1The Institute for Solid State Physics, the University of Tokyo, Kashiwa, Chiba 277-8581, Japan2RIKEN Center for Emergent Matter Science (CEMS), Saitama 351-0198, Japan3J-PARC Center, Japan Atomic Energy Agency, Tokai, Ibaraki 319-1195, Japan4Department of Physics, Advanced Sciences, G.S.H.S. Ochanomizu University, Bunkyo, Tokyo 112-8610, Japan5Institute of Materials Structure Science, High Energy Accelerator Research Organization, Tsukuba, Ibaraki 305-0801, Japan6Trans-scale Quantum Science Institute, The University of Tokyo, Bunkyo, Tokyo 113-0033, Japan
Received October 10, 2023; Accepted March 28, 2024; Published May 28, 2024
Neutron scattering is a powerful tool to study magnetic structures and cross-correlated phenomena originated from symmetry of the magnetic structures in matter. Among a number of neutron scattering techniques, polarized neutron scattering is quite sensitive to the orientations of the magnetic moments, which are essential to understand microscopic mechanisms of the spin-driven emergent phenomena. Here, we present POlarized Neutron Triple-Axis spectrometer PONTA in the Japan Research Reactor 3 (JRR-3), and show the present capabilities of polarized and unpolarized neutron scattering by introducing recent results from the instrument.
This article is published by the Physical Society of Japan under the terms of the Creative Commons Attribution 4.0 License. Any further distribution of this work must maintain attribution to the author(s) and the title of the article, journal citation, and DOI.
Emergent phenomena induced by nontrivial magnetic structures have been extensively studied in condensed matter physics. One of the most well-known examples is the spin-driven ferroelectricity of TbMnO3,1) in which the cycloidal magnetic order breaks the inversion symmetry of the crystal and accounts for the ferroelectricity.2) The helical magnetic orders in itinerant systems have also attracted attention since the discovery of magnetic skyrmion lattice,3) which is described by a superposition of multiple helimagnetic modulations running along different directions. The skyrmion lattice is composed of vortex-like spin objects, which induce emergent magnetic fields acting only on the conduction electrons; this effect is known as topological Hall effect.4) There are also a variety of emergent phenomena originated from nontrivial magnetic orders.5,6) To establish the microscopic couplings between the spins and other degrees of freedom in matter, it is indispensable to experimentally determine the magnetic structures.
Neutron scattering is one of the best tools to determine the magnetic structures, because it can directly probe Fourier-transformed magnetic moments in the samples. The most straightforward way to determine the magnetic structure by neutron scattering is to collect intensities of magnetic Bragg reflections in the unpolarized condition. By comparing the observed intensities with calculations, we can determine the orientations and magnitudes of the magnetic moments in the sample. A complementary technique is polarized neutron scattering. By measuring the spin polarization of the scattered neutrons, we can precisely determine the orientations of the magnetic moments. In this article, we introduce the POlarized Neutron Triple-Axis spectrometer PONTA installed in the Japan Research Reactor 3 (JRR-3). This instrument is utilized for both the polarized and unpolarized neutron scattering, and has been used for studies of magnetic materials exhibiting the emergent phenomena mentioned above. We review the present capabilities of PONTA and describe possible upgrades in the future.
2. Instrument Details
PONTA is a conventional triple-axis neutron spectrometer with longitudinal polarization analysis option. Figure 1 shows the instrument layout, which is normally used for both the unpolarized and polarized experiments. A unpolarized polychromatic neutron beam from the reactor was monochromatized by a pyrolytic graphite (PG) monochromator or a Cu2MnAl Heusler crystal monochromator. The former and latter produce unpolarized and polarized neutron beams, respectively. These monochromators are installed in a vertical translation stage, which enables us to switch the incident neutron beam between unpolarized and polarized conditions during an experiment. The incident neutron beam monochromatized by the crystal monochromators contains higher order neutrons with the wavelengths of \(\lambda/2\) and \(\lambda/3\), where λ is the primally wavelength reflected by the (002) and (111) planes of PG and Heusler monochromators, respectively. To minimize the higher-order neutrons, we employ a PG filter, which is put before (after) the sample position in elastic (inelastic) experiments. In addition, we put a sapphire filter, which eliminates neutrons in the epithermal regime (\(\lambda < 0.4\) Å), between the monochromator and the sample. The incident neutrons are scattered by the sample put on the goniometer. To analyze the energy of the scattered neutrons, we employ a PG(002) or Heusler(111) analyzer; the former and latter are used for unpolarized and polarized measurements, similarly to the monochromators. In addition to these fundamental elements of the spectrometer, there are also equipments only for polarized measurements, specifically a spin flipper and a Helmholtz coil. The spin flipper is installed between the monochromator and the sample stage, and has a thin rectangular-shaped coil to apply a horizontal magnetic field. The polarized incident neutrons exhibit precession in the coil. By adjusting the magnitude of the horizontal magnetic field, the spins of the incident neutrons are changed from up state to down state. The Helmholtz coil consists of a pair of vertical field coils and three horizontal field coils. In polarized experiments, the direction of the neutron spins is controlled by the magnetic fields applied by these coils. As we explain in the following sections, we mainly use two different configurations regarding the spin directions of the neutrons.
Figure 1. (Color online) Schematic illustration showing the typical instrument layout and the definitions of the Cartesian coordinates xyz with respect to the wavevectors of the incident and scattered neutrons (\(\boldsymbol{k}\) and \(\boldsymbol{k}'\)) and the scattering vector \({\boldsymbol{\kappa}}\).
As for the sample environment, we normally use a closed-cycle 4He refrigerator, which can reach the lowest temperature of 2.2 K, in zero magnetic field. PONTA can also be operated with most of the equipment for low-temperature and in-field measurements available in the Neutron Science Laboratory of the Institute for Solid State Physics of University of Tokyo (ISSP-NSL). For instance, we have performed polarized and unpolarized neutron scattering measurements with 3He refrigerator with the lowest temperature of 0.3 K, high-temperature closed-cycle 4He refrigerator with accessible temperature range of 10–700 K, 1.1 T-electromagnet, and 6 T vertical-field superconducting cryomagnet. The available equipment is listed the website of ISSP-NSL.7)
3. Polarized Neutron Scattering
Preliminary details
We briefly recall the basic formulas of polarized neutron scattering from magnetic materials. The elastic scattering cross section of neutrons \(d\sigma/d\Omega|_{\textit{el}}\) is described as follows:8) \begin{equation} \frac{d\sigma}{d\Omega}\bigg|_{\textit{el}} \propto |\langle \boldsymbol{k}' s'|V|\boldsymbol{k}s\rangle|^{2}. \end{equation} (1) \(\boldsymbol{k}\) and \(\boldsymbol{k}'\) are wave vectors of the incident and scattered neutrons. s and \(s'\) are the initial and final spin state of the neutrons. V represents the interaction potential for a neutron during the scattering process. For magnetic scattering, V is given by \begin{equation} V = - {\boldsymbol{\mu}}_{n}\cdot\boldsymbol{B}, \end{equation} (2) where \({\boldsymbol{\mu}}_{n}\) and \(\boldsymbol{B}\) show the magnetic dipole moment of a neutron and the magnetic field from the magnetic moments in the sample, respectively. By executing the integral with respect to the space coordinates, the cross section is written in the form of \begin{equation} \frac{d\sigma}{d\Omega}\bigg|_{\textit{el}} \propto |\langle s'|{\boldsymbol{\sigma}}\cdot\boldsymbol{M}^{\bot}({\boldsymbol{\kappa}})| s\rangle|^{2}. \end{equation} (3) \({\boldsymbol{\sigma}}\) is the Pauli spin operator for the neutron spin. \(\boldsymbol{M}^{\bot}({\boldsymbol{\kappa}})\) is a square of Fourier-transformed magnetic moments projected onto a plane perpendicular to the scattering vector \({\boldsymbol{\kappa}}\) (\(\equiv\boldsymbol{k}-\boldsymbol{k}'\)). To describe the polarization direction of the neutrons, we introduce a Cartesian coordinate system xyz in which the x axis is parallel to \({\boldsymbol{\kappa}}\), the z axis is perpendicular to the scattering plane, and the y axis completes the right-hand set (see the inset of Fig. 1). By using this coordinate system, \(\boldsymbol{M}^{\bot}({\boldsymbol{\kappa}})\) is written as \(\boldsymbol{M}^{\bot}({\boldsymbol{\kappa}})=(0,M^{\bot}_{y}({\boldsymbol{\kappa}}),M^{\bot}_{z}({\boldsymbol{\kappa}}))\). Note that the x component is always absent, because only the magnetic moments projected onto the yz plane contribute to \(\boldsymbol{M}^{\bot}({\boldsymbol{\kappa}})\). Taking the z axis as the quantization axis of the neutron spins, the right side of Eq. (3) is described as follows: \begin{align} |\langle{\uparrow}|\sigma_{y} M^{\bot}_{y}({\boldsymbol{\kappa}})+\sigma_{z} M^{\bot}_{z}({\boldsymbol{\kappa}})|{\uparrow}\rangle|^{2} &= |M^{\bot}_{z}({\boldsymbol{\kappa}})|^{2}, \end{align} (4) \begin{align} |\langle{\uparrow}|\sigma_{y} M^{\bot}_{y}({\boldsymbol{\kappa}})+\sigma_{z} M^{\bot}_{z}({\boldsymbol{\kappa}})|{\downarrow}\rangle|^{2} &= |M^{\bot}_{y}({\boldsymbol{\kappa}})|^{2}. \end{align} (5) This means that the magnetic moments perpendicular to the neutron spins induce spin-flip (SF) scattering, while those parallel to the neutron spin account for non-spin-flip (NSF) scattering. This experimental configuration is referred to as \(P_{zz}\) configuration. To realize this experimental configuration in the instrument, we apply weak guiding fields along the z axis throughout the scattering path. The incident neutrons obtained by the Heusler monochromator are in the ↑ state, which can be flipped to ↓ state by the spin flipper. Finally, only the neutrons in the ↑ state are reflected by the Heusler(111) analyzer and observed at the detector. Therefore, the intensities measured when the spin flipper is on (\(I_{\text{ON}}\)) and off (\(I_{\text{OFF}}\)) correspond to the SF and NSF scatterings, respectively.
By measuring the SF and NSF intensities, we can distinguish magnetic structures having different orientations of the magnetic moments. For instance, a magnetic Bragg reflection from a collinear antiferromagnetic order with magnetic moments parallel to the z axis produces only NSF scattering signals, which are detected when the spin flipper is off, as shown in Fig. 2(a). As an another example, a cycloidal magnetic structure with magnetic moments on the xy plane leads to only SF scattering signals, which are observed as \(I_{\text{ON}}\), as shown in Fig. 2(b). As for a screw-type magnetic order with a q-vector on the xy plane, the NSF scattering is always observed owing to the finite z component of the Fourier-transformed magnetic moments. On the other hand, the SF intensity changes with the angle between the scattering vector and the q-vector, as we see in the next section.
Figure 2. (Color online) Schematics showing possible SF and NSF scattering processes for magnetic Bragg reflections from (a) a collinear antiferromagnetic order with magnetic moments parallel to the z axis and (b) a cycloidal magnetic order with the magnetic moments in the xy plane.
Besides the \(P_{zz}\) configuration, we can also set the polarization direction of the incident neutrons to be parallel to the scattering vector by applying a weak horizontal field by the Helmholtz coil. In this case, both the y and z components of \(\boldsymbol{M}^{\bot}({\boldsymbol{\kappa}})\) are perpendicular to the neutron spins, and thus observed as the SF scattering. On the other hand, the nuclear scattering is always observed in the NSF channel, because it does not change the spin state of the neutrons. This configuration is referred to as \(P_{xx}\), which is used for separating magnetic scattering from nuclear scattering. The \(P_{xx}\) configuration is also used for determining sense of spin rotation in helical magnetic orders.9–14)
These selection rules regarding the spin rotation is generalized to include the inelastic scattering. Thus polarized neutron inelastic scattering can separate magnon excitations from the phonon excitations, and can detect the sense of the spin precession in magnon excitations.15)
Examples of the \(P_{xx}\) and \(P_{zz}\) polarization analyses for elastic scattering
We applied the \(P_{zz}\) polarization analysis to determine the magnetic structure in the ground state of the centrosymmetric magnetic skyrmion host Gd2PdSi3.16) This compound has a hexagonal crystal structure, in which magnetic Gd3+ ions are arranged to form triangular lattice layers stacked along the c axis. The recent resonant x-ray magnetic scattering study by Kurumaji et al. revealed that the ground state has screw-type magnetic modulations with the q-vectors of \((q,0,0)\), \((0,q,0)\), and \((q,-q,0)\), which are equivalent to each other under the threefold rotational symmetry about the c axis.17) The modulation wavenumber q is approximately 0.14 (r.l.u.). In general, resonant x-ray scattering is sensitive not only to magnetic modulations but also charge and orbital modulations. We thus performed polarized neutron scattering measurements, which can exclusively detect magnetic moments, to verify the magnetic structure. For this experiment, we grew an isotope-enriched 160Gd2PdSi3 single crystal to reduce the strong neutron absorption effect of natural Gd. The sample was mounted in a closed-cycle 4He refrigerator with the \((H,K,0)\) scattering plane. Other details of the experimental conditions are described in Ref. 16.
Figure 3(a) shows the polarized neutron scattering profile of a magnetic Bragg reflection at \((0,q,0)\) measured in the \(P_{zz}\) configuration. Both the SF and NSF are observed, which is consistent with the screw-type magnetic structure. However, the presence or absence of the magnetic moments along the q-vector direction was not concluded from this data, because the scattering vector is parallel to the q-vector, as shown in Fig. 3(b). We thus measured another magnetic reflection at \((-1+q,2,0)\), where the scattering vector and the q-vector are nearly perpendicular to each other, as shown in Fig. 3(d). If the spin component parallel to the q-vector exists, it should be observed in the SF channel. Figure 3(c) shows the observed profile, revealing that the SF scattering is absent at this reflection. This unambiguously shows that the spin spiral plane is perpendicular to the q-vector. We also note that the NSF and SF intensities at \((0,q,0)\) are not equal to each other. This means that the screw structure has ellipticity, which reflects the existence of the magnetic anisotropy for the magnetic moments on the Gd sites.
Figure 3. (Color online) (a) Polarized neutron scattering profile of the magnetic Bragg reflection at \((0,q,0)\) in 160Gd2PdSi3. (b) Schematic showing the directions of \({\boldsymbol{\kappa}}\), q-vector and guide field \(\boldsymbol{H}\) with respect to the xyz coordinates when measuring the reflection at \((0, q, 0)\). (c, d) Those for the reflection at \((-1+q,2,0)\). (e) Polarized neutron scattering profiles of the reflection at \((0,q,0)\) measured with the \(P_{xx}\) configuration. (f) Schematic showing the directions of \({\boldsymbol{\kappa}}\), q-vector and guide field \(\boldsymbol{H}\) for the \(P_{xx}\) measurement at \((0,q,0)\). All the data were measured at 2.5 K. The above profiles and schematics are replotted and redrawn using the data presented in Ref. 16.
We also note here that the polarization analysis with the \(P_{zz}\) configuration is also suited for studying magnetic structures when the sample has relatively strong neutron absorption. The ratio between NSF and SF intensities does not depend on the absorption effect, because they are measured through the same scattering path. We applied this technique to investigate magnetic structures in another centrosymmetric skyrmion host EuAl4.18)
In the above discussion, the NSF intensities observed at the magnetic Bragg peaks in the \(P_{zz}\) configuration were interpreted as the presence of the c component of the magnetic modulation. However, the NSF signal can arise from nonmagnetic modulations such as lattice modulations associated with charge density waves. To exclude this possibility, we also measured the magnetic Bragg reflection at \((0,q,0)\) with the \(P_{xx}\) configuration, in which a weak horizontal guide field was applied along the x axis [see Fig. 3(f)]. As shown in Fig. 3(e), the intensities were observed only in the SF channel, demonstrating that the data observed in the \(P_{zz}\) configuration does not contain nonmagnetic scattering.
Example of the \(P_{xx}\) polarization analysis for inelastic scattering
The longitudinal polarization analysis is applicable not only to elastic scattering but also inelastic scattering. For instance, we used the \(P_{xx}\) polarization analysis to study magnetic excitation of NiO, which exhibits a commensurate antiferromagnetic order with the q-vectors of \((\frac{1}{2},\frac{1}{2},\frac{1}{2})\) and its equivalents.19) In this experiment, we employed the PG(002) monochromator, which the has vertical focusing option, to increase the neutron flux at the sample position. To polarize the incident neutron beam, a supermirror spin polarizer was installed between the monochromator and the sample. The scattered neutrons were analyzed by the Heusler(111) analyzer. Three co-aligned single crystals of NiO with the total mass of approximately 10 g were mounted in an Al cell, which was attached onto the cold head of a closed-cycle 4He cryostat.
The measurement was performed with the \(E_{f}\)-fix mode, in which we fixed the energy of the scattered neutrons to be 14.7 meV. Accordingly, we changed the energy transfer by varying the energy of the incident neutrons. At each energy transfer, we measured the signals from the sample until the monitor count reached \(3.96\times 10^{6}\), which took approximately 9.3 and 7.7 min when the energy transfer was 0 and 17 meV, respectively. Figure 4 shows the energy-transfer dependence of the NSF and SF intensities at \((\frac{1}{2},\frac{1}{2},-\frac{1}{2})\), at which a magnetic elastic peak is observed at the elastic condition. We observed that the inelastic scattering intensity develops above 5 meV, and that the signals are mostly observed in the SF channel. This indicates the presence of the magnon excitation with the wave vector of \((\frac{1}{2},\frac{1}{2},-\frac{1}{2})\).
Figure 4. (Color online) Polarized inelastic scattering profile of the constant-Q scan at \((\frac{1}{2},\frac{1}{2},-\frac{1}{2})\) in NiO at 35 K.
4. Unpolarized Neutron Diffraction Measurements with a 2D Position Sensitive Detector
As we describe in the previous sections, polarized neutron scattering is a powerful technique to determine the orientations of the magnetic moments in the sample. However, the intensity of the polarized neutron beam is approximately two orders of magnitude weaker than that of the unpolarized beam at PONTA. We thus complementary use the polarized and unpolarized neutrons in most of the experiments. A major demand for unpolarized measurements is to efficiently measure the peak positions and intensities of magnetic reflections. To meet the demand, we recently started commissioning of a two-dimensional position sensitive detector (2D-PSD). We employ the detector which was originally developed for SENJU diffractometer in Materials and Life science experimental Facility (MLF) of J-PARC.20) The detector has a detection area of \(256\times 256\) mm, which is divided into \(64\times 64\) pixels. We put the detector at a distance of 733 mm from the sample, so that the horizontal coverage of the scattering angle is approximately 20 degrees. We selected the 160Gd2PdSi3 crystal for the test measurement, and measured satellite reflections near the 100 nuclear reflection. The intensities at each pixel were recorded with varying ω angle of the sample, and were transformed to an intensity map on the reciprocal space. Figure 5 shows the intensity distribution on the horizontal plane, namely the \((H,K,0)\) plane. The 100 nuclear reflection and surrounding by six satellite reflections are clearly observed.
Figure 5. (Color online) Intensity map on the \((H,K,0)\) plane of the 160Gd2PdSi3 single crystal measured by the 2D-PSD. Inset shows the definitions of the \(Q_{x}\) and \(Q_{y}\) axes, which are a Cartesian coordinate system with the unit of Å−1, and the positions of the nuclear and magnetic reflections. The shaded area in the inset corresponds the area measured by the ω rotation scans with the 2D-PSD.
We note here that the 2D-PSD will be useful when searching for magnetic reflections in unknown magnetic phases. The vertical coverage of the 2D-PSD enables us to observe magnetic reflections out of the horizontal plane. Moreover, the 2D-PSD would be useful to investigate magnetic orders in external fields such as magnetic field, electric field, uniaxial stress, etc. The apparatuses to apply magnetic field or uniaxial stress often bring the restrictions regarding the scattering plane. For instance, most of the high-field magnets are designed to have vertical magnetic field, and thus the scattering plane for a standard triple-axis spectrometer has to be perpendicular to the magnetic field. The accessibility to the out-of-plane reflections is essential to overcome this problem, and will pave the way for studying spin-driven cross-correlated phenomena induced by these external fields.
5. Summary and Outlooks
We introduced the polarized and unpolarized neutron scattering environments at PONTA spectrometer in JRR-3. The polarized neutron scattering is useful to determine the orientation of the magnetic moments and thus applied for studies on non-collinear/non-coplanar magnetic orders.18,21) The major drawback of the polarized neutron scattering is the extremely weak intensity. We are now planning to introduce supermirror spin analyzer for scattered neutrons, which is expected to have larger transmission as compared to the reflectivity of the Heusler(111) crystal analyzer. The 2D-PSD for unpolarized measurement will be useful to efficiently measure the intensity distributions in the reciprocal space. We note here that this 2D-PSD is also applicable for time-resolved neutron scattering measurements, because it was originally developed for the event-based data recording system in J-PARC. This feature will enable us to study transient phenomena under time-dependent external fields, such as rapid sweeps of temperature or magnetic fields. In fact, recent studies show that strongly correlated systems can a variety of metastable electronic phases.22) Neutron scattering studies on the metastable states would be fascinating topics in the future.
Acknowledgements
We gratefully acknowledge the supports from E-IMR and the Center of Neutron Science for Advanced Materials, Institute for Materials Research, Tohoku University. We also thank Mr. M. Ohkawara (Tohoku Univ.) for his kind support on the neutron experiments at JRR-3. The neutron scattering experiments at PONTA in JRR-3 were carried out along the proposals (Nos. 21507, 21401, 22518, and 22401).
References
1 T. Kimura, T. Goto, H. Shintani, K. Ishizaka, T. Arima, and Y. Tokura, Nature 426, 55 (2003). 10.1038/nature02018 Crossref, Google Scholar
2 M. Kenzelmann, A. B. Harris, S. Jonas, C. Broholm, J. Schefer, S. B. Kim, C. L. Zhang, S.-W. Cheong, O. P. Vajk, and J. W. Lynn, Phys. Rev. Lett. 95, 087206 (2005). 10.1103/PhysRevLett.95.087206 Crossref, Google Scholar
3 S. Mühlbauer, B. Binz, F. Jonietz, C. Pfleiderer, A. Rosch, A. Neubauer, R. Georgii, and P. Böni, Science 323, 915 (2009). 10.1126/science.1166767 Crossref, Google Scholar
4 A. Neubauer, C. Pfleiderer, B. Binz, A. Rosch, R. Ritz, P. G. Niklowitz, and P. Böni, Phys. Rev. Lett. 102, 186602 (2009). 10.1103/PhysRevLett.102.186602 Crossref, Google Scholar
5 S. Nakatsuji, N. Kiyohara, and T. Higo, Nature 527, 212 (2015). 10.1038/nature15723 Crossref, Google Scholar
6 T. Yokouchi, F. Kagawa, M. Hirschberger, Y. Otani, N. Nagaosa, and Y. Tokura, Nature 586, 232 (2020). 10.1038/s41586-020-2775-x Crossref, Google Scholar
8 G. L. Squires, Introduction to the Theory of Thermal Neutron Scattering (Dover, New York, 1978). Google Scholar
9 Y. Yamasaki, H. Sagayama, T. Goto, M. Matsuura, K. Hirota, T. Arima, and Y. Tokura, Phys. Rev. Lett. 98, 147204 (2007). 10.1103/PhysRevLett.98.147204 Crossref, Google Scholar
10 Y. Yamasaki, H. Sagayama, T. Goto, M. Matsuura, K. Hirota, T. Arima, and Y. Tokura, Phys. Rev. Lett. 100, 219902 (2008). 10.1103/PhysRevLett.100.219902 Crossref, Google Scholar
11 S. Seki, Y. Yamasaki, M. Soda, M. Matsuura, K. Hirota, and Y. Tokura, Phys. Rev. Lett. 100, 127201 (2008). 10.1103/PhysRevLett.100.127201 Crossref, Google Scholar
12 H. Sagayama, K. Taniguchi, N. Abe, T.-h. Arima, M. Soda, M. Matsuura, and K. Hirota, Phys. Rev. B 77, 220407 (2008). 10.1103/PhysRevB.77.220407 Crossref, Google Scholar
13 M. Soda, K. Kimura, T. Kimura, M. Matsuura, and K. Hirota, J. Phys. Soc. Jpn. 78, 124703 (2009). 10.1143/JPSJ.78.124703 Link, Google Scholar
14 T. Nakajima, S. Mitsuda, S. Kanetsuki, K. Tanaka, K. Fujii, N. Terada, M. Soda, M. Matsuura, and K. Hirota, Phys. Rev. B 77, 052401 (2008). 10.1103/PhysRevB.77.052401 Crossref, Google Scholar
15 Y. Nambu, J. Barker, Y. Okino, T. Kikkawa, Y. Shiomi, M. Enderle, T. Weber, B. Winn, M. Graves-Brook, J. M. Tranquada, T. Ziman, M. Fujita, G. E. W. Bauer, E. Saitoh, and K. Kakurai, Phys. Rev. Lett. 125, 027201 (2020). 10.1103/PhysRevLett.125.027201 Crossref, Google Scholar
16 J. Ju, H. Saito, T. Kurumaji, M. Hirschberger, A. Kikkawa, Y. Taguchi, T.-H. Arima, Y. Tokura, and T. Nakajima, Phys. Rev. B 107, 024405 (2023). 10.1103/PhysRevB.107.024405 Crossref, Google Scholar
17 T. Kurumaji, T. Nakajima, M. Hirschberger, A. Kikkawa, Y. Yamasaki, H. Sagayama, H. Nakao, Y. Taguchi, T.-h. Arima, and Y. Tokura, Science 365, 914 (2019). 10.1126/science.aau0968 Crossref, Google Scholar
18 R. Takagi, N. Matsuyama, V. Ukleev, L. Yu, J. S. White, S. Francoual, J. R. L. Mardegan, S. Hayami, H. Saito, K. Kaneko, K. Ohishi, Y. Onuki, T.-H. Arima, Y. Tokura, T. Nakajima, and S. Seki, Nat. Commun. 13, 1472 (2022). 10.1038/s41467-022-29131-9 Crossref, Google Scholar
19 W. L. Roth, Phys. Rev. 110, 1333 (1958). 10.1103/PhysRev.110.1333 Crossref, Google Scholar
20 T. Nakamura, T. Kawasaki, T. Hosoya, K. Toh, K. Oikawa, K. Sakasai, M. Ebine, A. Birumachi, K. Soyama, and M. Katagiri, Nucl. Instrum. Methods Phys. Res., Sect. A 686, 64 (2012). 10.1016/j.nima.2012.05.038 Crossref, Google Scholar
21 H. Takagi, R. Takagi, S. Minami, T. Nomoto, K. Ohishi, M.-T. Suzuki, Y. Yanagi, M. Hirayama, N. D. Khanh, K. Karube, H. Saito, D. Hashizume, R. Kiyanagi, Y. Tokura, R. Arita, T. Nakajima, and S. Seki, Nat. Phys. 19, 961 (2023). 10.1038/s41567-023-02017-3 Crossref, Google Scholar
22 F. Kagawa and H. Oike, Adv. Mater. 29, 1601979 (2017). 10.1002/adma.201601979 Crossref, Google Scholar