetry may result in strong g-factor anisotropy that can be probed experimentally by electron paramagnetic resonance (EPR) spectroscopy. The X-band (v=9.634 GHz) EPR spectrum of an [IrF6]2− doped Zn(viz)4[ZrF6] single crystal (1% Ir) was measured at 5 K (Supplementary Figs 12–16; Table 2). The experimental eigenvalues of the g-tensor are indeed remarkably anisotropic with gz=1.37 and gxy=2.11 in good agreement with the CASSCF calculations leading to gz=1.30 and gxy=2.24 (Supplementary Table 2).

Table 2 Hyperfine and superhyperfine coupling parameters for [IrF6]2− doped in Zn(viz)4[ZrF6] (1% Ir).

These combined experimental and theoretical results establish that the electronic ground state of the molecular [IrF6]2− species is Jeff=1/2 as suggested for the {IrO6}8− unit, present in all oxido-iridates.

Dynamic magnetic properties

The magnetization dynamics of the molecular fluorido-iridates was studied by a.c. susceptibility of 2 at different temperatures below 20 K (Fig. 4 and Supplementary Figs. 17–22 for 1, 2 and 3). The presence of peaks in the a.c. frequency (v) dependence of the imaginary component, χT (Fig. 4b), which shift with temperature, clearly indicates the slow relaxation of the magnetization, while the vanishing of the real component, χT, in the adiabatic limit (v→∞) reveals a blocking of the magnetization that concerns the whole volume of the material. Note that for the three compounds, the slow dynamics of the magnetization is observed only on the application of a small static magnetic field that likely serves to decouple the IrIV Jeff=1/2 from nuclear spins as justified by a comparable magnitude of the applied field (75 mT) to the width of the EPR spectra (Supplementary Fig. 15).

Figure 4: Magnetization dynamics.
figure 4

Two-dimensional frequency/temperature maps of the real (χT, a) and imaginary (χT, b) components of the a.c. susceptibility-temperature product for a polycrystalline sample of 2 obtained under μ0H=75 mT. Left part: representative 4-K data of the χT and χT frequency dependence with the solid black line being the best fit to the generalized Debye model27, that was used for each temperature to determine the relaxation time (τ) shown as green dots on the contour plots. The best fit of τ−1 versus T is shown as solid white lines (see text). Dashed lines are guides for the eyes.

The real and imaginary components of the a.c. susceptibility could be well-fitted to a generalized Debye model27 with small values of the distribution parameter (Supplementary Fig. 23) reflecting a single characteristic relaxation time. The temperature dependence of the extracted relaxation time is shown for 2 in Fig. 4 (and also Supplementary Fig. 24 for 13). In these molecular iridates, the spin-lattice relaxation rates τ–1 were modelled considering Raman and phonon-bottlenecked direct processes, described as a sum of power laws, τ–1=CT7+DT2, and leading to C=46 × 10–3 s–1 K–7, D=17 s–1 K–2 for 2 (ref. 28). The magnetization dynamics appears to be similar in 1, 2 and 3 (Supplementary Fig. 24), suggesting that this property is an intrinsic characteristic of the IrIV electronic structure. The slow relaxation of the magnetization in 2 was confirmed by muon spin relaxation (μ+SR) measurements at low temperatures (above 1.9 K; Supplementary Figs 25 and 26), where the implanted muons probe the local dynamics of magnetic fields, thereby ruling out any long-range order. This conclusion is further supported by magnetization measurements on an oriented single crystal of 1 and 2 below 2 K (Supplementary Figs 27–30). The magnetic susceptibility (χ=dM/dH) estimated from these experiments in the zero-field limit follows a Curie–Weiss law, χ=C/(Tθ), down to lowest accessible temperature of 0.03 K, confirming the absence of magnetic order (Supplementary Fig. 31). For 2, the magnetization reaches saturation faster when the d.c. field is applied perpendicular to the C4 axis rather than along this axis (Supplementary Figs 29 and 30), confirming the easy-plane magnetic anisotropy in agreement with the extracted g-factors (vide supra). The temperature and magnetic field sweep rate dependence of the magnetization revealed the existence of a weak hysteretic behaviour. At the lowest temperatures (<0.4 K) and in agreement with the a.c. susceptibility data (Fig. 4), butterfly-shaped M versus μ0H hysteresis loops for both 1 and 2 (Supplementary Figs 27–30) confirm the presence of phonon-bottlenecked direct processes that dominate the magnetization relaxation29.

Discussion

The principle of dimensional reduction in solid-state structures is based on the formation of a derived compound, AnaMXx+n, from a parent MXx precursor and an AaX salt, thereby forcing termination of M–X–M polymerization30. If A is voluminous, for instance an organic cation, child compounds with structurally and magnetically isolated molecular units can be formed. For many oxides, the ultimate dimensional reduction to molecular {MOx}y is impeded by the progressive development of large localized negative charges (for example, for {IrO6}8−). Thus, truly single-metal ion analogues of most oxides are difficult, if not impossible, to isolate. In the case of octahedral species, no example has been reported so far. As discussed herein for IrIV, the exchange of oxide with a less negatively charged fluoride results in dimensional reduction of the iridate as compared with the ternary iridium oxide parents. Although their local structures are obviously comparable, the electronic resemblance between these compounds should be discussed here. The branching ratios extracted from XAS (0.85; Table 1) are identical for 1 and 2, and very close to the values found for oxido-iridates, for example, 0.87 for Sr2IrO4 and 0.85 for Y2Ir2O7 (refs 31, 32). In addition, the L2/L3 XMCD intensity ratio of 4.7% for 2 is almost identical to the value determined for Sr2IrO4 (5%) (ref. 33). This striking agreement demonstrates the resemblance of the electronic structure of the {IrO6}8− moieties in oxido-iridates and the molecular [IrF6]2− unit. A Morbital/Mspin=〈Lz〉/2〈Sz〉 ratio of 2.5 in Sr2IrO4 was recently obtained by non-resonant magnetic X-ray scattering that circumvents the need for the estimation of 〈Tz〉 (ref. 34). For comparison, Morbital/Mspin ratios of 3.3, 3.3 and 2.1 are obtained for 1, 2 and 3, respectively. The slightly larger ratio for 1 and 2 as compared with Sr2IrO4, and 3 is attributed to the weaker covalency of the Ir–X bond for X=F than for X=O, Cl, as reflected in the nephelauxetic series35. It is worth emphasizing that the fluoride ion is the least nephelauxetic of all known ligands, and thus the [IrF6]2− moiety has the closest proximity to a perfectly ionic iridate. As a less-pronounced covalency would induce a larger Morbital/Mspin ratio36, a larger orbital magnetic moment is found in [IrF6]2− over {IrO6}8− and [IrCl6]2−, as expected.

In oxido-iridates, the purity of the Jeff=1/2 state depends crucially on the structural deviation from the octahedral symmetry, rendering oxido-iridate systems with weakly distorted {IrO6}8− octahedra highly interesting37. Remarkably, the [IrF6]2− moiety in 2 is structurally closer to cubic than any reported oxido-iridate system. Therefore fluorido-iridates are promising materials for stabilizing an ideal Jeff=1/2 Mott ground state as was already concluded from the experimentally determined values of Morbital and Mspin. Paramagnetic relaxation in oxido-iridates has been reported for diamagnetically doped honeycomb systems where the imperfect stoichiometry resulted in the blocking of the magnetization at low temperature38,39. Our low-temperature experiments reveal that the slow dynamics of the magnetization is indeed intrinsic to the IrIV centre, that might have implications for quantum magnetism.

Herein, we have established that a molecular [IrF6]2− species and the {IrO6}8− unit, present in all oxido-iridates, possess the same electronic Jeff=1/2 ground state. The dominating role of the SO coupling on the peculiar magnetism intrinsic to IrIV was elucidated by studying spatially isolated, magnetically decoupled, [IrF6]2− ions in solids. Whereas strongly perturbed by long-range magnetic interactions in iridates, the 5d spin and orbital moments, as well as their local anisotropy and dynamics have been quantified. These experimental results provide a solid basis for building a realistic model to describe the unusual physics discovered in oxido-iridates. Moreover, as illustrated by the one-dimensional Zn(viz)4[IrF6] system (2), the [IrF6]2− species appears to be a robust building-block for the construction of molecule-based architectures. It offers an unprecedented synthetic strategy using the versatile tools of soft chemistry to engineer new materials with exotic properties anticipated to be exclusively reserved for oxido-iridates.

Methods

Synthetic methods

Na2[IrCl6]·6H2O was heated to 150 °C in vacuo for 6 h, cooled to room temperature (exp. (calc. for Na2[IrCl6]) weight loss=19% (19%)) and subsequently heated to 200 °C in a flow reactor under F2 gas (10% diluted in Ar) and kept for 6 h. The last step was repeated to improve the crystallinity of the white product (exp. (calc. for Na2[IrF6]) weight loss=19% (22%)) that was identified as Na2[IrF6] by powder X-ray diffraction (Supplementary Fig. 32). A solution of Na2[IrF6] (68 mg, 0.19 mmol) in water (1 ml) was added a solution of PPh4Cl (200 mg, 0.53 mmol) in water (3 ml). The resulting solution was left undisturbed for 1 h to produce crystalline 1 (114 mg, 59%). Anal. calcd. (found) for C48H44F6IrP2O2: C: 56.46% (56.58%), H: 4.34% (4.19%). 1 can reversibly be oxidized to [IrF6] (E1/2=+0.8 V versus Fc+/Fc; Supplementary Fig. 33). The addition of 1-vinylimidazole (0.80 g, 8.5 mmol) to a methanol solution (40 ml) of 1 (100 mg, 0.098 mmol) and Zn(NO3)2·6H2O (100 mg, 0.34 mmol) afforded crystals of 2 after 1 day. Yield: 60 mg (82%). Anal. calcd. (found) for C20H24F6IrN8Zn: C: 32.11% (32.36%), H: 3.23% (3.07%), N: 14.98% (14.67%). Complex 3 was obtained from (NH4)2[IrCl6] (203 mg, 0.460 mmol) that was first dissolved in aqueous HCl (0.1 M, 100 ml). The solution was heated to 50 °C and filtered hot. To this solution, a solution of PPh4Cl (360 mg, 0.963 mmol) in aqueous HCl (0.1 M, 70 ml) was added slowly, inducing an immediate precipitation of orange-brown crystals. On complete addition, the reaction mixture was cooled to room temperature and stirred for 30 min leaving only a slightly coloured solution. The product was filtered off and washed three times with water and dried in a dynamic vacuum. Yield: 453 mg (91%). Anal. calcd. (found) for C48H40Cl6IrP2: C: 53.20% (53.24%), H: 3.72% (3.67%).

Crystallography

Powder X-ray diffraction patterns of Na2[IrF6] were obtained on samples mounted under nitrogen in an air-tight aluminium alloy cell and collected on a PANanalytical Bragg–Brentano θ-2θ geometry diffractometer (Cu Kα radiation) equipped with a secondary monochromator over an angular 2θ range of 5–80°. Rietveld refinement of the data was performed in the TOPAS (version 4) program40. Atomic positions were allowed to refine freely without restraints, while thermal parameters for all atoms were fixed. A suitable fit (Rwp=13.46%) was obtained using a structural model similar to that of Na2[GeF6] (ref. 41), with an expanded unit cell (space group P321, a=b=9.32858(24) Å, c=5.13417(19) Å, α=β=90°, γ=120°; Supplementary Fig. 32). Single-crystal X-ray diffraction studies on 13 (Supplementary Table 1) were performed at 122(1) K on a Bruker D8 VENTURE diffractometer equipped with Mo Kα high-brilliance IμS radiation (λ=0.71073 Å), a multilayer X-ray mirror, a PHOTON 100 CMOS detector, and an Oxford Cryosystems low-temperature device. The instrument was controlled with the APEX2 software package. The structures were solved in Olex2 using the olex2.solve structure solution program (Charge Flipping) and refined using the olex2.refine program42. All non-hydrogen atoms were refined anisotropically and hydrogen atoms were placed at calculated positions and refined as riding atoms with isotropic displacement parameters.

Magnetometry

The magnetic measurements were carried out between 1.8 and 280 K with applied d.c. fields ranging from –9 to +9 T, with the use of a MPMS-XL Quantum Design SQUID magnetometer and a PPMS-9 Quantum Design susceptometer. Measurements were performed on polycrystalline samples sealed in polyethylene bags (typically 3 × 0.5 × 0.02 cm; 15–30 mg) and immobilized in mineral oil. a.c. susceptibility measurements were performed with an oscillating field of 1–6 Oe with a frequency from 1 to 10,000 Hz. Before the experiments, the field-dependent magnetization was measured at 100 K to confirm the absence of any bulk ferromagnetic impurities. The magnetic data were corrected for the diamagnetic contributions from the sample, sample holder and mineral oil. Micro-SQUID magnetic measurements on single crystals were carried out in the field range from –1.4 to +1.4 T and in the 5.0–0.03 K temperature range. The field was applied in any direction of the micro-SQUID plane with precision >0.1° by separately driving three orthogonal coils43.

X-ray spectroscopy

Polarization-dependent XAS spectra at the Ir L2 and L3 edges were acquired at the ID12 beam line at the European Synchrotron Radiation Facility (ESRF), Grenoble, France. A helical undulator of the APPLE-II type allowed the circular polarization of incoming X-rays to be changed for each spectrum. The spectra were obtained by using the total fluorescence yield detection mode in magnetic fields up to ±17 T. Transmission detection was additionally used for 2 to check the validity of the self-absorption correction used for all samples, which took into account the chemical composition, the geometry of the experiment and the solid angle of the X-ray fluorescence detector. XMCD spectra were obtained as the difference of consecutive XAS spectra obtained with opposite photon helicities. In addition, the XMCD spectra were systematically recorded in both field directions to ensure the absence of experimental artefacts. The isotropic spectra were obtained as the sum of spectra with right and left circularly polarized X-rays and were normalized between zero before the absorption edge and one above the edge. The position of the step function describing transitions into the continuum was defined by the following procedure. The maximum of the L3 white line was assigned to zero energy and the L2 edge spectrum was shifted in energy to have the first EXAFS oscillations perfectly overlapping for the two spectra. The step function was placed by visual comparison of the shifted spectra. It was systematically checked that shifting the energy of the inflection point of the step function around the chosen position (±2 eV) was not significantly altering the white line integrals. A step function broadening of 6 eV was used, but this parameter does not affect the integrated values of the white line intensities. The white line integral at the L2 edge was divided by 2.15 reflecting factors of (i) 2 from the occupation ratios of 2p3/2 and 2p1/2 core states and (ii) 0.15 from the fact that the sum rules are applied to the line intensities I(ħω) (where ħω is the photon energy) and not to the absorption cross sections, μ(ħω), which are related by I(ħω)=1/(4π2α ħω) × μ(ħω) (binding energies for L3 and L2 edges are 11.225 and 12.835 keV, respectively). The same normalization procedure was applied to the corresponding XMCD integrals. The precise sample temperature of 2.6–2.9 K and the absolute magnitude of the magnetic moment at μ0H=±17 T was determined by scaling the field and temperature dependent XMCD signal intensity to coincide with the M versus μ0HT−1 master curve obtained from bulk magnetometry. The normalized spectra were analysed using the magneto-optical sum rules given by equation 2 for L2 and L3 absorption edges22:

Ligand field and ab initio calculations

The AOM calculations were performed using the Ligfield programme on the full d5 configuration44,45. All matrices were constructed in the weak-field basis and subsequently transformed to the eigen-basis corresponding to the parameter vectors given below. The octahedral component of the ligand field, ΔO=27,000 cm–1, would agree with the AOM parameter eσ=9,000 cm–1 in the cubic model. In the more realistic model, a eσ/eπ ratio of 5, and a distance dependence of eσdIr-F–5 and eπdIr-F–7 (where dIr-F is the Ir-F bond length) was assumed (yielding eσz=11,904 cm–1, eσxy=12,457 cm–1, eπz=2,356 cm–1 and eπxy=2,507 cm–1). All ab initio calculations were performed with the ORCA 3.0 programme package46 using ZORA47,48 for relativistic corrections and the def2-TZVP basis set for the [IrX6]2− moieties and the def2-SVP basis set for the 1-vinylimidazole ligands in calculations when requested in the model. State-averaged CASSCF calculations were performed on experimental geometries using Kohn–Sham starting orbitals obtained with the TPSSh functional. Active spaces of CAS(5,3) (Ir t2g orbitals—3 doublet CSFs) and CAS(5,5) (Ir d-orbitals—12 doublet configuration state functions (CSFs), 6 quartet CSFs, 1 sextet CSF) were used and the SO interaction taken into account through a mean field approximation (Supplementary Table 2)49.

EPR spectroscopy

The spectra were acquired at T=5 K on a Bruker Elexsys E500 spectrometer equipped with a Bruker ER 4116 DM dual mode cavity, an EIP 538B frequency counter and an ER035M NMR Gauss-meter, and with a microwave radiation frequency of 9.634 GHz. The spectra were simulated using home-written software considering an effective electronic Jeff of 1/2 and taking into account the natural isotopic abundance of 191Ir (37.3%) and 193Ir (62.3%), which have both a nuclear IIr=3/2 spin, as well as the 19F nuclear spin (IF=1/2, 100%). The spectra were computed using the following spin Hamiltonian for each iridium, 191Ir and 193Ir, isotope situated in a tetragonally distorted octahedral environment:

Here, operators refer to the effective spin of Jeff=1/2 pertinent to the electronic ground state. , and operators refer to the nuclear spin of the iridium nucleus and , , (n=1, 2, 3, 4, 5 or 6) refer to the nuclear spin of the fluoride ligands. The fluoride ligands on the positive and negative X-axis are designated with subscripts ‘1’ and ‘2’, respectively. Similarly, subscripts ‘3’ and ‘4’, and subscripts ‘5’ and ‘6’, designate the fluoride ligands on the Y- and Z-axis, respectively. The extracted parameter values are given in Table 2. Remarkably, the estimated hyperfine interactions appear to be almost isotropic, in contrast to the strongly anisotropic superhyperfine interactions to the fluoride ions, which exhibit averaged values on the same order of magnitude as the hyperfine ones. Performing the same experiment on a concentrated crystal of 2 gave virtually identical g-factors, but the hyperfine interactions could not be resolved.

Muon-spin relaxation

The μ+SR measurements were performed on the Dolly spectrometer at the Paul Scherrer Institut (Villigen, Switzerland). Fully spin-polarized muons were implanted into the sample and acted as a local probe of the internal magnetic fields. The time evolution of the polarization, which was monitored via the anisotropic β-decay of the implanted muons (lifetime 2.2 μs), is determined by the temporal and spatial properties of the local magnetic field50. For example, fluctuating magnetic moments in the vicinity of a muon would produce a fluctuating dipolar magnetic field, and therefore, a polarization relaxation from its initial value to zero. The time scale of the relaxation can be extracted from the field dependence of the muon spin-lattice relaxation, λ, considering the following equation51,

where, Δ is the magnitude of the local magnetic field, γ is the gyromagnetic ratio of the muon, μ0H is the applied magnetic field and τ is the fluctuation time scale. The equation above assumes that Δ and τ are field-independent. From the fit of λ as a function of μ0H (Supplementary Fig. 26) τ0.016 μs is obtained. The small value of τ compared with that obtained from a.c. susceptibility measurements is due to the fact that μSR reflects the integrated contribution over q-space as a local, point-like probe, while a.c. susceptibility measures the response for q=0 (ref. 52).

Data availability

The X-ray crystallographic coordinates for structures reported in this Article have been deposited at the Cambridge Crystallographic Data Centre (CCDC), under deposition number CCDC 1431835–1431837. These data can be obtained free of charge from The Cambridge Crystallographic Data Centre via www.ccdc.cam.ac.uk/data_request/cif. All other relevant data are available from the authors on request.

Additional information

How to cite this article: Pedersen, K. S. et al. Iridates from the molecular side. Nat. Commun. 7:12195 doi: 10.1038/ncomms12195 (2016).