Location via proxy:   [ UP ]  
[Report a bug]   [Manage cookies]                
License: CC BY 4.0
arXiv:2303.03682v2 [cond-mat.mtrl-sci] 05 Dec 2023

Stacking disorder and thermal transport properties of Î±đ›Œ\alphaitalic_α-RuCl33{}_{3}start_FLOATSUBSCRIPT 3 end_FLOATSUBSCRIPT

Heda Zhang Materials Science and Technology Division, Oak Ridge National Laboratory, Oak Ridge, Tennessee 37831, USA    Michael A McGuire Materials Science and Technology Division, Oak Ridge National Laboratory, Oak Ridge, Tennessee 37831, USA    Andrew F May Materials Science and Technology Division, Oak Ridge National Laboratory, Oak Ridge, Tennessee 37831, USA    Joy Chao Center for Nanophase Materials Sciences, Oak Ridge National Laboratory, Oak Ridge, Tennessee 37831, USA    Qiang Zheng Materials Science and Technology Division, Oak Ridge National Laboratory, Oak Ridge, Tennessee 37831, USA Department of Materials Science and Engineering, University of Tennessee, Knoxville, TN 37996, USA    Miaofang Chi Center for Nanophase Materials Sciences, Oak Ridge National Laboratory, Oak Ridge, Tennessee 37831, USA    Brian C Sales Materials Science and Technology Division, Oak Ridge National Laboratory, Oak Ridge, Tennessee 37831, USA    David G Mandrus Materials Science and Technology Division, Oak Ridge National Laboratory, Oak Ridge, Tennessee 37831, USA Department of Materials Science and Engineering, University of Tennessee, Knoxville, TN 37996, USA    Stephen E Nagler Neutron Scattering Division, Oak Ridge National Laboratory, Oak Ridge, Tennessee 37831, USA    Hu Miao Materials Science and Technology Division, Oak Ridge National Laboratory, Oak Ridge, Tennessee 37831, USA    Feng Ye yef1@ornl.gov Neutron Scattering Division, Oak Ridge National Laboratory, Oak Ridge, Tennessee 37831, USA    Jiaqiang Yan yanj@ornl.gov Materials Science and Technology Division, Oak Ridge National Laboratory, Oak Ridge, Tennessee 37831, USA
(December 5, 2023)
Abstract

Î±đ›Œ\alphaitalic_α-RuCl33{}_{3}start_FLOATSUBSCRIPT 3 end_FLOATSUBSCRIPT, a well-known candidate material for Kitaev quantum spin liquid, is prone to stacking disorder due to the weak van der Waals bonding between the honeycomb layers. After a decade of intensive experimental and theoretical studies, the detailed correlation between stacking degree of freedom, structure transition, magnetic and thermal transport properties remains unresolved. In this work, we reveal the effects of a small amount of stacking disorder inherent even in high quality Î±đ›Œ\alphaitalic_α-RuCl33{}_{3}start_FLOATSUBSCRIPT 3 end_FLOATSUBSCRIPT crystals. This small amount of stacking disorder results in the variation of the magnetic ordering temperature, suppresses the structure transition and thermal conductivity. Crystals with minimal amount of stacking disorder have a T>N{}_{N}>start_FLOATSUBSCRIPT italic_N end_FLOATSUBSCRIPT >7.4 K and exhibit a well-defined structure transition around 140 K upon cooling. For those with more stacking faults and a TN𝑁{}_{N}start_FLOATSUBSCRIPT italic_N end_FLOATSUBSCRIPT below 7 K, the structure transition occurs well below 140 K upon cooling and is incomplete, manifested by the diffuse streaks and the coexistence of both high temperature and low temperature phases down to the lowest measurement temperature. Both types of crystals exhibit oscillatory field dependent thermal conductivity and a plateau-like feature in thermal Hall resistivity in the field-induced quantum spin liquid state. However, Î±đ›Œ\alphaitalic_α-RuCl33{}_{3}start_FLOATSUBSCRIPT 3 end_FLOATSUBSCRIPT crystals with minimal amount of stacking disorder have a higher thermal conductivity that pushes the thermal Hall conductivity to be closer to the half-integer quantized value. These findings demonstrate a strong correlation between layer stacking, structure transition, magnetic and thermal transport properties, underscoring the importance of interlayer coupling in Î±đ›Œ\alphaitalic_α-RuCl33{}_{3}start_FLOATSUBSCRIPT 3 end_FLOATSUBSCRIPT despite the weak van der Waals bonding.

Shortly after the first experimental investigation of Î±đ›Œ\alphaitalic_α-RuCl33{}_{3}start_FLOATSUBSCRIPT 3 end_FLOATSUBSCRIPT as the candidate material for Kitaev quantum spin liquidPlumb et al. (2014), it was realized that the stacking sequence of the RuCl33{}_{3}start_FLOATSUBSCRIPT 3 end_FLOATSUBSCRIPT honeycomb layers can affect the magnetic properties Banerjee et al. (2016). This is typical for van der Waals bonded cleavable magnets where nowadays the stacking degree of freedom of the two dimensional structure units has been employed to engineer magnetic ground states Huang et al. (2017); Chen et al. (2019); Yan (2022). The stacking of RuCl33{}_{3}start_FLOATSUBSCRIPT 3 end_FLOATSUBSCRIPT honeycomb layers relative to each other leads to different proposed crystal structures including the monoclinic C2/m, the rhombohedral R⁹3¯𝑅¯3R\bar{3}italic_R overÂŻ start_ARG 3 end_ARG, and the trigonal P⁹31⁹12𝑃subscript3112P3_{1}12italic_P 3 start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT 12. At room temperature, most studies reported a C⁹2/mđ¶2𝑚C2/mitalic_C 2 / italic_m structure for Î±đ›Œ\alphaitalic_α-RuCl33{}_{3}start_FLOATSUBSCRIPT 3 end_FLOATSUBSCRIPT. Upon cooling, this monoclinic structure becomes unstable and around 150 K Î±đ›Œ\alphaitalic_α-RuCl33{}_{3}start_FLOATSUBSCRIPT 3 end_FLOATSUBSCRIPT goes through a first order structure transition most likely to the rhombohedral R⁹3¯𝑅¯3R\bar{3}italic_R overÂŻ start_ARG 3 end_ARG structure Mu et al. (2022). In early studies of Î±đ›Œ\alphaitalic_α-RuCl33{}_{3}start_FLOATSUBSCRIPT 3 end_FLOATSUBSCRIPT, more than one magnetic anomaly is typically observed in the temperature range 7-14 K Kubota et al. (2015); Johnson et al. (2015); Sears et al. (2015); Majumder et al. (2015); Banerjee et al. (2016). The 14 K magnetic order seems to result from the C⁹2/mđ¶2𝑚C2/mitalic_C 2 / italic_m type stacking Cao et al. (2016); Johnson et al. (2015). And it is believed that the mixing of different types of stacking gives rise to the multiple magnetic anomalies in the temperature range 7 K-14 K.

With the intense materials synthesis effort, most Î±đ›Œ\alphaitalic_α-RuCl33{}_{3}start_FLOATSUBSCRIPT 3 end_FLOATSUBSCRIPT crystals being studied nowadays have only one single magnetic transition around TN𝑁{}_{N}start_FLOATSUBSCRIPT italic_N end_FLOATSUBSCRIPT=7 K. This has been employed as a simple and convenient criteria for a quick check of crystal quality. However, TN𝑁{}_{N}start_FLOATSUBSCRIPT italic_N end_FLOATSUBSCRIPT has been found to vary from 6 K to 8 K Kim et al. (2022); Sears et al. (2017); Do et al. (2017). The origin of this discrepancy is still unknown. Also unknown is whether and how this magnetic order is correlated with the first order structure transition at high temperatures.

Recently, the intensively debated thermal transport properties of Î±đ›Œ\alphaitalic_α-RuCl33{}_{3}start_FLOATSUBSCRIPT 3 end_FLOATSUBSCRIPT calls for a thorough revisit to the materials issues of this fascinating compound Lee (2021). The half-integer quantized thermal Hall conductance was believed to be one of the fingerprints for Majorana fermions of the fractionalized spin excitations in Î±đ›Œ\alphaitalic_α-RuCl33{}_{3}start_FLOATSUBSCRIPT 3 end_FLOATSUBSCRIPT. However, the thermal Hall conductance in the field-induced disordered or quantum spin liquid state was found to be sample dependent and the experimental observation of half-integer quantized value requires using Î±đ›Œ\alphaitalic_α-RuCl33{}_{3}start_FLOATSUBSCRIPT 3 end_FLOATSUBSCRIPT crystals with high longitudinal thermal conductivity Yokoi et al. (2021); Kasahara et al. (2018); Yamashita et al. (2020); Lefrançois et al. (2022); Czajka et al. (2022); Bruin et al. (2022a); Kasahara et al. (2022). The other intriguing experimental observation is the oscillatory features of thermal conductivity as a function of in-plane magnetic field. These features were reproduced by different groups but the origin is under hot debate. While some believed this is an intrinsic character of the magnetic phase of T≈N{}_{N}\approxstart_FLOATSUBSCRIPT italic_N end_FLOATSUBSCRIPT ≈7 K Zhang et al. (2023); Czajka et al. (2021) and attributed the observed oscillations to quantum oscillations of putative charge-neutral fermions Czajka et al. (2021), others believed that the oscillatory features have an extrinsic origin and result from a sequence of field-induced magnetic phase transitions in crystals with stacking disorder Bruin et al. (2022b); Lefrançois et al. (2023).

All these interesting but debated results and interpretations highlight the importance of understanding the materials issues in Î±đ›Œ\alphaitalic_α-RuCl33{}_{3}start_FLOATSUBSCRIPT 3 end_FLOATSUBSCRIPT. In this work, we performed a careful investigation of the correlation between the layer stacking, structure transition, magnetic and thermal transport properties of different Î±đ›Œ\alphaitalic_α-RuCl33{}_{3}start_FLOATSUBSCRIPT 3 end_FLOATSUBSCRIPT crystals with TN𝑁{}_{N}start_FLOATSUBSCRIPT italic_N end_FLOATSUBSCRIPT varying from 6 K to 7.6 K. The amount of stacking disorder discussed in this work is far less than that in previous studies Kubota et al. (2015); Johnson et al. (2015); Sears et al. (2015); Majumder et al. (2015); Banerjee et al. (2016) and may not induce significant magnetic anomalies at 10-14 K in magnetic and specific heat measurements. This small amount of stacking disorder causes the variations in TN𝑁{}_{N}start_FLOATSUBSCRIPT italic_N end_FLOATSUBSCRIPT in different Î±đ›Œ\alphaitalic_α-RuCl33{}_{3}start_FLOATSUBSCRIPT 3 end_FLOATSUBSCRIPT crystals. Based on the characterizations, we categorize our Î±đ›Œ\alphaitalic_α-RuCl33{}_{3}start_FLOATSUBSCRIPT 3 end_FLOATSUBSCRIPT crystals into two types. Type-I Î±đ›Œ\alphaitalic_α-RuCl33{}_{3}start_FLOATSUBSCRIPT 3 end_FLOATSUBSCRIPT single crystals with TN𝑁{}_{N}start_FLOATSUBSCRIPT italic_N end_FLOATSUBSCRIPT above 7.2 K have minimal amount of stacking disorder and show a well defined structure transition around 140 K upon cooling. For type-II crystals with more stacking disorder and a lower TN𝑁{}_{N}start_FLOATSUBSCRIPT italic_N end_FLOATSUBSCRIPT, the structure transition occurs below 140 K upon cooling and is incomplete, manifested by the coexistence of both high temperature and low temperature phases and the observation of diffuse streaks by neutron scattering below the structure transition. For both types of crystals, oscillatory field dependent thermal conductivity and plateau like feature in thermal hall resistivity were observed in the field-induced quantum spin liquid state. However, Î±đ›Œ\alphaitalic_α-RuCl33{}_{3}start_FLOATSUBSCRIPT 3 end_FLOATSUBSCRIPT single crystals with TN𝑁{}_{N}start_FLOATSUBSCRIPT italic_N end_FLOATSUBSCRIPT above 7.4 K have a higher thermal conductivity, which pushes the thermal Hall conductivity to be closer to the half integer quantized value. Our results demonstrate a strong correlation between the layer stacking, structure transition, magnetic and thermal transport properties and highlight the importance of interlayer coupling in Î±đ›Œ\alphaitalic_α-RuCl33{}_{3}start_FLOATSUBSCRIPT 3 end_FLOATSUBSCRIPT.

I Experimental details

Î±đ›Œ\alphaitalic_α-RuCl33{}_{3}start_FLOATSUBSCRIPT 3 end_FLOATSUBSCRIPT used in this study were grown using two vapor transport techniques: self-selecting vapor growth with a small temperature gradient near the powder, and the conventional vapor transport technique with a large temperature gradient. The former is employed to grow thicker crystals for neutron single crystal diffraction measurements and the growth details can be found elsewhere Yan and McGuire (2022). The latter was performed in a two-zone tube furnace with the hot end kept at 1000°°\degree°C and the cold end at 750°°\degree°C. This yields millimeter sized single crystals with minimal amount of stacking disorder that are ideal for measurements of magnetic, thermodynamic, and thermal transport properties. The growth details are reported previouslyZhang et al. (2023). Varying the growth temperature and temperature gradient along the growth ampoule impacts the layer stacking and thus physical properties. For example, Î±đ›Œ\alphaitalic_α-RuCl33{}_{3}start_FLOATSUBSCRIPT 3 end_FLOATSUBSCRIPT crystals with TN𝑁{}_{N}start_FLOATSUBSCRIPT italic_N end_FLOATSUBSCRIPT around 6 K were obtained by keeping the hot end of the growth ampoule at 1075°°\degree°C while the cold end at 1025°°\degree°C. We noticed that crystal quality and properties are sensitive to the total vapor pressure inside of the growth ampoule. Our growths suggest a higher vapor pressure facilitates the formation of stacking disorder, although a detailed relation is not developed yet. Therefore, in order to obtain Î±đ›Œ\alphaitalic_α-RuCl33{}_{3}start_FLOATSUBSCRIPT 3 end_FLOATSUBSCRIPT crystals with controlled degree of stacking disorder, all factors affecting the total vapor pressure inside of the growth ampoule should be considered, such as, the purity and amount of starting powder, the volume of growth ampoule, growth temperature and temperature gradient along the growth ampoule. In our growths, high pure RuCl33{}_{3}start_FLOATSUBSCRIPT 3 end_FLOATSUBSCRIPT powder synthesized by reacting RuO22{}_{2}start_FLOATSUBSCRIPT 2 end_FLOATSUBSCRIPT powder with Al22{}_{2}start_FLOATSUBSCRIPT 2 end_FLOATSUBSCRIPTO33{}_{3}start_FLOATSUBSCRIPT 3 end_FLOATSUBSCRIPT-KCl saltYan et al. (2017) or purchased from Furuya Metals (Japan) was used and they are found to be comparable with respect to disorder control.

Elemental analysis on cleaved surfaces was performed using the wavelength dispersive (WDS) spectroscopy techniques. WDS measurement was performed using a JEOL JXA-8200X electron microprobe analyzer instrument equipped with five crystal-focusing spectrometers for wavelength dispersive x-ray spectroscopy. Magnetic properties were measured with a Quantum Design (QD) Magnetic Property Measurement System in the temperature range 2.0 K≀\leq≀T≀\leq≀ 300 K. Specific heat data below 30 K were collected using a QD Physical Property Measurement System (PPMS). Thermal transport measurements were performed on a custom-built PPMS puck as described beforeZhang et al. (2023). The temperature dependence of 0 0 l reflections was monitored on a flat surface by a PANalytical X’Pert Pro MPD powder x-ray diffractometer using Cu Kα⁹1đ›Œ1{}_{\alpha 1}start_FLOATSUBSCRIPT italic_α 1 end_FLOATSUBSCRIPT radiation. An Oxford PheniX closed cycle cryostat was used to measure from 300 K to 20 K.

Neutron diffraction experiments were carried out using the single crystal diffractometer CORELLI Ye et al. (2018) at SNS to monitor the structural phase transition and to search for potential diffuse scattering arising from possible stacking disorder in crystals grown under different conditions. Single crystals are mounted inside a closed-cycle refrigerator with a base temperature of 6 K. The Mantidworkbench package was used for data reduction and analysis. We define the momentum transfer Q in 3⁹D3đ·3D3 italic_D reciprocal space in Å−1superscriptitalic-Å1\AA^{-1}italic_Å start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT as Q=H⁹a*+K⁹b*+L⁹c*đ‘„đ»superscriptđ‘ŽđŸsuperscript𝑏𝐿superscript𝑐Q=Ha^{*}+Kb^{*}+Lc^{*}italic_Q = italic_H italic_a start_POSTSUPERSCRIPT * end_POSTSUPERSCRIPT + italic_K italic_b start_POSTSUPERSCRIPT * end_POSTSUPERSCRIPT + italic_L italic_c start_POSTSUPERSCRIPT * end_POSTSUPERSCRIPT, in which Hđ»Hitalic_H, KđŸKitalic_K and L𝐿Litalic_L are Miller indices, and a*superscript𝑎a^{*}italic_a start_POSTSUPERSCRIPT * end_POSTSUPERSCRIPT, b*superscript𝑏b^{*}italic_b start_POSTSUPERSCRIPT * end_POSTSUPERSCRIPT, c*superscript𝑐c^{*}italic_c start_POSTSUPERSCRIPT * end_POSTSUPERSCRIPT are the lattice vectors in reciprocal space. For each type of crystals mentioned below, we checked one small piece around 6 mg and one large piece around 200 mg. No difference was observed between the small and large pieces.

II Results

Refer to caption
Figure 1: (color online) Magnetic order and structure transition in two typical types of Î±đ›Œ\alphaitalic_α-RuCl33{}_{3}start_FLOATSUBSCRIPT 3 end_FLOATSUBSCRIPT single crystals. As described in the text, type-I crystals have magnetic ordering temperatures above 7.2 K, while type-II crystals normally order magnetically below 7 K. (a,b) Temperature dependence of magnetization and specific heat (Cp) below 20 K. The magnetization was measured in a magnetic field of 1 kOe applied along the zig-zag direction (perpendicular to the Ru-Ru bond). The vertical dashed lines highlight the Neel temperature, TN𝑁{}_{N}start_FLOATSUBSCRIPT italic_N end_FLOATSUBSCRIPT, defined as the temperature where Cp peaks. TN𝑁{}_{N}start_FLOATSUBSCRIPT italic_N end_FLOATSUBSCRIPT=7.6 K in (a) is 1.1 K higher than that in (b). (c-e) Temperature dependence of magnetization and 005 reflection in the temperature range 20 K-200 K for the crystal with TN𝑁{}_{N}start_FLOATSUBSCRIPT italic_N end_FLOATSUBSCRIPT=7.6 K, a representative for type-I crystals. (f-h) Temperature dependence of magnetization and 005 reflection in the temperature range 20 K-200 K for the crystal with TN𝑁{}_{N}start_FLOATSUBSCRIPT italic_N end_FLOATSUBSCRIPT=6.5 K, an example for type-II crystals. The temperature dependence of magnetization in (d,g) was measured in a magnetic field of 10 kOe applied perpendicular to the honeycomb plane.

Figure 1 shows the temperature dependence of magnetization, specific heat, and intensity of the 005 structural Bragg reflection (in C2/m notation) illustrating the magnetic ordering and structure transition temperatures for two typical types of single crystals. We define the magnetic ordering temperature as the temperature where Cp(T) peaks. Type-I crystals have T>N{}_{N}>start_FLOATSUBSCRIPT italic_N end_FLOATSUBSCRIPT >7.2 K and show sharp anomalies at TN𝑁{}_{N}start_FLOATSUBSCRIPT italic_N end_FLOATSUBSCRIPT in both temperature dependence of magnetization and specific heat. Results of one crystal with TN𝑁{}_{N}start_FLOATSUBSCRIPT italic_N end_FLOATSUBSCRIPT=7.6 K (see Fig. 1a) are presented in this work as a representative of type-I crystals. For this crystal, abrupt changes in the temperature dependence of magnetization (Fig. 1d) and 005 reflection suggest the structure transition occurs at about 140 K upon cooling (Fig. 1c) and about 170 K upon warming(Fig. 1e). Upon cooling below 140 K, the intensity of the peak sitting at 2ξ𝜃\thetaitalic_ξ = 85.15 degree decreases quickly upon further cooling and disappears around 130 K. Meanwhile, a peak sitting at 2ξ𝜃\thetaitalic_ξ = 85.62 degree appears and its intensity increases quickly upon cooling and saturates below about 120 K. Step-like features associated with the structure transition are observed in the high temperature magnetization at 140 K and 170 K upon cooling and warming, respectively. Similar features are observed for all Î±đ›Œ\alphaitalic_α-RuCl33{}_{3}start_FLOATSUBSCRIPT 3 end_FLOATSUBSCRIPT crystals with TN𝑁{}_{N}start_FLOATSUBSCRIPT italic_N end_FLOATSUBSCRIPT above 7.2 K.

Type-II crystals (Fig. 1b) normally have a TN𝑁{}_{N}start_FLOATSUBSCRIPT italic_N end_FLOATSUBSCRIPT below 7 K and show broader transitions and weaker anomalies near TN𝑁{}_{N}start_FLOATSUBSCRIPT italic_N end_FLOATSUBSCRIPT in magnetization and specific heat. Physical properties of one crystal with TN𝑁{}_{N}start_FLOATSUBSCRIPT italic_N end_FLOATSUBSCRIPT=6.5 K were presented in this work as an example of type-II crystals. Above TN𝑁{}_{N}start_FLOATSUBSCRIPT italic_N end_FLOATSUBSCRIPT, some weak features were observed in both magnetization and specific heat in (b), in sharp contrast to (a). The structure transition also shows anomalous features distinct from those for type-I crystals. Upon cooling, the structure transition of the crystal with TN𝑁{}_{N}start_FLOATSUBSCRIPT italic_N end_FLOATSUBSCRIPT=6.5 K (see Fig. 1f) occurs around 60 K and the transition is not complete which is well illustrated by the coexistence of reflections from both high temperature and low temperature phases to the lowest temperature 20 K of our x-ray diffraction measurement. Upon warming from 20 K, the phase coexistence persists until about 170 K above which only the high temperature monoclinic phase remains(Fig. 1h). Accordingly, a wide loop was observed in the temperature dependence of magnetization (Fig. 1g). It is interesting to note that for both Î±đ›Œ\alphaitalic_α-RuCl33{}_{3}start_FLOATSUBSCRIPT 3 end_FLOATSUBSCRIPT crystals the structure transition occurs around 170 K upon warming. When screening crystals, we noticed that the structure transition can occur in a wide temperature range 50 K-140 K when measuring upon cooling. For type-II crystals, the phase coexistence is always observed for crystals with TN𝑁{}_{N}start_FLOATSUBSCRIPT italic_N end_FLOATSUBSCRIPT below 7 K. Fig.S1 of Supporting Materials shows the x-ray diffraction results of two different crystals. Both crystals have TN𝑁{}_{N}start_FLOATSUBSCRIPT italic_N end_FLOATSUBSCRIPT=7.1 K (from specific heat) and the structure transition at 120 K upon cooling. However, one crystal shows a complete structure transition while the other one shows phase coexistence below 120 K. 120 K seems to be the lowest structure transition temperature that one Î±đ›Œ\alphaitalic_α-RuCl33{}_{3}start_FLOATSUBSCRIPT 3 end_FLOATSUBSCRIPT crystal can have before showing a sluggish first order structure transition and phase coexistence at low temperatures.

Refer to caption
Figure 2: (color online) Neutron single crystal diffraction found a sharp structure transition for the crystal with TN𝑁{}_{N}start_FLOATSUBSCRIPT italic_N end_FLOATSUBSCRIPT=7.6 K but a sluggish structure transition and diffuse streaks for the crystal with TN𝑁{}_{N}start_FLOATSUBSCRIPT italic_N end_FLOATSUBSCRIPT=6.5 K. (a) The thermal evolution of the (1,1,6)116(1,1,6)( 1 , 1 , 6 ) reflection measured in cooling and warming confirms the sharp structural transition in the crystal with TN𝑁{}_{N}start_FLOATSUBSCRIPT italic_N end_FLOATSUBSCRIPT=7.6 K. (b, c) The neutron diffraction patterns at 200 K (in the high temperature C/2⁹mđ¶2𝑚C/2mitalic_C / 2 italic_m phase) and 100 K (in the low temperature R⁹3¯𝑅¯3R\bar{3}italic_R overÂŻ start_ARG 3 end_ARG phase) for the crystal with TN𝑁{}_{N}start_FLOATSUBSCRIPT italic_N end_FLOATSUBSCRIPT=7.6 K. Extra reflections in (c) confirm the symmetry change at low temperatures. (d) The temperature evolution of (1,1,3)113(1,1,3)( 1 , 1 , 3 ) reflection measured in cooling and warming of the crystal with TN𝑁{}_{N}start_FLOATSUBSCRIPT italic_N end_FLOATSUBSCRIPT=6.5 K. (e, f) The neutron diffraction patterns at 200 K (in the high temperature C/2⁹mđ¶2𝑚C/2mitalic_C / 2 italic_m phase) and 10 K (in the low temperature R⁹3¯𝑅¯3R\bar{3}italic_R overÂŻ start_ARG 3 end_ARG phase) for the crystal with TN𝑁{}_{N}start_FLOATSUBSCRIPT italic_N end_FLOATSUBSCRIPT=6.5 K. The mosaic in (e) and the diffuse streaks in (f) are highlighted by the line cut shown in the right half of each panel (b,c,e,f). The transformation matrix between the high-T𝑇Titalic_T C⁹2/mđ¶2𝑚C2/mitalic_C 2 / italic_m and low-T𝑇Titalic_T R⁹3¯𝑅¯3R{\bar{3}}italic_R overÂŻ start_ARG 3 end_ARG unit cell are: 𝐚r=𝐚msubscript𝐚𝑟subscript𝐚𝑚{\bf a}_{r}={\bf a}_{m}bold_a start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT = bold_a start_POSTSUBSCRIPT italic_m end_POSTSUBSCRIPT, 𝐛r=1/2⁹(−𝐚m+𝐛m)subscript𝐛𝑟12subscript𝐚𝑚subscript𝐛𝑚{\bf b}_{r}=1/2(-{\bf a}_{m}+{\bf b}_{m})bold_b start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT = 1 / 2 ( - bold_a start_POSTSUBSCRIPT italic_m end_POSTSUBSCRIPT + bold_b start_POSTSUBSCRIPT italic_m end_POSTSUBSCRIPT ), 𝐜r=𝐚m+3ⁱ𝐜msubscript𝐜𝑟subscript𝐚𝑚3subscript𝐜𝑚{\bf c}_{r}={\bf a}_{m}+3{\bf c}_{m}bold_c start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT = bold_a start_POSTSUBSCRIPT italic_m end_POSTSUBSCRIPT + 3 bold_c start_POSTSUBSCRIPT italic_m end_POSTSUBSCRIPT, where the subscript “r” and “m” denote the rhombhedral and monoclinic structures.

Considering the limited penetration depth of x-ray diffraction, we studied the structure of different pieces of pristine Î±đ›Œ\alphaitalic_α-RuCl33{}_{3}start_FLOATSUBSCRIPT 3 end_FLOATSUBSCRIPT crystals using single crystal neutron diffraction. All crystals studied with neutron diffraction were well characterized by measuring magnetic properties and/or specific heat before exposing to neutron beam. Multiple pieces with mass ranging from 6 mg to 200 mg and TN𝑁{}_{N}start_FLOATSUBSCRIPT italic_N end_FLOATSUBSCRIPT ranging from 6 K to 7.6 K were checked. All crystals crystallize in C2/m at room temperature but have different degree of mosaic. We are aware of previous reports of successful growth of Î±đ›Œ\alphaitalic_α-RuCl33{}_{3}start_FLOATSUBSCRIPT 3 end_FLOATSUBSCRIPT crystals with the P⁹31⁹12𝑃subscript3112P3_{1}12italic_P 3 start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT 12 structure Bruin et al. (2022b). Our neutron results do not rule out the possibility that some small crystals suitable for x-ray single crystal diffraction can have P⁹31⁹12𝑃subscript3112P3_{1}12italic_P 3 start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT 12 type structure. Figures 2(a,b) show the temperature dependence of (116) or (113) reflection to determine the structure transition in both crystals. For the crystal with TN𝑁{}_{N}start_FLOATSUBSCRIPT italic_N end_FLOATSUBSCRIPT=7.6 K, the structure transition occurs around 140 K upon cooling and 170 K upon warming, consistent with those determined from magnetic and x-ray diffraction measurements shown in Fig. 1. For the crystal with TN𝑁{}_{N}start_FLOATSUBSCRIPT italic_N end_FLOATSUBSCRIPT=6.5 K, the structure transition occurs below 80 K upon cooling and 170 K upon warming. The latter temperature agrees well with those determined from magnetic and X-ray diffraction measurements shown in Fig. 1. The slightly different transition temperature when measuring cooling is in line with the fact that this transition temperature can vary in a wide temperature range 50 K-140 K depending on the concentration and detailed distribution of stacking imperfection.

Refer to caption
Figure 3: (color online) Thermal transport measurement results of Î±đ›Œ\alphaitalic_α-RuCl33{}_{3}start_FLOATSUBSCRIPT 3 end_FLOATSUBSCRIPT. (a) Thermal conductivity of Î±đ›Œ\alphaitalic_α-RuCl33{}_{3}start_FLOATSUBSCRIPT 3 end_FLOATSUBSCRIPT crystal. Heat current is applied along the zig-zag direction (perpendicular to the Ru-Ru bond). (b) Oscillatory features of thermal conductivity. Data collected with both heat current and magnetic field along the zig-zag direction. (c, d) Thermal Hall resistivity (c) and conductivity (d) at 5 K. Both heat current and magnetic field are applied along the zig-zag direction. The horizontal dashed lines in (d) highlight the thermal Hall conductivity relative to the fraction of quantized thermal Hall ÎșQ⁹Hsubscriptđœ…đ‘„đ»\kappa_{QH}italic_Îș start_POSTSUBSCRIPT italic_Q italic_H end_POSTSUBSCRIPT/n, where n = 2,4,and 8 are shown.

Detailed reciprocal space maps provide better understanding of the nature of average and local structure for both Î±đ›Œ\alphaitalic_α-RuCl33{}_{3}start_FLOATSUBSCRIPT 3 end_FLOATSUBSCRIPT single crystals. Figures 2b and c show the neutron diffraction patterns collected at 200 K (above the structure transition) and 100 K (below the structure transition) for the type-I crystal with TN𝑁{}_{N}start_FLOATSUBSCRIPT italic_N end_FLOATSUBSCRIPT=7.6 K. As reported previously, the high temperature monoclinic C⁹2/mđ¶2𝑚C2/mitalic_C 2 / italic_m structure becomes unstable upon cooling and transforms into a rhombohedral R⁹3¯𝑅¯3R\bar{3}italic_R overÂŻ start_ARG 3 end_ARG structure Mu et al. (2022). This can be appreciated by the appearance of additional Bragg peaks along the L𝐿Litalic_L direction (see Fig. 2c). The right half of each panel shows the line cut along [0,2,L]02𝐿[0,2,L][ 0 , 2 , italic_L ] highlighting that (1) this crystal is of high quality with negligible mosaic and (2) the absence of any diffuse feature in the temperature range investigated. The correlation length ÎŸđœ‰\xiitalic_Ο derived from the full width half maximum (FWHM) of the (0,2,0)020(0,2,0)( 0 , 2 , 0 ) reflection at 200 K using Lorentzian line-shape is ∌450±10ⁱÅsimilar-toabsentplus-or-minus45010Å\rm\sim 450\pm 10~{}\AA∌ 450 ± 10 roman_Å, which is at the resolution limit of the neutron measurement. The peak broadens slightly at 100 K and suggests a reduction of the correlation length to about 350±10ⁱÅplus-or-minus35010Å\rm 350\pm 10~{}\AA350 ± 10 roman_Å. Figures 2(e)-2(f) show the neutron diffraction patterns at 200 K (above the structure transition) and 10 K (below the structure transition) for the type-II crystal with TN𝑁{}_{N}start_FLOATSUBSCRIPT italic_N end_FLOATSUBSCRIPT=6.5 K. The crystal clearly has a much larger mosaic and more interestingly, diffuse streaks are evident below the structural transition [Fig. 2(f)]. The correlation length is about 78±5ⁱÅplus-or-minus785Å\rm 78\pm 5~{}\AA78 ± 5 roman_Å at 200 K and reduces to 32±3ⁱÅplus-or-minus323Å\rm 32\pm 3~{}\AA32 ± 3 roman_Å at 10 K. Both values are much smaller compared to those for the crystal with TN𝑁{}_{N}start_FLOATSUBSCRIPT italic_N end_FLOATSUBSCRIPT=7.6 K. Despite the apparent difference in the correlation lengths, the local environment surrounding the Ru ions for both single crystals are very similar in the high temperature C⁹2/mđ¶2𝑚C2/mitalic_C 2 / italic_m phase. As can be determined from the structure information listed in Table I in Supporting Materials, those two crystals have the same lattice parameters, Cl-Ru-Cl bond angles, Ru-Cl bond lengths at 200 K. Given the surprisingly similar crystal structure and lack of any unidentified phase, the observed diffuse streaks in the crystal with TN𝑁{}_{N}start_FLOATSUBSCRIPT italic_N end_FLOATSUBSCRIPT=6.5 K are attributed to the stacking faults that are prevalent in van der Waals bonded layered materials. As discussed later, the stacking disorder can act as the pinning center preventing a uniform structure phase transition. Thus the structure transition occurs at a lower temperature. The inhomogeneous distribution of stacking disorder might be responsible for the sluggish transition occurring in a wider temperature range.

After warming above the structure transition, both crystals seem to get back to the original pristine state. In Supporting Materials, we show the neutron diffraction patterns of both crystals after warming up to 200 K from below the structure transition. Interestingly, the diffraction pattern and the correlation length are comparable before and after the thermal cycling. We also monitored possible effects of thermal cycling on the structure transition and magnetic order by measuring the temperature dependence of magnetization and specific heat. As shown in the Supporting Materials, the thermal cycling does not seem to affect the magnetic order after 10 thermal cycles.

Figure 3(a) shows the typical temperature dependence of thermal conductivity (Îșasubscript𝜅𝑎\kappa_{a}italic_Îș start_POSTSUBSCRIPT italic_a end_POSTSUBSCRIPT, thermal current along a-axis) of both single crystals below 20 K. The small difference in TN𝑁{}_{N}start_FLOATSUBSCRIPT italic_N end_FLOATSUBSCRIPT can be well resolved. Both samples show a recovery of heat transport upon cooling through TN𝑁{}_{N}start_FLOATSUBSCRIPT italic_N end_FLOATSUBSCRIPT and a peak around 5 K in the temperature dependence of thermal conductivity. This is consistent with previous reportsHentrich et al. (2018); Hirobe et al. (2017); Leahy et al. (2017); Kasahara et al. (2022); Lefrançois et al. (2022). The crystal with TN𝑁{}_{N}start_FLOATSUBSCRIPT italic_N end_FLOATSUBSCRIPT=7.6 K has a higher thermal conductivity below TN𝑁{}_{N}start_FLOATSUBSCRIPT italic_N end_FLOATSUBSCRIPT. This magnitude of thermal conductivity is higher than the threshold value for the observation of half-integer quantized thermal Hall effect Kasahara et al. (2022).

Figure 3(b) shows the oscillatory features in the field dependence of longitudinal thermal conductivity (Îșasubscript𝜅𝑎\kappa_{a}italic_Îș start_POSTSUBSCRIPT italic_a end_POSTSUBSCRIPT) at 2 K, shown as the scaled derivative (Îș−1superscript𝜅1\kappa^{-1}italic_Îș start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPTdÎș𝜅\kappaitalic_Îș/dB T−11{}^{-1}start_FLOATSUPERSCRIPT - 1 end_FLOATSUPERSCRIPT) of thermal conductivity with respect to applied magnetic field. This scaling procedure brings the oscillatory features of various samples with different thermal conductivities to the same scale and emphasizes the critical fields instead of absolute oscillatory amplitudes. We observed that the oscillatory features, especially those after the system has entered the field-induced quantum spin liquid phase ( 7.5 T), are very similar for both samples. This indicates that the oscillatory features in the field induced quantum spin liquid region are dominated by in-plane physics.

We also studied the thermal Hall effect of both crystals. Figures. 3c and d show a plateau region in both thermal resistivity and (scaled) conductivity measured at 5 K. The plateau spans the field region in which the field-induced disordered or quantum spin liquid phase dominates. The profile of the thermal Hall signal is better shown by resistivity. The scaled thermal Hall conductivity is highly sample dependent. We noticed in our sample screening that the plateau value could vary by a factor of 5 for different samples with different thermal conductivities. The crystal with TN𝑁{}_{N}start_FLOATSUBSCRIPT italic_N end_FLOATSUBSCRIPT=7.6 K has less stacking disorder and hence a higher thermal conductivity. This leads to a near half-integer quantized Îșx⁹ysubscriptđœ…đ‘„đ‘Š\kappa_{xy}italic_Îș start_POSTSUBSCRIPT italic_x italic_y end_POSTSUBSCRIPT/T. In contrast, Îșx⁹ysubscriptđœ…đ‘„đ‘Š\kappa_{xy}italic_Îș start_POSTSUBSCRIPT italic_x italic_y end_POSTSUBSCRIPT/T for the crystal with TN𝑁{}_{N}start_FLOATSUBSCRIPT italic_N end_FLOATSUBSCRIPT=6.5 K is further away from the half-integer quantized value as illustrated by the data shown in Fig. 3(d). A more detailed study of the anisotropic thermal transport properties and the origin of the oscillatory features will be reported separately.

III Discussions

The detailed physical properties of two Î±đ›Œ\alphaitalic_α-RuCl33{}_{3}start_FLOATSUBSCRIPT 3 end_FLOATSUBSCRIPT crystals were reported in this work as the representatives of two types of Î±đ›Œ\alphaitalic_α-RuCl33{}_{3}start_FLOATSUBSCRIPT 3 end_FLOATSUBSCRIPT crystals. Typically, type-I crystals have a TN𝑁{}_{N}start_FLOATSUBSCRIPT italic_N end_FLOATSUBSCRIPT above 7.2 K, whereas type-II crystals have a TN𝑁{}_{N}start_FLOATSUBSCRIPT italic_N end_FLOATSUBSCRIPT below 7 K. Specific heat results shown in Fig. 1 and Fig. S5 in Supporting Materials demonstrate that TN𝑁{}_{N}start_FLOATSUBSCRIPT italic_N end_FLOATSUBSCRIPT of our Î±đ›Œ\alphaitalic_α-RuCl33{}_{3}start_FLOATSUBSCRIPT 3 end_FLOATSUBSCRIPT crystals can range from 6 K to 7.6 K. Our WDS measurements confirmed that all crystals have the same atomic ratio and the nonstoichiometry or chemical composition should not be responsible for the variation in TN𝑁{}_{N}start_FLOATSUBSCRIPT italic_N end_FLOATSUBSCRIPT among different Î±đ›Œ\alphaitalic_α-RuCl33{}_{3}start_FLOATSUBSCRIPT 3 end_FLOATSUBSCRIPT crystals. Therefore, the differences in properties between type-I and type-II crystals may shed some light on the origin of TN𝑁{}_{N}start_FLOATSUBSCRIPT italic_N end_FLOATSUBSCRIPT variation in different Î±đ›Œ\alphaitalic_α-RuCl33{}_{3}start_FLOATSUBSCRIPT 3 end_FLOATSUBSCRIPT crystals.

A careful comparison between type-I and -II crystals indicates that stacking disorder causes a sluggish structure transition and leads to phase coexistence at low temperatures. Type-II crystals distinguish themselves from type-I crystals by showing a sluggish structure transition most likely originating from the stacking disorder. The larger mosaic (see Fig. 2e) suggests the presence of stacking disorder in type-II crystals. From magnetic and x-ray and neutron diffraction measurements of different crystals, the structure transition when being measured upon cooling occurs at a lower temperature with increasing mosaic, and two-phase coexistence is always observed when the structure transition occurs below 120 K. The phase coexistence was observed at 20 K, the lowest temperature of our X-ray diffraction measurements. It will not be a surprise if the phase coexistence persists at even lower temperatures, for example, below TN𝑁{}_{N}start_FLOATSUBSCRIPT italic_N end_FLOATSUBSCRIPT. Our x-ray and neutron diffraction studies provide solid evidence for the phase coexistence but cannot tell the detailed concentration and distribution of those two phases or the stacking disorder. However, one can expect the stacking disorder acts as a pinning center for the sliding of RuCl33{}_{3}start_FLOATSUBSCRIPT 3 end_FLOATSUBSCRIPT honeycomb layers during the structure phase transition, leading to more complex stacking disorder as the temperature decreases. This expectation is consist with the observation of diffuse streaks for type-II crystals shown in Fig. 2(f). Such diffuse streaks were observed in previous studies and attributed to stacking disorder between RuCl33{}_{3}start_FLOATSUBSCRIPT 3 end_FLOATSUBSCRIPT honeycomb layers Kim et al. (2022); Johnson et al. (2015). The concentration and detailed distribution of stacking disorder in the starting crystals affect the mosaic, structure transition temperature, the concentration and distribution of both high temperature and low temperature phases below the structure transition. Figure S1 presents the x-ray diffraction results for crystals with TN𝑁{}_{N}start_FLOATSUBSCRIPT italic_N end_FLOATSUBSCRIPT near 7.1 K, which highlights the effect of the concentration and detailed distribution of stacking disorder.

The neutron single crystal diffraction confirms that type-I crystals are of high quality as evidenced by a large correlation length beyond the resolution limit of our neutron measurement. However, this doesn’t mean that these crystals are free of stacking disorder. Previously, a scanning transmission electron microscope (STEM) study observed P⁹31⁹12𝑃subscript3112P3_{1}12italic_P 3 start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT 12 type stacking in Î±đ›Œ\alphaitalic_α-RuCl33{}_{3}start_FLOATSUBSCRIPT 3 end_FLOATSUBSCRIPT flakes from one crystal with C⁹2/mđ¶2𝑚C2/mitalic_C 2 / italic_m structure determined by neutron single crystal diffraction Ziatdinov et al. (2016). This result was reproduced in our recent STEM studies. STEM probes a small region of a thin flake, while neutron diffraction studies the whole crystal in neutron beam averaging over many similar and different regions. The discrepancy between STEM and neutron diffraction indicates the presence of stacking disorder in type-I crystals and the disorder can even exist in local regions in the form of other types of layer stacking with comparable energyCao et al. (2016). Nevertheless, the long coherence length implies the concentration of stacking disorder in high quality type-I crystals is low.

The stacking disorder can be responsible for the slightly suppressed TN𝑁{}_{N}start_FLOATSUBSCRIPT italic_N end_FLOATSUBSCRIPT. The phase coexistence or stacking disorder may affect both the interlayer and intralayer magnetic interactions. The interlayer magnetic interactions are disrupted by inhomogeneous layer spacing and exchange paths, while theoretical studies have shown that the intralayer exchange interactions are influenced by the stacking structure Kim and Kee (2016). Neutron single crystal diffraction at 200 K revealed the same local structure in both types of Î±đ›Œ\alphaitalic_α-RuCl33{}_{3}start_FLOATSUBSCRIPT 3 end_FLOATSUBSCRIPT crystals. However, whether the local structure similarity persists at low temperatures is unknown. Despite tremendous efforts, the low temperature structure is not yet determined due to the appearance of diffuse scattering below the structure transitionCao et al. (2016); Johnson et al. (2015). Solving the low temperature structure is beyond the scope of current effort. However, we would point out that this is essential to resolve possible local distortion of RuCl66{}_{6}start_FLOATSUBSCRIPT 6 end_FLOATSUBSCRIPT octahedra that determines the relative magnitude of those terms in magnetic Hamiltonian and also critical for a quantitative understanding of the suppression of TN𝑁{}_{N}start_FLOATSUBSCRIPT italic_N end_FLOATSUBSCRIPT. In addition, the phase coexistence scenario should be considered when solving the low temperature structure, depending on the quality of crystals employed.

The above discussion of how stacking disorder suppresses TN𝑁{}_{N}start_FLOATSUBSCRIPT italic_N end_FLOATSUBSCRIPT does not contradict the previous proposal that mechanical deformation can induce magnetic anomalies in the temperature range 10-14 KCao et al. (2016); Zhang et al. (2023). In previous studies, mechanical deformation was employed to create a large concentration of stacking disorder, which resulted in crystals with one single magnetic order at 14 K. However, the concentration of stacking disorder in both types of crystals studied in this work is much lower and this is especially true to type-I crystals. On the other hand, the suppression of TN𝑁{}_{N}start_FLOATSUBSCRIPT italic_N end_FLOATSUBSCRIPT from 7 K and the appearance of weak magnetic anomalies in the temperature range 10-14 K seem to be strongly correlated and may have the same origin. The Supporting Materials present specific heat data for more samples, with separate panels highlighting the features below and above 8 K. For those crystals with suppressed TN𝑁{}_{N}start_FLOATSUBSCRIPT italic_N end_FLOATSUBSCRIPT (for example lower than 7 K) or with more than one lambda type anomaly, some weak anomalies can be observed above 8 K. All the observations suggest that a small amount of stacking disorder suppresses the long range magnetic order with TN𝑁{}_{N}start_FLOATSUBSCRIPT italic_N end_FLOATSUBSCRIPT around 7 K, and increasing the population of stacking faults leads to magnetic phases with TN𝑁{}_{N}start_FLOATSUBSCRIPT italic_N end_FLOATSUBSCRIPT above 10 K.

The stacking disorder is also responsible for the sample dependent thermal transport properties. Upon cooling through TN𝑁{}_{N}start_FLOATSUBSCRIPT italic_N end_FLOATSUBSCRIPT, type-I crystals show a stronger recovery of thermal conductivity (see Fig. 3a). The high thermal conductivity and TN𝑁{}_{N}start_FLOATSUBSCRIPT italic_N end_FLOATSUBSCRIPT of type-I crystals indicate the importance of interlayer coupling and interactions. As reported previously Kasahara et al. (2022); Yamashita et al. (2020), a large thermal conductivity is necessary for the observation of half-integer quantized thermal Hall effect. In Fig. 3(d), the crystal with TN𝑁{}_{N}start_FLOATSUBSCRIPT italic_N end_FLOATSUBSCRIPT=7.6 K shows a high thermal Hall conductivity Îșx⁹ysubscriptđœ…đ‘„đ‘Š\kappa_{xy}italic_Îș start_POSTSUBSCRIPT italic_x italic_y end_POSTSUBSCRIPT/T, closer to the half integer quantized value. It is interesting to note that half-integer quantized thermal Hall effect was recently observed for a deformed Î±đ›Œ\alphaitalic_α-RuCl33{}_{3}start_FLOATSUBSCRIPT 3 end_FLOATSUBSCRIPT crystal with TN𝑁{}_{N}start_FLOATSUBSCRIPT italic_N end_FLOATSUBSCRIPT=14 K which exhibits a high thermal Hall resistivity despite a low thermal conductivityZhang et al. (2023). These two studies suggest two different ways to observe thermal Hall conductivity at half-integer quantized value: using high quality samples with minimal amount of stacking disorder and high thermal conductivity, or using samples with a high thermal Hall resistivity. Obviously, thermal Hall conductivity can be rather sample dependent and the stacking disorder can have a dramatic effect on it. Despite the sample dependent thermal Hall conductivity, both types of crystals have the similar critical fields for oscillatory features in magnetothermal conductivity and also the same field range in which thermal Hall effect and oscillatory features in thermal conductivity are observed.

The correlation between stacking disorder, magnetic and thermal transport properties developed in this work suggests the following can be valid criteria when selecting crystals with minimal amount of stacking disorder: (1) the structure transition temperature measured upon cooling. This can be determined by measuring the temperature dependence of magnetization in the paramagnetic state, or by monitoring the evolution with temperature of an appropriate nuclear reflection by x-ray or neutron diffraction. (2) the magnetic ordering temperature determined by specific heat measurements. In addition to a well defined TN𝑁{}_{N}start_FLOATSUBSCRIPT italic_N end_FLOATSUBSCRIPT, specific heat measurements can also tell whether extra anomalies exist near TN𝑁{}_{N}start_FLOATSUBSCRIPT italic_N end_FLOATSUBSCRIPT (see Fig. S5). We noticed that the anisotropic magnetic susceptibility around TN𝑁{}_{N}start_FLOATSUBSCRIPT italic_N end_FLOATSUBSCRIPT is not that sensitive to the small variation of TN𝑁{}_{N}start_FLOATSUBSCRIPT italic_N end_FLOATSUBSCRIPT or the presence of multiple magnetic transitions as revealed by specific heat shown in Fig. S5. From the characterizations of different crystals in the last decade, crystals with minimal amount of stacking disorder should have a structure transition around 140 K when being measured upon cooling and one single magnetic anomaly around 7.6 K in specific heat.

IV Summary

In summary, we study the effects of small amount of stacking disorder on the structure, magnetic and thermal transport properties in Î±đ›Œ\alphaitalic_α-RuCl33{}_{3}start_FLOATSUBSCRIPT 3 end_FLOATSUBSCRIPT single crystals with TN𝑁{}_{N}start_FLOATSUBSCRIPT italic_N end_FLOATSUBSCRIPT varying from 6.0 K to 7.6 K. The amount of stacking disorder may not be enough to induce magnetic anomalies in the temperature range 10-14 K, but can still have a dramatic effect on the physical properties. The stacking disorder in as-grown crystals can suppress the structure transition temperature (upon cooling), the Neel temperature, and lattice thermal conductivity. The similar oscillatory field dependent thermal conductivity and plateau like feature in thermal Hall resistivity were observed in all Î±đ›Œ\alphaitalic_α-RuCl33{}_{3}start_FLOATSUBSCRIPT 3 end_FLOATSUBSCRIPT single crystals studied in this work despite the variation in TN𝑁{}_{N}start_FLOATSUBSCRIPT italic_N end_FLOATSUBSCRIPT. However, Î±đ›Œ\alphaitalic_α-RuCl33{}_{3}start_FLOATSUBSCRIPT 3 end_FLOATSUBSCRIPT single crystals with minimal amount of stacking disorder have a higher thermal conductivity, which pushes the thermal Hall conductivity to be closer to the half-integer quantized value. Our results demonstrate a strong correlation between the layer stacking, structure transition, magnetic and thermal transport properties and highlight the importance of interlayer coupling in Î±đ›Œ\alphaitalic_α-RuCl33{}_{3}start_FLOATSUBSCRIPT 3 end_FLOATSUBSCRIPT despite the weak van der Waals bonding. Our work also suggests well-defined criteria for selecting Î±đ›Œ\alphaitalic_α-RuCl33{}_{3}start_FLOATSUBSCRIPT 3 end_FLOATSUBSCRIPT crystals with minimal amount of stacking disorder: a structure transition around 140 K when being measured upon cooling and one single magnetic anomaly around 7.6 K in specific heat. Moreover, the presence of stacking disorder may be universal in all layered, cleavable magnets. The implications of this work can be extended to other compounds and help resolve and understand their intrinsic properties and the novel quantum phenomena.

V Acknowledgment

JY would thank discussions with Tom Berlijn, Huibo Cao, Hwan Do and Alan Tennant. The authors thank Michael Lance for WDS measurements. HZ, SN, MM, and JY were supported by the U.S. Department of Energy, Office of Science, National Quantum Information Science Research Centers, Quantum Science Center. QZ, DM, AM, HM, and BS were supported by the US Department of Energy, Office of Science, Basic Energy Sciences, Materials Sciences and Engineering Division. JC and MC were supported by an Early Career project supported by DOE Office of Science FWP ERKCZ55. A portion of this research used resources at the Spallation Neutron Source, a DOE Office of Science User Facility operated by the Oak Ridge National Laboratory.

This manuscript has been authored by UT-Battelle, LLC, under Contract No. DE-AC0500OR22725 with the U.S. Department of Energy. The United States Government retains and the publisher, by accepting the article for publication, acknowledges that the United States Government retains a non-exclusive, paid-up, irrevocable, world-wide license to publish or reproduce the published form of this manuscript, or allow others to do so, for the United States Government purposes. The Department of Energy will provide public access to these results of federally sponsored research in accordance with the DOE Public Access Plan (http://energy.gov/ downloads/doe-public-access-plan).

VI references

References

  • Plumb et al. (2014) KW Plumb, JP Clancy, LJ Sandilands, V Vijay Shankar, YF Hu, KS Burch, Hae-Young Kee,  and Young-June Kim,Â â€œÎ±đ›Œ\alphaitalic_α- rucl 3: A spin-orbit assisted mott insulator on a honeycomb lattice,” Phys. Rev. B 90, 041112 (2014).
  • Banerjee et al. (2016) A Banerjee, CA Bridges, J-Q Yan, AA Aczel, L Li, MB Stone, GE Granroth, MD Lumsden, Y Yiu, Johannes Knolle, et al., “Proximate kitaev quantum spin liquid behaviour in a honeycomb magnet,” Nat. Mater. 15, 733–740 (2016).
  • Huang et al. (2017) Bevin Huang, Genevieve Clark, EfrĂ©n Navarro-Moratalla, Dahlia R Klein, Ran Cheng, Kyle L Seyler, Ding Zhong, Emma Schmidgall, Michael A McGuire, David H Cobden, et al., “Layer-dependent ferromagnetism in a van der waals crystal down to the monolayer limit,” Nature 546, 270–273 (2017).
  • Chen et al. (2019) Weijong Chen, Zeyuan Sun, Zhongjie Wang, Lehua Gu, Xiaodong Xu, Shiwei Wu,  and Chunlei Gao, “Direct observation of van der waals stacking–dependent interlayer magnetism,” Science 366, 983–987 (2019).
  • Yan (2022) J-Q Yan, “Perspective–the elusive quantum anomalous hall effect in mnbi2te4: Materials,” ECS Journal of Solid State Science and Technology 11, 063007 (2022).
  • Mu et al. (2022) Sai Mu, Kiranmayi D Dixit, Xiaoping Wang, Douglas L Abernathy, Huibo Cao, Stephen E Nagler, Jiaqiang Yan, Paula Lampen-Kelley, David Mandrus, Carlos A Polanco, et al., “Role of the third dimension in searching for majorana fermions in Î±đ›Œ\alphaitalic_α- rucl 3 via phonons,” Physical Review Research 4, 013067 (2022).
  • Kubota et al. (2015) Yumi Kubota, Hidekazu Tanaka, Toshio Ono, Yasuo Narumi,  and Koichi Kindo, “Successive magnetic phase transitions in Î±đ›Œ\alphaitalic_α- rucl 3: Xy-like frustrated magnet on the honeycomb lattice,” Phys. Rev. B 91, 094422 (2015).
  • Johnson et al. (2015) Roger D Johnson, SC Williams, AA Haghighirad, J Singleton, V Zapf, P Manuel, II Mazin, Y Li, Harald Olaf Jeschke, R ValentĂ­, et al., “Monoclinic crystal structure of Î±đ›Œ\alphaitalic_α- rucl 3 and the zigzag antiferromagnetic ground state,” Physical Review B 92, 235119 (2015).
  • Sears et al. (2015) Jennifer A Sears, M Songvilay, KW Plumb, JP Clancy, Yiming Qiu, Yang Zhao, D Parshall,  and Young-June Kim, “Magnetic order in Î±đ›Œ\alphaitalic_α- rucl 3: A honeycomb-lattice quantum magnet with strong spin-orbit coupling,” Physical Review B 91, 144420 (2015).
  • Majumder et al. (2015) M Majumder, M Schmidt, Helge Rosner, AA Tsirlin, H Yasuoka,  and M Baenitz, “Anisotropic ru 3+ 4 d 5 magnetism in the Î±đ›Œ\alphaitalic_α- rucl 3 honeycomb system: Susceptibility, specific heat, and zero-field nmr,” Phys. Rev. B 91, 180401 (2015).
  • Cao et al. (2016) Huibo B Cao, A Banerjee, J-Q Yan, CA Bridges, MD Lumsden, DG Mandrus, DA Tennant, BC Chakoumakos,  and SE Nagler, “Low-temperature crystal and magnetic structure of Î±đ›Œ\alphaitalic_α- rucl 3,” Phys. Rev. B 93, 134423 (2016).
  • Kim et al. (2022) Subin Kim, Bo Yuan,  and Young-June Kim,Â â€œÎ±đ›Œ\alphaitalic_α-rucl3 and other kitaev materials,” APL Mater. 10, 080903 (2022).
  • Sears et al. (2017) Jennifer A Sears, Yang Zhao, Zhijun Xu, Jeffrey W Lynn,  and Young-June Kim, “Phase diagram of Î±đ›Œ\alphaitalic_α- rucl 3 in an in-plane magnetic field,” Physical Review B 95, 180411 (2017).
  • Do et al. (2017) Seung-Hwan Do, Sang-Youn Park, Junki Yoshitake, Joji Nasu, Yukitoshi Motome, Yong Seung Kwon, DT Adroja, DJ Voneshen, Kyoo Kim, T-H Jang, et al., “Majorana fermions in the kitaev quantum spin system Î±đ›Œ\alphaitalic_α-rucl3,” Nat. Phys. 13, 1079–1084 (2017).
  • Lee (2021) Patrick Lee, “Quantized (or not quantized) thermal hall effect and oscillations in the thermal conductivity in the kitaev spin liquid candidate rucl3,” Journal Club for Condensed Matter Physics , DOI: 10.36471/JCCM–November–2021–02 (2021).
  • Yokoi et al. (2021) T Yokoi, S Ma, Y Kasahara, S Kasahara, T Shibauchi, N Kurita, H Tanaka, J Nasu, Y Motome, C Hickey, et al., “Half-integer quantized anomalous thermal hall effect in the kitaev material candidate Î±đ›Œ\alphaitalic_α-rucl3,” Science 373, 568–572 (2021).
  • Kasahara et al. (2018) Y Kasahara, T Ohnishi, Y Mizukami, O Tanaka, Sixiao Ma, K Sugii, N Kurita, H Tanaka, J Nasu, Y Motome, et al., “Majorana quantization and half-integer thermal quantum hall effect in a kitaev spin liquid,” Nature 559, 227–231 (2018).
  • Yamashita et al. (2020) M Yamashita, J Gouchi, Y Uwatoko, N Kurita,  and H Tanaka, “Sample dependence of half-integer quantized thermal hall effect in the kitaev spin-liquid candidate Î±đ›Œ\alphaitalic_α- rucl 3,” Physical Review B 102, 220404 (2020).
  • Lefrançois et al. (2022) É Lefrançois, G Grissonnanche, J Baglo, P Lampen-Kelley, J-Q Yan, C Balz, D Mandrus, SE Nagler, S Kim, Young-June Kim, et al., “Evidence of a phonon hall effect in the kitaev spin liquid candidate Î±đ›Œ\alphaitalic_α- rucl 3,” Physical Review X 12, 021025 (2022).
  • Czajka et al. (2022) Peter Czajka, Tong Gao, Max Hirschberger, Paula Lampen-Kelley, Arnab Banerjee, Nicholas Quirk, David G Mandrus, Stephen E Nagler,  and N Phuan Ong, “Planar thermal hall effect of topological bosons in the kitaev magnet Î±đ›Œ\alphaitalic_α-rucl3,” Nature Materials , 1–6 (2022).
  • Bruin et al. (2022a) JAN Bruin, RR Claus, Y Matsumoto, N Kurita, H Tanaka,  and H Takagi, “Robustness of the thermal hall effect close to half-quantization in Î±đ›Œ\alphaitalic_α-rucl3,” Nature Physics 18, 401–405 (2022a).
  • Kasahara et al. (2022) Y Kasahara, S Suetsugu, T Asaba, S Kasahara, T Shibauchi, N Kurita, H Tanaka,  and Y Matsuda, “Quantized and unquantized thermal hall conductance of the kitaev spin liquid candidate Î±đ›Œ\alphaitalic_α- rucl 3,” Physical Review B 106, L060410 (2022).
  • Zhang et al. (2023) Heda Zhang, Andrew May, Hu Miao, Brian Sales, David Mandrus, Stephen Nagler, Michael McGuire,  and Jiaqiang Yan, “The sample-dependent and sample-independent thermal transport properties of Î±đ›Œ\alphaitalic_α- rucl 3,” arXiv e-prints , arXiv–2303.02098 (2023).
  • Czajka et al. (2021) Peter Czajka, Tong Gao, Max Hirschberger, Paula Lampen-Kelley, Arnab Banerjee, Jiaqiang Yan, David G Mandrus, Stephen E Nagler,  and NP Ong, “Oscillations of the thermal conductivity in the spin-liquid state of Î±đ›Œ\alphaitalic_α-rucl3,” Nature Physics 17, 915–919 (2021).
  • Bruin et al. (2022b) JAN Bruin, RR Claus, Y Matsumoto, J Nuss, S Laha, BV Lotsch, N Kurita, H Tanaka,  and H Takagi, “Origin of oscillatory structures in the magnetothermal conductivity of the putative kitaev magnet Î±đ›Œ\alphaitalic_α-rucl3,” APL Materials 10, 090703 (2022b).
  • Lefrançois et al. (2023) Étienne Lefrançois, Jordan Baglo, Q BarthĂ©lemy, S Kim, Young-June Kim,  and Louis Taillefer, “Oscillations in the magnetothermal conductivity of Î±đ›Œ\alphaitalic_α- rucl 3: Evidence of transition anomalies,” Physical Review B 107, 064408 (2023).
  • Yan and McGuire (2022) J-Q Yan and MA McGuire, “Self-selecting vapor growth of transition metal halide single crystals,” arXiv e-prints , arXiv–2211 (2022).
  • Yan et al. (2017) J-Q Yan, Brian C Sales, Michael A Susner,  and Michael A Mcguire, “Flux growth in a horizontal configuration: an analog to vapor transport growth,” Physical Review Materials 1, 023402 (2017).
  • Ye et al. (2018) F. Ye, Y. Liu, R. Whitfield, R. Osborn,  and S. Rosenkranz, “Implementation of cross correlation for energy discrimination on the time-of-flight spectrometer CORELLI,” J. Appl. Crystallogr. 51, 315 (2018).
  • Hentrich et al. (2018) Richard Hentrich, Anja UB Wolter, Xenophon Zotos, Wolfram Brenig, Domenic Nowak, Anna Isaeva, Thomas Doert, Arnab Banerjee, Paula Lampen-Kelley, David G Mandrus, et al., “Unusual phonon heat transport in Î±đ›Œ\alphaitalic_α- rucl 3: strong spin-phonon scattering and field-induced spin gap,” Phys. Rev. Lett. 120, 117204 (2018).
  • Hirobe et al. (2017) Daichi Hirobe, Masahiro Sato, Yuki Shiomi, Hidekazu Tanaka,  and Eiji Saitoh, “Magnetic thermal conductivity far above the nĂ©el temperature in the kitaev-magnet candidate Î±đ›Œ\alphaitalic_α- rucl 3,” Physical Review B 95, 241112 (2017).
  • Leahy et al. (2017) Ian A Leahy, Christopher A Pocs, Peter E Siegfried, David Graf, S-H Do, Kwang-Yong Choi, B Normand,  and Minhyea Lee, ‘‘Anomalous thermal conductivity and magnetic torque response in the honeycomb magnet Î±đ›Œ\alphaitalic_α- rucl 3,” Physical review letters 118, 187203 (2017).
  • Ziatdinov et al. (2016) M Ziatdinov, Arnab Banerjee, A Maksov, Tom Berlijn, Wu Zhou, HB Cao, J-Q Yan, Craig A Bridges, DG Mandrus, Stephen E Nagler, et al., “Atomic-scale observation of structural and electronic orders in the layered compound Î±đ›Œ\alphaitalic_α-rucl3,” Nature communications 7, 13774 (2016).
  • Kim and Kee (2016) Heung-Sik Kim and Hae-Young Kee, “Crystal structure and magnetism in Î±đ›Œ\alphaitalic_α- rucl 3: An ab initio study,” Physical Review B 93, 155143 (2016).