Abstract
Defects are ubiquitous in solids, often introducing new properties that are absent in pristine materials. One of the opportunities offered by these crystal imperfections is an extrinsically induced long-range magnetic ordering,1 a long-time subject of theoretical investigations.1–3 Intrinsic, two-dimensional (2D) magnetic materials4–7 are attracting increasing attention for their unique properties including layer-dependent magnetism4 and electric field modulation6. Yet, inducing magnetism into otherwise non-magnetic 2D materials remains a challenge. Here, we investigate magneto-transport properties of ultrathin PtSe2 crystals and demonstrate unexpected magnetism. Our electrical measurements show the existence of either ferromagnetic or anti-ferromagnetic ground state orderings depending on the number of layers in this ultra-thin material. The change in the device resistance upon application of a ~ 25 mT magnetic field is as high as 400 Ω with a magnetoresistance (MR) value of 5%. Our first-principles calculations suggest that surface magnetism induced by the presence of Pt vacancies and the Ruderman-Kittel-Kasuya-Yosida (RKKY) exchange couplings across ultra-thin films of PtSe2 are responsible for the observed layer-dependent magnetism. Considering the existence of such unavoidable growth-related vacancies in 2D materials, 8,9 these findings can expand the range of 2D ferromagnets into materials that would otherwise be overlooked.
While intrinsically magnetic materials are rare in nature, long-range magnetism can be introduced into non-magnetic 2D materials through adatom intercalation,10 proximity coupling,11 and defect engineering12 thanks to their large surface area to volume ratios. The latter strategy has been predicted for several 2D metallic materials including graphene,2 but this has not been experimentally realized yet. Recently, PtSe2 has been considered to be an ideal platform for investigating defect-induced magnetization.13,14 First-principles calculations show that Se and especially Pt vacancies in monolayer PtSe2 have a strong influence on its electronic properties; a large magnetic moment of up to 6 µB is expected for single or double Pt vacancies. Such defects in 2D materials can either form naturally during the growth/annealing processes8 or they can be intentionally created post-synthesis, e.g., under electron beam irradiation.15 Point defects including Pt vacancies are present at the topmost layers in chemical vapor transport-grown PtSe2.9 Considering the existence of these unavoidable point defects in as-grown material9 and its metallic behavior above a critical thickness,16,17 PtSe2 holds a great promise for spintronic applications.
We obtain thin, ribbon-shaped PtSe2 flakes by mechanical exfoliation from chemical vapor transport -grown bulk crystals (HQ Graphene) onto a Si/SiO2 (270 nm) substrate. In total, ten devices were characterized with PtSe2 thickness varying between ~ 4.15 nm to ~ 14 nm, as determined by atomic force microscopy (AFM), allowing us to study the thickness dependence of the magnetism. Non-magnetic metallic Palladium (Pd) contacts (80 nm thick) were formed using electron beam lithography and electron beam evaporation techniques. These non-magnetic Pd contacts were fabricated to exclude any influence of the contact on the magnetic response of the device. Due to the ribbon-like shape of our crystals, devices were fabricated in the two-terminal geometry. An AFM image of a completed device is shown in Figure 1a. Longitudinal device resistance was characterized as a function of back-gate voltage (VBG), source-drain bias (VSD), magnetic field and temperature.
We first discuss the charge transport properties of a device based on a 5.2 nm thick PtSe2 crystal (device A). Figure 1b shows the gate-voltage (VBG) dependence of source-drain current ISD at fixed bias-voltage (VSD) values of 50, 100 and 200 mV where we observe nearly gate-independent transport characteristics. This indicates that 5.2 nm PtSe2 is metallic, in good agreement with recent findings. This is also confirmed by the observation of nearly linear ISD-VSD relation even at low temperatures (Figure 1c) and by studying the temperature dependence of the device resistance (Supplementary Information 1). Since semiconducting behavior was observed in thinner samples (it is also challenging to electrically probe magnetism in such resistive samples (Supplementary Information 2-3)), we focus on crystals thicker than 4.15 nm for the remainder of this manuscript.
We utilize magnetoresistance measurements as a sensing tool to characterize the magnetism in PtSe2. For this, we sweep an out-of-plane magnetic field (B) while recording the device resistance. As shown in Figure 2a, we observe a hysteresis loop with minima at ± 25 mT under backward and forward scans. Such hysteretic behavior and the presence of a minor loop (Supplementary Information 4) are the hallmarks of ferromagnetism and were previously observed in ferromagnetic metals such as Cobalt (Co) under the same measurement geometry.18 We extract a coercive field of ~ 25 mT which is also comparable to the one observed in Co wires.18 We observe a similar magnetoresistance response in five different samples with comparable coercive field values (Supplementary Information 5). We also note that, just like the total device resistance, the change in the device resistance under a magnetic field is almost insensitive to the VBG (Supplementary Information 6).
Next, we study the dependence of the change of device resistance (ΔR) on the bias voltage and the temperature. Figure 2b shows the VSD dependence of ΔR measured at 1.6 K. The signal increases from ΔR = 30 Ω to 400 Ω as the bias is reduced from 50 mV to 1 mV. The extracted MR percental variation is 5% at low biases (Supplementary Information 7). As shown in Figure 2b, dependencies of both ΔR and device resistance on VSD are similar and could be due to slightly non-linear charge injection at low biases. Its physical origin needs further investigation. Following, we present the temperature dependence of the magnetoresistance measured at a fixed bias of VSD = 5 mV (Figure 2c). ΔR has a very weak temperature dependence in the 1.5 K < T < 13 K range. Raising the temperature just slightly over 13 K quickly suppresses the magnetic response of the device. The temperature dependence of device resistance also shows a similar sudden drop when T is raised above 13 K, which could be related to the suppression of spin-dependent scattering. The resistance keeps increasing as the temperature is further increased, due to the metallic nature of PtSe2 (Supplementary Information 1).
In contrast to this hysteretic magneto-transport response, another subset containing five out of ten characterized devices shows different characteristics (Supplementary Information 8-9). In Figure 3a, we show the magnetic field dependence (forward and backward sweep directions) of the device resistance for one of these devices (device B), which exhibits plateaus with two different values. Here, we observe the lower plateau at low fields in the -30 mT < B < 30 mT range, whereas a sharp jump to the high plateau is observed for fields above |30| mT. The height of the jump in this device is ΔR ~ 2 Ω and depends on the VSD, as shown in Figure 3b. The width and height of this plateau are also strongly sensitive to the temperature, as the signal completely disappears above 5 K (Figure 3c). Here, we are confident that the transition between these plateaus is a result of the switching between antiferromagnetic ordering at low fields into a fully spin polarized state at higher fields due to the metamagnetic effect.19 Such magneto-transport response with two plateaus was previously observed in semiconducting bilayer CrI3-based tunneling devices.20,21 It was shown that each CrI3 layer has initially opposite spin polarization at low fields and that the application of a field above the coercive field reverses the magnetization of one of CrI3 layers causing both layers to have the same spin polarization. Additionally, the observation of very sharp switching between low and high plateaus indicates the out-of-plane direction magneto-crystalline anisotropy of PtSe2, in good agreement with recent predictions.14 If the anisotropy had in-plane components, the switching would occur gradually with the field.
After presenting the existence of both ferromagnetic and antiferromagnetic ordering in metallic PtSe2, we next check if there is a layer dependent magnetization. The natural way to investigate this dependence would be to compare magneto-transport response of mono and bilayer PtSe2. PtSe2 is however extremely resistive in this thickness range.16 Instead, we have characterized two adjacent devices prepared under identical conditions. As confirmed by the AFM scan, thickness of PtSe2 crystals in these devices differs by only one layer, as shown in Figure 4a-b, which allows us to investigate layer-dependent magnetism. While the device with a 6.45 nm thick (10-11 layers-device C) PtSe2 shows the characteristic response with two plateaus (Figure 4c), the device with an extra layer (device D) shows ferromagnetic response (Figure 4d). This observation indicates the layer-dependent magnetism in PtSe2. To further investigate this effect, we prepared another PtSe2 device (device E) containing additional one-layer thick fragments on the channel surface, as revealed by AFM imaging (Figure 4e). Magneto-transport response of this sample is very intriguing: both ferromagnetic and antiferromagnetic ordering coexist (Figure 4f). This measurement also allows us to directly conclude that the switching fields for these two contrasting magnetic orderings are different. Here, the strong layer-dependent relationship between different mechanisms could be stabilizing magnetic ordering differently and hence results in different switching fields. This observation is in good agreement with the different coercive fields probed by magneto-optic Kerr effect (MOKE) measurements for even and odd layers of insulating CrI3 magnets.4 Existence of either ferromagnetic or antiferromagnetic ground state orderings depending on the number of layers is intriguing.
In the following, we propose a theoretical picture to interpret the observed effects. As pristine PtSe2 does not present any intrinsic spin-polarization, neither in bulk nor in multilayer form, we suggest that the observed magnetism originates from lattice imperfections (Supplementary Information 10). The appropriate model describing the interaction between magnetic impurities mediated by conduction electrons is the well-established RKKY model.22–24 According to this, the dependence of magnetic exchange coupling J on the distance R reads as J(R) ≈ cos(2kFR)/Rα, with kF being the Fermi wavevector of the metal hosting the magnetic impurities. The Fermi surface of PtSe2 is composed of closed electron and hole pockets located at kz = 0 and electron pockets at kz = ±0.6 π/c.25,26 The latter give rise to an oscillating and power-law decaying contribution to J(R) in the direction normal to the PtSe2 layers with a period of approximately two layers, as schematically shown in Figure 5a. Hence, we see that adding or removing one layer from metallic PtSe2 is sufficient to change the sign of the magnetic exchange coupling, thereby realizing a thickness-dependent ferro- or anti-ferromagnetic configuration. This observation is also in a good agreement with the summary of devices measured at fixed bias of VSD = 50 mV and 100 mV. As shown in Fig. 5a-inset and Supplementary Information 12, we observe that MR decreases as crystal thickness increases, layer-dependent magnetism exist and magnetism is not detectable in thicker crystals.
Next, we step from model interactions to first-principles calculations performed on realistic models in order to suggest which type of defects may be responsible for the magnetic response observed in our samples. Earlier experimental investigations revealed the abundance of vacancy point defects in PtSe2 films.9 Motivated by this, we assess the impact of such defects on the magnetic properties of PtSe2 by introducing Se (VSe) and Pt (VPt) vacancy in both metallic multilayer and bulk PtSe2 models. In analogy with other transition metal dichalcogenides,27 we find that the VSe defect does not lead to magnetism (Supplementary Information 10). On the other hand, the introduction of the VPt defect at the surface of the multilayer model induces a local magnetic moment of ~1.2 μB per defect. We further observe that the spin density around the VPt defect mostly localizes on the neighboring selenium atoms within its first coordination shell, as shown in Figure 5b. Additionally, the investigation of the electronic density of states reported in Figure 5c indicates that the local magnetic moments stem from defect-induced spin-split states that emerge around the Fermi level. Remarkably, we find that the same defect does not lead to a local magnetic moment when forming in the bulk (Supplementary Information 10,12). Similarly, extrinsic adsorbents do not induce magnetism (Supplementary Information 11). It is worth noting in this context that PtSe2 is more prone to hosting metal atom vacancies than conventional, group VI dichalcogenides (e.g. MoS2) as a consequence of the substantially lower formation energy that this defect exhibits in the former material9 as compared to the latter.27
In contrast to recently discovered intrinsic ferromagnets,4–6 PtSe2 has some key advantages for immediate research: it is stable in its pristine form (Supplementary Information 13) without the need for an encapsulation process,17 it can be easily grown with various scalable methods,28 and readily isolated from its as-prepared substrate which makes it transferrable onto any arbitrary substrate.28 Compared to these intrinsic metallic ferromagnets, the currently observed critical temperature in PtSe2 is small. This could be in principle improved by control of the vacancy concentration (to increase the concentration and total magnitude of the resulting magnetic moment), control of crystal thickness (to control the sign of the coupling and the switching fields as discussed above) and Fermi level energy (to modulate the density of states at the Fermi level) in future experiments.
Contrary to the assumption that defects in 2D materials are detrimental for the overall crystal quality, defect engineering could be a key approach to enriching their functionality by inducing magnetism in air-stable materials. In the case of PtSe2, combining the defect-induced magnetism with unique thickness-dependent properties could have several applications for realizing spintronic devices using atomically thin materials.29–31 Thin semiconducting PtSe2 could be utilized as a magnetic substrate for proximity studies.32,33 On the other hand, thick metallic PtSe2 could be integrated into lateral and vertical spin torque devices,34 tunneling magneto-resistance devices,35 and spin valve devices36 for generating spin polarized charge carriers.
Methods
Device fabrication
PtSe2 crystals were obtained by mechanical exfoliation from bulk crystals (HQ Graphene) onto a doped Si substrate with 270 nm of SiO2. The substrate was imaged by using a color camera equipped optical microscope (Olympus BX51M). The thickness of selected crystals was determined by using AFM topography imaging (Asylum Research Cypher). Metallic contacts were prepared using e-beam lithography (Raith EBPG 5000+, 100 keV thermal field emission gun with a beam does of ~ 950 µC/cm2) and e-beam evaporation of Pd (Alliance-Concept EVA 760, 80 nm thick), without the use of any adhesion layers such as Cr or Ti.
Measurements
Cryogenic measurements were performed in an ICE Oxford liquid helium continuous flow cryo-magnetic system with a base temperature of ~ 1.5 K. Drain currents were measured using a Keithley Sourcemeter 2450, while a Keithley Sourcemeter 2400 was used for applying bias through the SiO2 gate dielectric. The drain voltage was varying between 1 mV to 1 V, while the source was grounded. Electronic transport measurements were carried out in a two-terminal configuration.
First-principles calculations
First-principles calculations have been performed within the generalized gradient approximation to density functional theory devised by Perdew, Burke and Ernzerhof.37 Geometry optimizations have been carried out with the VASP code,38,39 using a kinetic energy cutoff of 400 eV and k-meshes equivalent to 18 k-points per unit cell along each periodic direction. During geometry optimizations, lattice constants and interlayer spacing were constrained to their experimental values of 3.73 Å and 5.08 Å, respectively, while the tolerance on atomic forces was set to 0.02 eV/Å. The electronic density of states has been calculated with the SIESTA code,40 using a fine mesh equivalent to 120 k-points and a smearing of 0.02 eV. The adopted multilayer and bulk PtSe2 models consist of a a 6-layer thick 3 × 3 supercell and a periodic 3 × 3 × 2 supercell, respectively. In the case of multilayer models, the defects are placed in the topmost layer and a vacuum region of 15 Å was introduced to separate periodic images.
Supplementary Material
Supplementary information is available for this paper at https://www.nature.com/nnano/.
Acknowledgements
We acknowledge A. H. C. Neto for fruitful insights and discussions. We acknowledge the help of Z. Benes (CMI) with electron-beam lithography and K. Marinov for training on measurement set-up. A.A., A.C., D.U., and A.K. would like to acknowledge support by the European Research Council (ERC, Grant 682332), Swiss National Science Foundation (Grant 153298) and Marie Curie-Sklodowska COFUND (grant 665667). A.K. acknowledges funding from the European Union’s Horizon 2020 research and innovation programme under grant agreement No 785219 (Graphene Flagship). M.P. and O.V.Y. acknowledge support by the Swiss National Science Foundation (Grants 162612 and 172543). First-principles simulations have been carried out at the Swiss National Supercomputing Centre (CSCS) under the project s832.
Footnotes
Data availability. The data that support the findings of this study are available from the corresponding authors on reasonable request.
Authors Contributions. A.A. and A.K. designed the experiments. A.A. and A.C. fabricated the samples. A.A. performed transport measurements. A.C. and D.U. performed Raman spectroscopy measurements. M.P. and O.V.Y. devised theoretical models and performed first-principles calculations. A.A., M.P. and A.K. wrote the manuscript with input from A.C.
Competing interests
The authors declare no competing interests.
Reprints and permissions information is available at www.nature.com/reprints.
Publisher’s note: Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.
References
- 1.Esquinazi P, Hergert W, Spemann D, Setzer A, Ernst A. Defect-induced magnetism in solids. IEEE Trans Magn. 2013;49:4668–4674. [Google Scholar]
- 2.Yazyev OV, Helm L. Defect-induced magnetism in graphene. Phys Rev B. 2007;75 125408. [Google Scholar]
- 3.Osorio-Guillén J, Lany S, Barabash SV, Zunger A. Magnetism without magnetic ions: percolation, exchange, and formation energies of magnetism-promoting intrinsic defects in CaO. Phys Rev Lett. 2006;96 doi: 10.1103/PhysRevLett.96.107203. 107203. [DOI] [PubMed] [Google Scholar]
- 4.Huang B, et al. Layer-dependent ferromagnetism in a van der Waals crystal down to the monolayer limit. Nature. 2017;546:270–273. doi: 10.1038/nature22391. [DOI] [PubMed] [Google Scholar]
- 5.Gong C, et al. Discovery of intrinsic ferromagnetism in two-dimensional van der Waals crystals. Nature. 2017;546:265–269. doi: 10.1038/nature22060. [DOI] [PubMed] [Google Scholar]
- 6.Deng Y, et al. Gate-tunable room-temperature ferromagnetism in two-dimensional Fe3GeTe2. Nature. 2018;563:94–99. doi: 10.1038/s41586-018-0626-9. [DOI] [PubMed] [Google Scholar]
- 7.Bonilla M, et al. Strong room-temperature ferromagnetism in VSe2 monolayers on van der Waals substrates. Nat Nanotechnol. 2018;13:289–293. doi: 10.1038/s41565-018-0063-9. [DOI] [PubMed] [Google Scholar]
- 8.Zhou W, et al. Intrinsic structural defects in monolayer molybdenum disulfide. Nano Lett. 2013;13:2615–2622. doi: 10.1021/nl4007479. [DOI] [PubMed] [Google Scholar]
- 9.Zheng H, et al. Intrinsic point defects in ultrathin 1T-PtSe2 layers. arXiv:1808.04719 [cond-mat] 2018 [Google Scholar]
- 10.Hardy WJ, et al. Very large magnetoresistance in Fe0.28TaS2 single crystals. Phys Rev B. 2015;91 054426. [Google Scholar]
- 11.Wang Z, Tang C, Sachs R, Barlas Y, Shi J. Proximity-induced ferromagnetism in graphene revealed by the anomalous Hall effect. Phys Rev Lett. 2015;114 doi: 10.1103/PhysRevLett.114.016603. 016603. [DOI] [PubMed] [Google Scholar]
- 12.Guguchia Z, et al. Magnetism in semiconducting molybdenum dichalcogenides. Sci Adv. 2018;4 doi: 10.1126/sciadv.aat3672. eaat3672. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 13.Gao J, et al. Structure, stability, and kinetics of vacancy defects in monolayer PtSe2: A first-principles study. ACS Omega. 2017;2:8640–8648. doi: 10.1021/acsomega.7b01619. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 14.Zhang W, et al. Magnetism and magnetocrystalline anisotropy in single-layer PtSe2: Interplay between strain and vacancy. J Appl Phys. 2016;120 013904. [Google Scholar]
- 15.Krasheninnikov AV, Nordlund K. Ion and electron irradiation-induced effects in nanostructured materials. J Appl Phys. 2010;107 071301. [Google Scholar]
- 16.Ciarrocchi A, Avsar A, Ovchinnikov D, Kis A. Thickness-modulated metal-to-semiconductor transformation in a transition metal dichalcogenide. Nat Commun. 2018;9:919. doi: 10.1038/s41467-018-03436-0. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 17.Zhao Y, et al. High-electron-mobility and air-stable 2D layered PtSe2 FETs. Adv Mater. 2017;29 doi: 10.1002/adma.201604230. 1604230. [DOI] [PubMed] [Google Scholar]
- 18.Leven B, Dumpich G. Resistance behavior and magnetization reversal analysis of individual Co nanowires. Phys Rev B. 2005;71 064411. [Google Scholar]
- 19.Huang F, Kief MT, Mankey GJ, Willis RF. Magnetism in the few-monolayers limit: A surface magneto-optic Kerr-effect study of the magnetic behavior of ultrathin films of Co, Ni, and Co-Ni alloys on Cu(100) and Cu(111) Phys Rev B. 1994;49:3962–3971. doi: 10.1103/physrevb.49.3962. [DOI] [PubMed] [Google Scholar]
- 20.Song T, et al. Giant tunneling magnetoresistance in spin-filter van der Waals heterostructures. Science. 2018;360:1214–1218. doi: 10.1126/science.aar4851. [DOI] [PubMed] [Google Scholar]
- 21.Klein DR, et al. Probing magnetism in 2D van der Waals crystalline insulators via electron tunneling. Science. 2018;360:1218–1222. doi: 10.1126/science.aar3617. [DOI] [PubMed] [Google Scholar]
- 22.Ruderman MA, Kittel C. Indirect exchange coupling of nuclear magnetic moments by conduction electrons. Phys Rev. 1954;96:99–102. [Google Scholar]
- 23.Kasuya T. A Theory of metallic ferro- and antiferromagnetism on Zener’s model. Prog Theor Phys. 1956;16:45–57. [Google Scholar]
- 24.Yosida K. Magnetic properties of Cu-Mn alloys. Phys Rev. 1957;106:893–898. [Google Scholar]
- 25.Zhang K, et al. Experimental evidence for type-II Dirac semimetal in PtSe2. Phys Rev B. 2017;96 125102. [Google Scholar]
- 26.Clark OJ, et al. Dual quantum confinement and anisotropic spin splitting in the multivalley semimetal PtSe2. Phys Rev B. 2019;99 045438. [Google Scholar]
- 27.Pizzochero M, Yazyev OV. Point defects in the 1T’ and 2H phases of single-layer MoS2: A comparative first-principles study. Phys Rev B. 2017;96 245402. [Google Scholar]
- 28.Yu X, et al. Atomically thin noble metal dichalcogenide: a broadband mid-infrared semiconductor. Nat Commun. 2018;9:1545. doi: 10.1038/s41467-018-03935-0. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 29.Žutić I, Fabian J, Das Sarma S. Spintronics: Fundamentals and applications. Rev Mod Phys. 2004;76:323–410. [Google Scholar]
- 30.Han W, Kawakami RK, Gmitra M, Fabian J. Graphene spintronics. Nat Nanotechnol. 2014;9:794–807. doi: 10.1038/nnano.2014.214. [DOI] [PubMed] [Google Scholar]
- 31.Wen H, et al. Experimental demonstration of xor operation in graphene magnetologic gates at room temperature. Phys Rev Applied. 2016;5 044003. [Google Scholar]
- 32.Žutić I, Matos-Abiague A, Scharf B, Dery H, Belashchenko K. Proximitized materials. Materials Today. 2019;22:85–107. [Google Scholar]
- 33.Scharf B, Xu G, Matos-Abiague A, Žutić I. Magnetic Proximity effects in transition-metal dichalcogenides: Converting excitons. Phys Rev Lett. 2017;119 doi: 10.1103/PhysRevLett.119.127403. 127403. [DOI] [PubMed] [Google Scholar]
- 34.Yang T, Kimura T, Otani Y. Giant spin-accumulation signal and pure spin-current-induced reversible magnetization switching. Nat Phys. 2008;4:851–854. [Google Scholar]
- 35.Moodera JS, Kinder LR, Wong TM, Meservey R. Large Magnetoresistance at room temperature in ferromagnetic thin film tunnel junctions. Phys Rev Lett. 1995;74:3273–3276. doi: 10.1103/PhysRevLett.74.3273. [DOI] [PubMed] [Google Scholar]
- 36.Avsar A, et al. Gate-tunable black phosphorus spin valve with nanosecond spin lifetimes. Nat Phys. 2017;13:888–893. [Google Scholar]
- 37.Perdew JP, Burke K, Ernzerhof M. Generalized gradient approximation made simple. Phys Rev Lett. 1996;77:3865–3868. doi: 10.1103/PhysRevLett.77.3865. [DOI] [PubMed] [Google Scholar]
- 38.Kresse G, Furthmüller J. Efficient iterative schemes for ab initio total-energy calculations using a plane-wave basis set. Phys Rev B. 1996;54:11169–11186. doi: 10.1103/physrevb.54.11169. [DOI] [PubMed] [Google Scholar]
- 39.Kresse G, Joubert D. From ultrasoft pseudopotentials to the projector augmented-wave method. Phys Rev B. 1999;59:1758–1775. [Google Scholar]
- 40.Soler JM, et al. The SIESTA method for ab initio order- N materials simulation. J Phys: Condens Matter. 2002;14:2745. [Google Scholar]
Associated Data
This section collects any data citations, data availability statements, or supplementary materials included in this article.