This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

Sr2IrO4/Sr3Ir2O7 superlattice for a model 2D quantum Heisenberg antiferromagnet

Hoon Kim These authors contributed equally to this work. Department of Physics, Pohang University of Science and Technology, Pohang 790-784, Republic of Korea Center for Artificial Low Dimensional Electronic Systems, Institute for Basic Science (IBS), 77 Cheongam-Ro, Pohang 37673, South Korea    Joel Bertinshaw11footnotemark: 1 These authors contributed equally to this work. Max Planck Institute for Solid State Research, Heisenbergstraße 1, D-70569 Stuttgart, Germany    J. Porras Max Planck Institute for Solid State Research, Heisenbergstraße 1, D-70569 Stuttgart, Germany    B. Keimer Max Planck Institute for Solid State Research, Heisenbergstraße 1, D-70569 Stuttgart, Germany    Jungho Kim Advanced Photon Source, Argonne National Laboratory 9700 Cass Ave, Lemont, IL 60439, USA    J.-W. Kim Advanced Photon Source, Argonne National Laboratory 9700 Cass Ave, Lemont, IL 60439, USA    Jimin Kim Department of Physics, Pohang University of Science and Technology, Pohang 790-784, Republic of Korea Center for Artificial Low Dimensional Electronic Systems, Institute for Basic Science (IBS), 77 Cheongam-Ro, Pohang 37673, South Korea    Jonghwan Kim Center for Artificial Low Dimensional Electronic Systems, Institute for Basic Science (IBS), 77 Cheongam-Ro, Pohang 37673, South Korea Department of Materials Science and Engineering, Pohang University of Science and Technology, Pohang 37673, South Korea    Gahee Noh    Gi-Yeop Kim Department of Materials Science and Engineering, Pohang University of Science and Technology, Pohang 37673, South Korea    Si-Young Choi Department of Materials Science and Engineering, Pohang University of Science and Technology, Pohang 37673, South Korea    B. J. Kim To whom correspondence should be addressed.
[email protected]
Department of Physics, Pohang University of Science and Technology, Pohang 790-784, Republic of Korea Center for Artificial Low Dimensional Electronic Systems, Institute for Basic Science (IBS), 77 Cheongam-Ro, Pohang 37673, South Korea
Abstract

Spin-orbit entangled pseudospins hold promise for a wide array of exotic magnetism ranging from a Heisenberg antiferromagnet to a Kitaev spin liquid depending on the lattice and bonding geometry, but many of the host materials suffer from lattice distortions and deviate from idealized models in part due to inherent strong pseudospin-lattice coupling. Here, we report on the synthesis of a magnetic superlattice comprising the single (nn=1) and the double (nn=2) layer members of the Ruddlesden-Popper series iridates Srn+1IrnO3n+1 alternating along the cc-axis, and provide a comprehensive study of its lattice and magnetic structures using scanning transmission electron microscopy, resonant elastic and inelastic x-ray scattering, third harmonic generation measurements and Raman spectroscopy. The superlattice is free of the structural distortions reported for the parent phases and has a higher point group symmetry, while preserving the magnetic orders and pseudospin dynamics inherited from the parent phases, featuring two magnetic transitions with two symmetry-distinct orders. We infer weaker pseudospin-lattice coupling from the analysis of Raman spectra and attribute it to frustrated magnetic-elastic couplings. Thus, the superlattice expresses a near ideal network of effective spin-one-half moments on a square lattice.

preprint: APS/123-QED

I Introduction

The physics of SS=1/2 antiferromagnet (AF) on a two-dimensional (2D) square lattice has a long history of research as it is widely believed to hold the key for the mechanism of high temperature superconductivity in copper oxide compounds [1, 2, 3, 4, 5]. However, it is rarely realized outside of the Cu-based compounds, and as a result its generic features are difficult to isolate from material specifics. Ruddlesden-Popper (RP) series iridates Srn+1IrnO3n+1, have recently emerged as a new material platform to study the same physics with spin-orbit entangled JeffJ_{\textrm{eff}}=1/2 pseudospins replacing the SS=1/2 moments in the cuprates [6, 7, 8, 9, 10]. Indeed, the single layer Sr2IrO4 has reproduced much of the cuprate phenomenology: a pseudogapped metal [11, 12, 13], and a nodal metal with a dd-wave symmetric gap indicative of possible unconventional superconductivity [14, 15] emerge upon electron doping from the parent phase that is approximately a Heisenberg AF [16, 17]. Further, experiments indicate existence of various symmetry-breaking orders: polarized neutron diffraction [18] detects time-reversal symmetry breaking above the Néel temperature (TN), magnetic torque [19] and second harmonic generation [20, 21] measurements indicate loss of C4 rotation symmetry, and resonant x-ray diffraction [22] observes splitting of a magnetic Bragg peak suggesting formation of a nearly commensurate density wave.

However, iridates are different from the cuprates in several aspects, and to what extent they are relevant to the essential physics is an open important issue. First, the dominant orbital character of the pseudospins leading to strong pseudospin-lattice coupling (PLC) [23, 24, 25], which accounts largely for the spin-wave gap, questions the validity of spin-only models. Second, structural distortions of kinds not found in cuprates [26, 27, 28] add complexity to theory models by allowing additional interactions. For example, the staggered tetragonal distortion of IrO6 octahedra in Sr2IrO4, breaking the vertical glide planes and thus lowering the space group from II41/acdacd to II41/aa [27, 26], leads to additional pseudospin exchange interactions, which provide a mechanism for locking of pseudospin canting angles and the octahedral rotation [29]. In the bilayer compound Sr3Ir2O7, the monoclinic distortion, lowering the space group from orthorhombic BbcaBbca to monoclinic CC2/cc [28] results in bending of otherwise straight Ir-O-Ir cc-axis bonds. This in turn leads to canting of the AF moments aligned along the cc-axis, manifesting as small but clearly measurable net in-plane ferromagnetic moments [30, 31]. Such distortions lead to deviation from the ideal cubic-symmetric JeffJ_{\textrm{eff}}=1/2 states on rectilinear superexchange pathways, which are assumed in theory models that predict, for example, realization of a Kitaev spin liquid in a honeycomb lattice [8, 32].

Refer to caption
Figure 1: (color online). Stacking pattern of the superlattice as imaged by STEM. (a) Wide field-of-view STEM image along [100] projection. The alternation between single-layers (blue) and double-layers (orange) is well maintained over the entire field of view. (b) Magnified HAADF-STEM image with single-layer (blue) and double-layer (orange) indicated. (c) A structural model for the superlattice. The single-layer and double-layer are shifted by a half unit cell on the SrO planes. IrO6 octahedra are rotated about the cc-axis as in the parent compounds. (d) [110]- and (e) [100]-projected HAADF (left)- and ABF (right)-STEM images overlaid with the atom positions (Sr, grey; Ir, blue, orange; O, red) from the model.

Here, we report on the synthesis of a Sr2IrO4/Sr3Ir2O7 superlattice, and provide a comprehensive study of its lattice and magnetic structures. The lattice structure is investigated by scanning transmission electron microscopy (STEM), resonant x-ray diffraction (RXD), and rotational anisotropy third harmonic generation measurements (RA-THG), and the magnetic structure by magnetometry, RXD, resonant inelastic x-ray scattering (RIXS), and Raman scattering. The superlattice is free of structural distortions reported for the parent phases while leaving their magnetic structures intact. The superlattice features two magnetic transitions with two different orders: canted abab-plane AF and cc-axis collinear AF inherited from Sr2IrO4 and Sr3Ir2O7, respectively. Their contrasting pseudospin dynamics, of Heisenberg and Ising types, also remain unchanged within our experimental resolutions. However, abab-plane magnetic anisotropy of the Heisenberg pseudospins is significantly reduced indicating weaker PLC, possibly due to the Ising pseudospins aligned along the cc-axis resisting the orthorhombic distortions. Our result shows that two distinct types of quasi-2D magnetism can be compounded in a superlattice to realize a pseudospin system closer to an ideal model.

II Lattice Structure

The superlattice has the nominal chemical composition Sr5Ir3O11 and can be regarded as a nn=1.5 member of the RP series. Although this phase is not quite thermodynamically stable, it forms transiently during a flux growth before either nn=1 or nn=2 phase eventually stabilizes depending on the starting composition of Sr2CO3 and IrO2 molten in SrCl2 flux and dwelling temperature. Thermal quenching at high temperature leads to intergrowth of both phases, and a highly ordered superlattice can be found by a careful control of the heating sequence. The resulting “crystals” have typically a few microns thickness, thicker than layer-by-layer grown thin films but thinner than typical bulk crystals. As the conventional structure refinement is limited by the small sample volume, we rely on STEM and RXD to verify the superlattice formation.

Figure 1(a) shows a wide field-of-view STEM image along [100] projection showing the stacking of single (SL) and double-layer (DL) units alternating over >> 40 unit cells. The stacking sequence is indicated in Fig. 1(b) and the unit cell is depicted in Fig. 1(c), which is modeled from the known structures of Sr2IrO4 (Ref. 33) and Sr3Ir2O7 (Ref. 34). Figures 1(d) and 1(e) show representative high-angle annular dark field (HAADF) and annular bright field (ABF) images along [110] projection and [100] projection, respectively. The images are overlaid with the atomic positions based on the unit cell in Fig. 1(c). In the [100] projection, the staggered rotation of IrO6 octahedra about the cc-axis as in the parent compounds is seen as diffuse rods (see also Figs. S1 and S2). Overall, our data is in good agreement with the model.

Refer to caption
Figure 2: (color online). RXD on the superlattice at the Ir L3L_{3}-edge. Sharp (a) (0 0 L) and (b) (0 2 L) charge reflections are centered at integer-L values (dashed lines), indicating the superlattice is well ordered across the bulk of the sample. Minor impurity peaks that match Sr2IrO4 and Sr3Ir2O7 are marked by cyan and orange sticks, respectively. Miller indices are in the orthorhombic notation; i.e., reciprocal lattice vectors corresponding to the unit cell shown in Fig. 1(c).

The superlattice formation is confirmed to represent the bulk of the sample via RXD conducted at the Ir L3L_{3}-edge. Figures 2(a) and 2(b) plot scans along (0 0 L) and (0 2 L), respectively, which show reflections centered at integer-L values with \sim 0.01 Å widths. Impurity peaks of intensities of the order of \lesssim 1 % of the main peaks are observed, which match the cc-axis lattice parameters of either Sr2IrO4 or Sr3Ir2O7. The sharp reflections and negligible impurity peaks indicate that the superlattice structure remains well correlated at a macroscopic level. Whereas the in-plane lattice parameters (\approx 5.502 Å) match those of the parent systems, the cc-axis is unique with a length of 16.93 Å. According to the study by Harlow et al. [35], however, diffraction patterns from a randomly-stacked intergrowth of nn=1 and nn=2 phases can misleadingly appear similar to those of an ordered phase. Such possibility is ruled out in our sample by selectively measuring the periodicity of DL, by exploiting the fact that only DLs are magnetically ordered in the temperature range between T= 220 K and 280 K (to be discussed in section III).

Refer to caption
Figure 3: (color online). RA-THG patterns of the superlattice (open circles) taken under (a) PPPP, (b) PSPS, (c) SPSP, (d) SSSS geometries. Incident 1200 nm light was used at room temperature. The third harmonic 400 nm light was collected as a function of azimuth-angle while the scattering plane rotates about cc-axis [36, 37]. The THG signals are normalized by the PPPP trace, overlaid with the best fits to bulk electric dipole induced THG tensor of 4/mmmmmm point group (navy lines).

Next, we further refine the structure using RA-THG, a nonlinear optical process highly sensitive to the bulk crystal symmetry through nonlinear susceptibility tensors [36, 37, 38]. This technique has been used for Sr2IrO4 and Sr3Ir2O7 to detect subtle structure distortions [27, 28, 21]. Figure 3 shows the azimuth-angle dependence of the third harmonic signals as the scattering plane is rotated about the cc-axis, for the four different polarization configurations of the incident and reflected light, which can be either parallel (PP) or perpendicular (SS) to the scattering plane. The patterns are symmetric with respect to mirror reflections aa\rightarrowa-a and bb\rightarrowb-b, and four-fold rotations about the cc-axis. The combination of both symmetries leads to eight-fold symmetric patterns for the PSPS and SPSP. To confirm, the patterns are overlaid with the best fit to electric-dipole induced THG tensor for 4/mmmmmm (navy), whose expression is given in Appendix A. We find an excellent agreement and conclude that the superlattice has a higher point group symmetry than Sr2IrO4 (Ref. 27) and Sr3Ir2O7 (Ref. 28), in both of which the patterns manifestly lack the mirror symmetries.

III Magnetic structure

Refer to caption
Figure 4: (color online). Magnetometry of the superlattice. (a) The superlattice magnetic order consistent with our data. (b) Field-cooled M-T curves of the superlattice (navy), Sr2IrO4 (cyan) and Sr3Ir2O7 (orange). (c) M-H hysteresis at 5 K comparing the superlattice (navy) and the bulk Sr2IrO4 (cyan). For a direct comparison, Sr2IrO4 curves are multiplied by the mass proportion (\approx 0.36) of SL in the superlattice. (d) M-T curves measured with fields applied along [100] and [110].
Refer to caption
Figure 5: (color online). Magnetic RXD study on the superlattice. (a) Magnetic (1 0 L) reflections appear at every integer L values, contributed by signals from both SL and DL. (b) (1 0 L) scans measured at every 5 K upon heating from 200 K to 245 K. (1 0 21) reflection dominated by SL disappears around T = 220 K. (c) At 250 K, the intensity modulation along (1 0 L) coincides with DL structure factor squared (dotted line), indicating that the magnetic intensities are dominated by DL. (d) Temperature dependence of (1 0 8) and (1 0 10) reveal two magnetic transitions at TNAT_{N}^{A} = 220 K and TNBT_{N}^{B} = 280 K. (e) Polarization analysis to separate AF signals from in-plane (blue) and out-of-plane (orange) moments.

Having established the lattice structure of the superlattice, we now turn to its magnetic structure. In short, the magnetic structure remains almost unchanged from the parent systems: SL has abab-plane canted AF structure while DL has cc-axis collinear AF structure, as shown in Fig. 4(a). The net ferromagnetic response to the dc field with the saturation moment close to one-third of that of Sr2IrO4 [Figs. 4(b) and 4(c)] suggests that the SL (which makes up one-third of the superlattice in terms of the number of Ir ions) has in-plane canted AF structure, while DL has cc-axis collinear AF structure. We note that the AF ordering in Sr3Ir2O7 is visible as a small jump in the magnetization due to its slight monoclinicity β\beta\sim 90.05  (and thereby canting of the moments [28, 30]), but no such anomaly indicative of DL magnetic ordering is seen in our tetragonal superlattice [Fig. 4(b)]. Unlike in Sr2IrO4, we observe a ferromagnetic hysteresis loop in the M-H curve shown in Fig. 4(c), which implies ferromagnetic staking of SL net moments. Based on our M-H and M-T curves [Figs. 4(c) and 4(d)] measured for fields along [100] and [110] directions, we are not able to identify the magnetic easy axis in the abab-plane, which in Sr2IrO4 is clearly along the aa-axis [24, 23]. This signifies reduced magnetic anisotropy, which will be discussed in more detail later on.

Refer to caption
Figure 6: (color online). Magnetic excitations in the superlattice. (a) RIXS map measured at T =10 K along high symmetry directions indicate that the SL and DL modes follow the dispersions of the parent systems (plotted with markers). (b) Spectra at select 𝐪\bf{q} points. The fitted peaks (solid lines) are compared with those of the parent systems (markers). (c) The superlattice spectrum at the zone corner (π\pi,0) (navy line) is well reproduced by a linear sum of Sr2IrO4 and Sr3Ir2O7 spectra (black line). (d) The two-magnon Raman spectrum (navy dots), measured in the B2gB_{\textrm{2g}} channel at T = 15 K, is also well approximated by summing the Sr2IrO4 and Sr3Ir2O7 spectral intensities (black line).

We confirm the magnetic structure shown in Fig. 4(a) using RXD. As in the case of Sr2IrO4 and Sr3Ir2O7, magnetic reflections are found along (1 0 L) [Fig. 5(a)], but at every integer L, which has contributions from both SL and DL. However, they can be separated by exploiting the DL structure factor. The ratio of Ir-O-Ir bond length along the cc-axis to cc lattice parameter returns oscillation period of \sim 4.15. For example, the DL contribution nearly vanishes for (1 0 8) and (1 0 21). Indeed, L scans shown in Fig. 5(b) shows that (1 0 21) peak disappears around T = 220 K as temperature increases while (1 0 22) peak is present up to T = 250 K. At this temperature, the intensity modulation well agrees with the DL structure factor squared [Fig. 5(c)], implying that the peaks are due to reflections from DL only. This is unambiguous evidence for coherent superlattice formation over the probing depth of x-ray (290 nm \sim 3.1 μ\mum as calculated in Ref. 39). The SL transition temperature is measured to be TNAT_{N}^{A} = 220 K from the temperature dependence of (1 0 8) peak shown in Fig. 5(d). At (1 0 10), two transitions are seen, the higher temperature one at TNBT_{N}^{B} = 280 K being the transition in DL.

Additional measurements were conducted using polarization analysis in order to separate the abab-plane and cc-axis components of the antiferromagnetic moments. By studying the magnetic (0 5 2) reflection in a horizontal scattering geometry [Fig. 5(e)], the π\pi-σ\sigma^{\prime} channel mostly detects in-plane moments, whereas π\pi-π\pi^{\prime} is sensitive to out-of-plane moments. The temperature dependence of the integrated intensities in the two channels, shown in Fig. 5(e), reveals that the out-of-plane (in-plane) magnetic signal arises below TNBT_{N}^{B} = 280 K (TNAT_{N}^{A} = 220 K), thereby located at DL (SL), consistent with the magnetic structure in Fig. 4(a).

IV Pseudospin dynamics

Having established the static magnetic structure, we now turn to the pseudospin dynamics. Figure 6(a) plots the pseudospin excitation spectrum along high symmetry directions. Select 𝐪\mathbf{q}-points are plotted in Fig. 6(b). The spectra are fitted with two peaks and their energy positions are compared with the spin-wave dispersions for the parent systems. Overall, the spectra are well described as having two peaks corresponding to spin waves in SL and DL. It is known that Sr2IrO4 is almost gapless at the magnetic zone center reflecting the quasi-continuous symmetry in the abab-plane [16], whereas Sr3Ir2O7 has an exceptionally large gap of \approx 90 meV (Ref. 40). Except in the immediate vicinity of the magnetic zone center, we find a good agreement with the parent systems. In particular, the spectra at the zone corner (π\pi, 0) is well reproduced by a linear sum of the spectra of Sr2IrO4 and Sr3Ir2O7 at the same momenta, as shown in Fig. 6(c). Further, the two-magnon spectrum measured by Raman scattering [Fig. 6(d)], which is dominated by zone-boundary modes, is in perfect agreement with a linear sum of those in Sr2IrO4 and Sr3Ir2O7. Together, these data indicate that the nearest-neighbor spin exchange couplings remain unaltered with respect to the parent systems.

Refer to caption
Figure 7: (color online). Pseudospin-lattice coupling estimated by Raman spectroscopy. (a) The magnetic mode at the zone center is observed in B2gB_{\textrm{2g}} scattering geometry (superlattice, navy; Sr2IrO4, blue gray). (b) The temperature dependent component of the spin-wave gap Δ(T)\Delta(T) extracted from Raman spectra in (a). The trend of the superlattice and Sr2IrO4 share the same functional form A1T/TN+BA\sqrt{1-T/T_{N}}+B, where AA is the offset in the log-log plot and proportional to the strength of PLC. Temperature dependence of A1gA_{\textrm{1g}} phonons in (c) Sr3Ir2O7 and (d) the superlattice. (e) Integrated intensity of the lower energy A1gA_{\textrm{1g}} phonon. The spin-dependent component scales with the ordered moment squared MAF2M_{AF}^{2}, and dashed lines are fits to functional form of 1(T/TN)1-(T/T_{N}), whose slopes are proportional to the PLC strength Λ\Lambda. All Raman spectra are corrected for laser heating and Bose factors.

It is perhaps not surprising that there is no significant change in the pseudospin dynamics across the majority of the Brillouin zone for both SL and DL, considering that these are quasi-2D systems and only the stacking sequence is altered. However, it is rather unusual that the two magnetic subsystems behave almost independently of each other. In particular, magnetic ordering in DL occurs across SL that remains paramagnetic down to TNAT_{N}^{A}. That no other static order exists in SL above TNAT_{N}^{A} is confirmed from its almost identical RIXS spectra measured across TNAT_{N}^{A} (Fig. S4). Our representation analysis (Appendix B) shows that the SL and the DL magnetic orders belong to different irreducible representations, which guarantees that they are not coupled in the linear order (this, however, does not rule out interactions between SL and DL). Further, we note that both TNAT_{N}^{A} and TNBT_{N}^{B} are reduced only by few degrees from their bulk counterparts, which confirms that interlayer couplings play a minor role in the critical phenomena [8, 41, 42].

While the magnetic properties of the superlattice mostly inherit from the parent Sr2IrO4 and Sr3Ir2O7, some differences are also found. Unlike in the case of Sr2IrO4, the in-plane magnetic anistropy of SL is hardly visible in the magnetometry data [Fig. 4(c)]. In Sr2IrO4, the magnetic anisotropy is known to arise mostly from PLC [23, 24] stabilizing the magnetic ordering along the aa- or bb-axis. The PLC dominates over the nearest-neighbor pseudospin exchange anisotropies favoring alignment along the bond directions (i.e. along [110]), and also accounts for a dominant part of the in-plane magnon gap [24]. To compare the strength of PLC, we analyze the temperature dependence of Raman spectra in B2gB_{\textrm{2g}} symmetry channel, which measures the magnon gap at the zone center [see Fig. 7(a)]. It is seen in Fig. 7(b) that the magnon gap of both Sr2IrO4 and the superlattice follow the same mean-field-like functional form of A1T/TN+BA\sqrt{1-T/T_{N}}+B, where the temperature-independent constant BB arises from anisotropic magnetic interactions, and AA measures the strength of PLC that scales with the size of the ordered moment [23, 24, 25]. From the intercept at T = 0 in the log-log plot shown in Fig. 7(b), we find that the PLC is drastically reduced by a factor of two. The reduced PLC in SL can be understood in terms of the structural rigidity provided by DL resisting orthorhombic distortions, which would be energetically unfavorable for the cc-axis collinear AF structure.

An indication of suppressed PLC can also be found from a phonon mode strongly coupled to the AF order in DL. In Sr3Ir2O7, a recent ultrafast optical spectroscopy study has shown that strong PLC manifests as an abrupt large enhancement across TNT_{N} of the amplitude of the coherent oscillation of the impulsive 4.4 THz (\approx 18 meV) A1gA_{\textrm{1g}} phonon mode [25]. The intensity follows the phenomenological relation I(T)Z(Λ/4μB2)MAF2\sqrt{I(T)}\propto Z-(\Lambda/4\mu_{B}^{2})M_{AF}^{2}, where Λ\Lambda and ZZ are the PLC strength and temperature-independent spectral weight, respectively [43]. The strength of PLC in Sr3Ir2O7 is estimated to be two orders of magnitude stronger than in cubic perovskite manganites with strong spin-lattice couplings [44]. Since the oscillation amplitude is linearly dependent on Raman tensor α/Q\partial\alpha/\partial Q [45], where α\alpha is the polarizability and QQ is the normal coordinate of the phonon mode, the enhancement should be also visible in Raman spectra. Indeed, we observe a strong enhancement upon cooling of the 18 meV A1gA_{\textrm{1g}} mode intensity as shown in Fig. 7(c). However, the corresponding mode in the superlattice at \approx 15 meV shows only a modest increase comparable to that of the 23 meV mode [Fig. 7(e)]. Taken as a whole, the absence of discernible anisotropy in the magnetization data, the reduced magnon gap, and the insensitivty of the A1gA_{\textrm{1g}} mode to the magnetic order all consistently point to significant reduction of PLC in the superlattice.

V Summary

Recent advances in the control and understanding of 2D materials has recently expanded into the research of magnetism in the extreme quantum 2D limit and in artificial heterostructures. In this work, we have demonstrated a successful growth of a Sr2IrO4/Sr3Ir2O7 magnetic superlattice, whose constituent bulk phases exhibit novel quantum magnetism in the limit of strong spin-orbit coupling. While intergrowth of different phases in a RP series is frequently found, a natural formation of an ordered superlattice is extremely rare. The superlattice has a lattice symmetry higher than those of the bulk phases and realize an undistorted square lattice geometry. Thus, the superlattice offers a unique opportunity to study the pseudospin physics in an ideal setting.

The superlattice preserves the bulk magnetic orders and interleaves abab-plane canted AF and cc-axis collinear AF alternately stacked along the cc-axis. The two mutually orthogonal orders, however, behave independently of each order reflecting weak interlayer couplings expected for quasi-2D systems. In particular, there is a temperature range where only one magnetic subsystem develops an order across the other that remains paramagnetic. Further, the pseudsospin dynamics remains unchanged from the parent systems for the most part of the Brillouin zone.

Instead, the incompatible nature of the magnetic orders manifests as a strong suppression of the PLC, which is expected to be generally strong for iridates. The reduced PLC in SL is inferred from the smaller zone-center magnon gap, which in Sr2IrO4 is largely accounted for by the PLC through a pseudo-Jahn-Teller effect that results in an orthorhombic distortion as the abab-plane magnetic order breaks the tetragonal symmetry of the lattice. This effect, however, is counteracted in the superlattice by DL. The magnetic order in DL is collinear along the straight cc-axis bond, which in Sr3Ir2O7 is bent due to a slight monoclinicity.

The origin of the distortions in the parent compounds and their absence in the superlattice is a subject of future research. To the best of our knowledge, the breaking of glide symmetries as seen in Sr2IrO4, attributed to staggered tetragonal distortions, is not found in any other transition-metal compounds of the RP type, which suggests that the distortion is not due to an instability of the lattice, but rather its interactions with pseudospins whose magnetic moment size strongly depends on lattice distortions. Similarly, it remains to be understood if PLC plays any role in stablizing the monoclinic structure in Sr3Ir2O7. At any rate, many iridates in bulk form exhibit lattice distortions of some sort and deviate from idealized models, and in this regard the superlattice stands out as a realization of pseudospin one-half physics on an undistorted square lattice.

The persistent magnetic orders in the superlattice also allows investigation of their evolution upon carrier doping. It is known that Sr3Ir2O7 is on the verge of a metal-insulator transition, and therefore it may well be that DL turns into a Fermi liquid metal while SL remains a correlated fluid, a situation akin to the case of electron-doped Sr2IrO4 via surface alkali metal dosing, where Fermi arcs and a dd-wave gap are observed, which possibly involve screening effects from the surface metallic layer. Our results presented here provide a solid foundation for these future investigations.

VI Methods

STEM measurements were conducted using JEMARM200F, JEOL at 200 kV with 5th-order probe-corrector (ASCOR, CEOS GmbH, Germany). The specimens were prepared by dual-beam focused ion beam (FIB) with Ga ion beams of 1 \sim 5 kV (Nanolab G3 CX, FEI), and further milled by Ar ion beams of 100 meV (PIPS II, Gatan) to minimize the surface damage. RXD and RIXS measurements were carried out at Beamline 6-ID-B and Beamline 27-ID, respectively, at the Advanced Photon Source, Argonne National Laboratory. For RA-THG measurements, we adapted the scheme reported in [37] by replacing the phase mask by a pair of wedge prisms to accommodate a wider wavelength range of incident light (from 800 nm to 1200 nm). The incident 1200 nm light was provided by an optical parametric amplifier (Orpheus-F, Light Conversion) powered by a femtosecond laser of 100 kHz repetition rate (Pharos, Light Conversion). Raman spectroscopy is measured with 633 nm He-Ne laser, which beam of 0.8 mW is focused on \sim2 μ\mum.

Acknowledgement

This project is supported by IBS-R014-A2 and National Research Foundation (NRF) of Korea through the SRC (no. 2018R1A5A6075964). We acknowledge financial support by the European Research Council under Advanced Grant No. 669550 (Com4Com). The use of the Advanced Photon Source at the Argonne National Laboratory was supported by the U. S. DOE under Contract No. DE-AC02-06CH11357. We acknowledge DESY (Hamburg, Germany), a member of the Helmholtz Association HGF, for the provision of experimental facilities. J. B. was supported by the Alexander von Humboldt Foundation. G. Noh, G.-Y. Kim, and S.-Y. Choi acknowledge the support of the Global Frontier Hybrid Interface Materials by the NRF (Grant No. 2013M3A6B1078872).

APPENDIX A: Third Harmonic Generation Tensor

We obtain the general form of the bulk electric dipole induced THG tensor for the 4/mmmmmm point group by starting from the most general form of a fourth-rank Cartesian tensor and requiring it to be invariant under all of the symmetry operations contained in the group as:

χijklED=((xxxx000xxyy000xxzz)(0xxyy0xxyy00000)(00xxzz000xxzz0)(0xxyy0xxyy00000)(xxyy000xxxx000xxzz)(00000xxzz0xxzz0)(00zxxz000zxxz00)(00000zxxz0zxxz0)(zxxz000zxxz000zzzz)),\chi^{ED}_{ijkl}=\begin{pmatrix}\begin{pmatrix}xxxx&0&0\\ 0&xxyy&0\\ 0&0&xxzz\end{pmatrix}&\begin{pmatrix}0&xxyy&0\\ xxyy&0&0\\ 0&0&0\end{pmatrix}&\begin{pmatrix}0&0&xxzz\\ 0&0&0\\ xxzz&0&\end{pmatrix}\\ \begin{pmatrix}0&xxyy&0\\ xxyy&0&0\\ 0&0&0\end{pmatrix}&\begin{pmatrix}xxyy&0&0\\ 0&xxxx&0\\ 0&0&xxzz\end{pmatrix}&\begin{pmatrix}0&0&0\\ 0&0&xxzz\\ 0&xxzz&0\end{pmatrix}\\ \begin{pmatrix}0&0&zxxz\\ 0&0&0\\ zxxz&0&0\end{pmatrix}&\begin{pmatrix}0&0&0\\ 0&0&zxxz\\ 0&zxxz&0\end{pmatrix}&\begin{pmatrix}zxxz&0&0\\ 0&zxxz&0\\ 0&0&zzzz\end{pmatrix}\par\end{pmatrix}\;, (1)

where jkljkl(ii) stands for incident(scattered) light polarizations. In our experiment, the light is incident on the sample surface normal to the cc-axis with the incidence angle θ\theta and azimuth angle ψ\psi, which is defined to be zero when the scattering plane contains the aa and cc axes). Then, the polarization vectors are given by

ϵs\displaystyle\vec{\epsilon}_{s}\, =(sinψ,cosψ, 0),\displaystyle=\,(\sin\psi,\,-\cos\psi,\,0)\;,
ϵin,p\displaystyle\vec{\epsilon}_{in,p}\, =(cosθcosψ,cosθsinψ,sinθ),\displaystyle=\,(-\cos\theta\cos\psi,\,-\cos\theta\sin\psi,\,\sin\theta)\;, (A2)
ϵout,p\displaystyle\vec{\epsilon}_{out,p}\, =(cosθcosψ,cosθsinψ,sinθ),\displaystyle=\,(\cos\theta\cos\psi,\,\cos\theta\sin\psi,\,\sin\theta)\;,

Multiplying the tensor with the polarization vectors, the expressions for the THG intensity simplify as

ISS(3ω)\displaystyle I^{SS}(3\omega)\,\sim\, |A+Bcos(4ψ)|2,\displaystyle|A+B\cos(4\psi)|^{2}\;,
IPS(3ω)\displaystyle I^{PS}(3\omega)\,\sim\, |Bsin(4ψ)|2,\displaystyle|B\sin(4\psi)|^{2}\;, (A3)
ISP(3ω)\displaystyle I^{SP}(3\omega)\,\sim\, |Bsin(4ψ)|2,\displaystyle|B\sin(4\psi)|^{2}\;,
IPP(3ω)\displaystyle I^{PP}(3\omega)\,\sim\, |A+Bcos(4ψ)|2,\displaystyle|A^{\prime}+B\cos(4\psi)|^{2}\;,

where AA, AA^{\prime} and BB are adjustable parameters. The formulae are consistent with the 4/mmmmmm point group symmetry, invariant under both mirror reflections (ψ\psi \rightarrow πψ\pi-\psi and ψ\psi \rightarrow ψ-\psi) and C4 rotation about the cc-axis (ψ\psi \rightarrow π/2+ψ\pi/2+\psi). The full expressions and those for lower symmetry point groups are given in the supplemental materials.

Irrep. Single layer Double layer
a1 a2 b1 b2 b3 b4
Γ1\Gamma_{1} (m1, m2, 0) (-m1, m2, 0) (m3, m4, 0) (m5, m6, 0) (-m3, m4, 0) (-m5, m6, 0)
(-m2,  m1, 0) (-m2, -m1, 0) (-m6, -m5, 0) (-m4, -m3, 0) (-m6, m5, 0) (-m4, m3, 0)
Γ2\Gamma_{2} (0, 0, m1) (0, 0, -m1) (0, 0, m1) (0, 0, -m1)
Γ3\Gamma_{3} (0, 0, m1) (0, 0, m1) (0, 0, -m1) (0, 0, -m1)
Γ4\Gamma_{4} (0, 0, m1) (0, 0, m1) (0, 0, m2) (0, 0, m2) (0, 0, m2) (0, 0, m2)
Γ5\Gamma_{5} (0, 0, m1) (0, 0, -m1) (0, 0, -m2) (0, 0, m2) (0, 0, m2) (0, 0, -m2)
Table 1: Irreducible representations and basis vectors for the magnetic structures allowed in the superlattice. mi (i=1,,6=1,\cdots,6) are independent parameters for the magnetic moments. Iridium ions are located at a1:(0,0,0), a2:(1/2,1/2,0), b1:(0,1/2,δ\delta), b2:(0,1/2,1-δ\delta), b3:(1/2,0,δ\delta), b4:(1/2,0,1-δ\delta) in the unit cell Fig. 1(c), where δ\delta = 0.3791. The magnetic structure in Fig. 4(a) comprises canted abab-plane AF of Γ1\Gamma_{1} and cc-axis collinear AF of Γ5\Gamma_{5}.

APPENDIX B: Representation Analysis of the magnetic structure

We present a representation analysis based on the lattice structure shown in Fig. 1(c), assuming 𝐪\mathbf{q} = 0 propagation vector based on the fact that all observed magnetic reflections have integer Miller indices. Its space group is P4¯bP\bar{4}b2, which lacks inversion symmetry and thus its point group (4¯m2\bar{4}m2) is of lower symmetry than 4/mmmmmm. RA-THG is, however, insensitive to the inversion symmetry and has the same tensor structure for the two point groups. The inversion symmetry is broken by the way in which octahdera are rotated in DL in the current model. An inversion symmetric structure model can be made by doubling the unit cell in such a way that the octahedral rotation is opposite on two neighboring DLs. Thus, the determination of the presence of inversion symmetry requires full structure refinement including the octahedral rotation on each layers, which is beyond the scope of this work. However, our result that the magnetic orders in SL and DL belong to different irreducible representations (IR) is not affected by these details.

In the superlattice, iridium ions in SL and DL are not symmetry related and thus their magnetic structures are analyzed separately. The result of the analysis is shown in Table 1. In both SL and DL, the abab-plane and cc-axis components belong to different irreducible representations. This can easily be seen from the transformation property under the two-fold rotation about cc-axis; the abab-plane moments are flipped, whereas the cc-axis moments remain invariant. As long as the abab-plane and cc-axis components are not mixed by any of the symmetry operations contained in the space group, their different characters for the two-fold rotation guarantee that they belong to different IRs. A more detailed version of the analysis is presented in the supplemental material.

References

  • Anderson [1987] P. W. Anderson, The Resonating Valence Bond State in La2CuO4\mathrm{La}_{2}\mathrm{CuO}_{4} and SuperconductivityScience 235, 1196–1198 (1987).
  • Lee et al. [2006] P. A. Lee, N. Nagaosa, and X.-G. Wen, Doping a Mott insulator: Physics of high-temperature superconductivityRev. Mod. Phys. 78, 17–85 (2006).
  • Keimer et al. [2015] B. Keimer, S. A. Kivelson, M. R. Norman, S. Uchida, and J. Zaanen, From quantum matter to high-temperature superconductivity in copper oxidesNature 518, 179–186 (2015).
  • Plakida [2010] N. Plakida, High-Temperature Cuprate Superconductors (Springer, 2010).
  • Orendáčová et al. [2019] A. Orendáčová, R. Tarasenko, V. Tkáč, E. Čižmár, M. Orendáč, and A. Feher, Interplay of Spin and Spatial Anisotropy in Low-Dimensional Quantum Magnets with Spin 1/2Crystals 9, 6 (2019).
  • Bertinshaw et al. [2019] J. Bertinshaw, Y. K. Kim, G. Khaliullin, and B. J. Kim, Square Lattice IridatesAnnu. Rev. Condens. Matter Phys. 10, 315–336 (2019).
  • Kim et al. [2008] B. J. Kim, H. Jin, S. J. Moon, J.-Y. Kim, B.-G. Park, C. S. Leem, J. Yu, T. W. Noh, C. Kim, S.-J. Oh, J.-H. Park, V. Durairaj, G. Cao, and E. Rotenberg, Novel Jeff=1/2{J}_{\mathrm{eff}}=1/2 Mott State Induced by Relativistic Spin-Orbit Coupling in Sr2IrO4{\mathrm{Sr}}_{2}{\mathrm{IrO}}_{4}Phys. Rev. Lett. 101, 076402 (2008).
  • Jackeli and Khaliullin [2009] G. Jackeli and G. Khaliullin, Mott Insulators in the Strong Spin-Orbit Coupling Limit: From Heisenberg to a Quantum Compass and Kitaev ModelsPhys. Rev. Lett. 102, 017205 (2009).
  • Kim et al. [2009] B. J. Kim, H. Ohsumi, T. Komesu, S. Sakai, T. Morita, H. Takagi, and T. Arima, Phase-Sensitive Observation of a Spin-Orbital Mott State in Sr2IrO4{\mathrm{Sr}}_{2}{\mathrm{IrO}}_{4}Science 323, 1329–1332 (2009).
  • Wang and Senthil [2011] F. Wang and T. Senthil, Twisted Hubbard Model for Sr2IrO4{\mathrm{Sr}}_{2}{\mathrm{IrO}}_{4}: Magnetism and Possible High Temperature SuperconductivityPhys. Rev. Lett. 106, 136402 (2011).
  • Kim et al. [2014] Y. K. Kim, O. Krupin, J. D. Denlinger, A. Bostwick, E. Rotenberg, Q. Zhao, J. F. Mitchell, J. W. Allen, and B. J. Kim, Fermi arcs in a doped pseudospin-1/2 Heisenberg antiferromagnetScience 345, 187–190 (2014).
  • Yan et al. [2015] Y. J. Yan, M. Q. Ren, H. C. Xu, B. P. Xie, R. Tao, H. Y. Choi, N. Lee, Y. J. Choi, T. Zhang, and D. L. Feng, Electron-Doped Sr2IrO4{\mathrm{Sr}}_{2}{\mathrm{IrO}}_{4}: An Analogue of Hole-Doped Cuprate Superconductors Demonstrated by Scanning Tunneling MicroscopyPhys. Rev. X 5, 041018 (2015).
  • Cao et al. [2016] Y. Cao, Q. Wang, J. A. Waugh, T. J. Reber, H. Li, X. Zhou, S. Parham, S. R. Park, N. C. Plumb, E. Rotenberg, A. Bostwick, J. D. Denlinger, T. Qi, M. A. Hermele, G. Cao, and D. S. Dessau, Hallmarks of the Mott-metal crossover in the hole-doped pseudospin-1/2 Mott insulator Sr2IrO4{\mathrm{Sr}}_{2}{\mathrm{IrO}}_{4}Nat. Commun. 7, 11367 (2016).
  • Kim et al. [2016] Y. K. Kim, N. H. Sung, J. D. Denlinger, and B. J. Kim, Observation of a d-wave gap in electron-doped Sr2IrO4{\mathrm{Sr}}_{2}{\mathrm{IrO}}_{4}Nat. Phys. 12, 37–41 (2016).
  • Sumita et al. [2017] S. Sumita, T. Nomoto, and Y. Yanase, Multipole Superconductivity in Nonsymmorphic Sr2IrO4{\mathrm{Sr}}_{2}{\mathrm{IrO}}_{4}Phys. Rev. Lett. 119, 027001 (2017).
  • Kim et al. [2012a] J. Kim, D. Casa, M. H. Upton, T. Gog, Y.-J. Kim, J. F. Mitchell, M. van Veenendaal, M. Daghofer, J. van den Brink, G. Khaliullin, and B. J. Kim, Magnetic Excitation Spectra of Sr2IrO4{\mathrm{Sr}}_{2}{\mathrm{IrO}}_{4} Probed by Resonant Inelastic X-Ray Scattering: Establishing Links to Cuprate SuperconductorsPhys. Rev. Lett. 108, 177003 (2012a).
  • Fujiyama et al. [2012] S. Fujiyama, H. Ohsumi, T. Komesu, J. Matsuno, B. J. Kim, M. Takata, T. Arima, and H. Takagi, Two-Dimensional Heisenberg Behavior of Jeff=1/2{J}_{\mathrm{eff}}\mathbf{=}1/2 Isospins in the Paramagnetic State of the Spin-Orbital Mott Insulator Sr2IrO4{\mathrm{Sr}}_{2}{\mathrm{IrO}}_{4}Phys. Rev. Lett. 108, 247212 (2012).
  • Jeong et al. [2017] J. Jeong, Y. Sidis, A. Louat, V. Brouet, and P. Bourges, Time-reversal symmetry breaking hidden order in Sr2(Ir,Rh)O4{\mathrm{Sr}}_{2}{\mathrm{(Ir,Rh)O}}_{4}Nat. Commun. 8, 15119 (2017).
  • Murayama et al. [2021] H. Murayama, K. Ishida, R. Kurihara, T. Ono, Y. Sato, Y. Kasahara, H. Watanabe, Y. Yanase, G. Cao, Y. Mizukami, T. Shibauchi, Y. Matsuda, and S. Kasahara, Bond Directional Anapole Order in a Spin-Orbit Coupled Mott Insulator Sr2(Ir1xRhx)O4{\mathrm{Sr}}_{2}({\mathrm{Ir}}_{1-x}{\mathrm{Rh}}_{x}){\mathrm{O}}_{4}Phys. Rev. X 11, 011021 (2021).
  • Zhao et al. [2016] L. Zhao, D. H. Torchinsky, H. Chu, V. Ivanov, R. Lifshitz, R. Flint, T. Qi, G. Cao, and D. Hsieh, Evidence of an odd-parity hidden order in a spin–orbit coupled correlated iridateNat. Phys. 12, 32–36 (2016).
  • Seyler et al. [2020] K. L. Seyler, A. de la Torre, Z. Porter, E. Zoghlin, R. Polski, M. Nguyen, S. Nadj-Perge, S. D. Wilson, and D. Hsieh, Spin-orbit-enhanced magnetic surface second-harmonic generation in Sr2IrO4{\mathrm{Sr}}_{2}\mathrm{Ir}{\mathrm{O}}_{4}Phys. Rev. B 102, 201113(R) (2020).
  • Chen et al. [2018] X. Chen, J. L. Schmehr, Z. Islam, Z. Porter, E. Zoghlin, K. Finkelstein, J. P. C. Ruff, and S. D. Wilson, Unidirectional spin density wave state in metallic (Sr1xLax)2IrO4({\mathrm{Sr}}_{1-x}{\mathrm{La}}_{x})_{2}{\mathrm{IrO}}_{4}Nat. Commun. 9, 103 (2018).
  • Liu and Khaliullin [2019] H. Liu and G. Khaliullin, Pseudo-Jahn-Teller Effect and Magnetoelastic Coupling in Spin-Orbit Mott InsulatorsPhys. Rev. Lett. 122, 057203 (2019).
  • Porras et al. [2019] J. Porras, J. Bertinshaw, H. Liu, G. Khaliullin, N. H. Sung, J.-W. Kim, S. Francoual, P. Steffens, G. Deng, M. M. Sala, A. Efimenko, A. Said, D. Casa, X. Huang, T. Gog, J. Kim, B. Keimer, and B. J. Kim, Pseudospin-lattice coupling in the spin-orbit Mott insulator Sr2IrO4{\mathrm{Sr}}_{2}{\mathrm{IrO}}_{4}Phys. Rev. B 99, 085125 (2019).
  • Hu et al. [2019] L. L. Hu, M. Yang, Y. L. Wu, Q. Wu, H. Zhao, F. Sun, W. Wang, R. He, S. L. He, H. Zhang, R. J. Huang, L. F. Li, Y. G. Shi, and J. Zhao, Strong pseudospin-lattice coupling in Sr3Ir2O7{\mathrm{Sr}}_{3}{\mathrm{Ir}}_{2}{\mathrm{O}}_{7}: Coherent phonon anomaly and negative thermal expansionPhys. Rev. B 99, 094307 (2019).
  • Ye et al. [2013] F. Ye, S. Chi, B. C. Chakoumakos, J. A. Fernandez-Baca, T. Qi, and G. Cao, Magnetic and crystal structures of Sr2IrO4{\mathrm{Sr}}_{2}{\mathrm{IrO}}_{4}: A neutron diffraction studyPhys. Rev. B 87, 140406(R) (2013).
  • Torchinsky et al. [2015] D. H. Torchinsky, H. Chu, L. Zhao, N. B. Perkins, Y. Sizyuk, T. Qi, G. Cao, and D. Hsieh, Structural Distortion-Induced Magnetoelastic Locking in Sr2IrO4{\mathrm{Sr}}_{2}{\mathrm{IrO}}_{4} Revealed through Nonlinear Optical Harmonic GenerationPhys. Rev. Lett. 114, 096404 (2015).
  • Hogan et al. [2016] T. Hogan, L. Bjaalie, L. Zhao, C. Belvin, X. Wang, C. G. Van de Walle, D. Hsieh, and S. D. Wilson, Structural investigation of the bilayer iridate Sr3Ir2O7{\mathrm{Sr}}_{3}{\mathrm{Ir}}_{2}{\mathrm{O}}_{7}Phys. Rev. B 93, 134110 (2016).
  • Boseggia et al. [2013] S. Boseggia, H. C. Walker, J. Vale, R. Springell, Z. Feng, R. S. Perry, M. M. Sala, H. M. Rønnow, S. P. Collins, and D. F. McMorrow, Locking of iridium magnetic moments to the correlated rotation of oxygen octahedra in Sr2IrO4{\mathrm{Sr}}_{2}{\mathrm{IrO}}_{4} revealed by x-ray resonant scatteringJ. Phys.: Condens. Matter 25, 422202 (2013).
  • Cao et al. [2002] G. Cao, Y. Xin, C. S. Alexander, J. E. Crow, P. Schlottmann, M. K. Crawford, R. L. Harlow, and W. Marshall, Anomalous magnetic and transport behavior in the magnetic insulator Sr3Ir2O7{\mathrm{Sr}}_{3}{\mathrm{Ir}}_{2}{\mathrm{O}}_{7}Phys. Rev. B 66, 214412 (2002).
  • Nagai et al. [2007] I. Nagai, Y. Yoshida, S. I. Ikeda, H. Matsuhata, H. Kito, and M. Kosaka, Canted antiferromagnetic ground state in Sr3Ir2O7{\mathrm{Sr}}_{3}{\mathrm{Ir}}_{2}{\mathrm{O}}_{7}J. Phys.: Condens. Matter 19, 136214 (2007).
  • Chaloupka et al. [2010] J. Chaloupka, G. Jackeli, and G. Khaliullin, Kitaev-Heisenberg Model on a Honeycomb Lattice: Possible Exotic Phases in Iridium Oxides A2IrO3{A}_{2}{\mathrm{IrO}}_{3}Phys. Rev. Lett. 105, 027204 (2010).
  • Crawford et al. [1994] M. K. Crawford, M. A. Subramanian, R. L. Harlow, J. A. Fernandez-Baca, Z. R. Wang, and D. C. Johnston, Structural and Magnetic Studies of Sr2IrO4{\mathrm{Sr}}_{2}{\mathrm{IrO}}_{4}Phys. Rev. B 49, 9198–9201 (1994).
  • Matsuhata et al. [2004] H. Matsuhata, I. Nagai, Y. Yoshida, S. Hara, S.-I. Ikeda, and N. Shirakawa, Crystal structure of sr3ir2o7{\mathrm{sr}}_{3}{\mathrm{ir}}_{2}{\mathrm{o}}_{7} investigated by transmission electron microscopy, J. Solid State Chem. 177, 3776–3783 (2004).
  • Harlow et al. [1995] R. L. Harlow, Z. G. Li, W. J. Marshall, M. K. Crawford, and M. A. Subramanian, Effects of stacking faults on refinement of single crystal X-ray diffraction data for Sr5Ir3O11{\mathrm{Sr}}_{5}{\mathrm{Ir}}_{3}{\mathrm{O}}_{11}Materials research bulletin 30, 217–223 (1995).
  • Torchinsky et al. [2014] D. H. Torchinsky, H. Chu, T. Qi, G. Cao, and D. Hsieh, A low temperature nonlinear optical rotational anisotropy spectrometer for the determination of crystallographic and electronic symmetriesRev. Sci. Instrum. 85, 083102 (2014).
  • Harter et al. [2015] J. W. Harter, L. Niu, A. J. Woss, and D. Hsieh, High-speed measurement of rotational anisotropy nonlinear optical harmonic generation using position-sensitive detectionOpt. Lett. 40, 4671–4674 (2015).
  • Sipe et al. [1987] J. E. Sipe, D. J. Moss, and H. M. van Driel, Phenomenological theory of optical second- and third-harmonic generation from cubic centrosymmetric crystalsPhys. Rev. B 35, 1129–1141 (1987).
  • Mazzone et al. [2021] D. G. Mazzone, D. Meyers, Y. Cao, J. G. Vale, C. D. Dashwood, Y. Shi, A. J. A. James, N. J. Robinson, J. Lin, V. Thampy, Y. Tanaka, A. S. Johnson, H. Miao, R. Wang, T. A. Assefa, J. Kim, D. Casa, R. Mankowsky, D. Zhu, R. Alonso-Mori, S. Song, H. Yavas, T. Katayama, M. Yabashi, Y. Kubota, S. Owada, J. Liu, J. Yang, R. M. Konik, I. K. Robinson, J. P. Hill, D. F. McMorrow, M. Först, S. Wall, X. Liu, and M. P. M. Dean, Laser-induced transient magnons in Sr3Ir2O7{\mathrm{Sr}}_{3}{\mathrm{Ir}}_{2}{\mathrm{O}}_{7} throughout the Brillouin zoneProc. Natl. Acad. Sci. USA 118, e2103696118 (2021).
  • Kim et al. [2012b] J. Kim, A. H. Said, D. Casa, M. H. Upton, T. Gog, M. Daghofer, G. Jackeli, J. van den Brink, G. Khaliullin, and B. J. Kim, Large Spin-Wave Energy Gap in the Bilayer Iridate Sr3Ir2O7{\mathrm{Sr}}_{3}{\mathrm{Ir}}_{2}{\mathrm{O}}_{7}: Evidence for Enhanced Dipolar Interactions Near the Mott Metal-Insulator TransitionPhys. Rev. Lett. 109, 157402 (2012b).
  • Vale et al. [2015] J. G. Vale, S. Boseggia, H. C. Walker, R. Springell, Z. Feng, E. C. Hunter, R. S. Perry, D. Prabhakaran, A. T. Boothroyd, S. P. Collins, H. M. Rønnow, and D. F. McMorrow, Importance of XYXY anisotropy in Sr2IrO4{\mathrm{Sr}}_{2}{\mathrm{IrO}}_{4} revealed by magnetic critical scattering experimentsPhys. Rev. B 92, 020406(R) (2015).
  • Irkhin and Katanin [1999] V. Y. Irkhin and A. A. Katanin, Kosterlitz-Thouless and magnetic transition temperatures in layered magnets with a weak easy-plane anisotropyPhys. Rev. B 60, 2990–2993 (1999).
  • Suzuki and Kamimura [1973] N. Suzuki and H. Kamimura, Theory of Spin-Dependent Phonon Raman Scattering in Magnetic CrystalsJ. Phys. Soc. Jpn. 35, 985–995 (1973).
  • Granado et al. [1999] E. Granado, A. García, J. A. Sanjurjo, C. Rettori, I. Torriani, F. Prado, R. D. Sánchez, A. Caneiro, and S. B. Oseroff, Magnetic ordering effects in the Raman spectra of La1xMn1xO3{\mathrm{La}}_{1-x}{\mathrm{Mn}}_{1-x}{\mathrm{O}}_{3}Phys. Rev. B 60, 11879–11882 (1999).
  • Zeiger et al. [1992] H. J. Zeiger, J. Vidal, T. K. Cheng, E. P. Ippen, G. Dresselhaus, and M. S. Dresselhaus, Theory for displacive excitation of coherent phononsPhys. Rev. B 45, 768–778 (1992).