Room temperature ferromagnetic ordering from bound magnetic polarons in rare-earth-doped ultrathin MoS2 nanosheets

Eng Tuan Poh a, Rajasree Das a, Manikandan Marimuthu a, Zheng Zhang b, Ramanathan Mahendiran *a and Chorng Haur Sow *a
aDepartment of Physics, National University of Singapore, 2 Science Drive 3, Singapore 117551. E-mail: phyrm@nus.edu.sg; physowch@nus.edu.sg; Tel: +(65) 65162616 Tel: (+65) 65162957
bInstitute of Materials Research and Engineering, A*STAR (Agency for Science, Technology and Research), Singapore 138634

Received 21st March 2025 , Accepted 27th June 2025

First published on 4th July 2025


Abstract

Dilute magnetic semiconductors (DMSs) endowed with room-temperature ferromagnetic capabilities that arise from bound magnetic polarons (BMPs) have been attractive for spintronics, information storage and magneto-optical applications. In particular, the inclusion of rare-earth lanthanide ions as magnetic dopants in semiconductors presents significant potential owing to their strong free-ion magnetic moments. Herein, we evaluate the influence of various rare-earth dopants (Tb3+, Er3+, and Eu3+) magnetically coupled to vacancy carrier spins in MoS2 nanosheets. The manifested ferromagnetic magnitudes adopt a trend that differs from that of free ion moments, with 5% Tb-doped MoS2 exhibiting maximum saturation magnetization. The characteristic dependence of sample magnetization on dopant concentration and stringent annealing conditions (defect concentration) justifies the BMP model in describing the system. The rational creation of these ferromagnetic nanosheets is expected to provide value in low-temperature, solution-based processing of spintronic components into monolithic integrated electronics and multifunctional devices. These findings are also expected to contribute to a comprehensive understanding of the design of rare-earth-doped transition metal dichalcogenides (TMDs) as DMSs for such future spintronic applications.


1. Introduction

Since the initial predictions of room temperature ferromagnetic properties in dilute magnetic semiconductors (DMSs),1,2 numerous trials across a range of dopant and host combinations have prompted continual research efforts in this domain. At low doping concentrations, the acquired ferromagnetic magnitudes in these systems have challenged the prevailing notions of magnetism. It is expected that short-range interactions in conventional super-exchange and double-exchange processes are incapable of explaining the long-range magnetic order from such a low-density distribution of magnetic inclusions. Only with the subsequent studies involving bound magnetic polarons (BMP),3–8 a reasonable model became available to account for the magnetic performance of DMSs involving wide bandgap semiconductor hosts embedded with transition metal (TM)9–12 and rare-earth dopants.

The subsequent surge in experimental explorations initiated many nano-systems inclusive of Mn-doped CdX (X = S, Se, and Te),13,14 CaMn7O12,15 Ce-doped BaTiO3,16 Co-doped Dy2O317 and M-doped ZnO (M = Yb,18 Mn,19 Tb,20 Eu,21 Co,22,23 Cu,24 Ni,25 and Sm26), each with magnetic characteristics befitting the BMP theory. Furthermore, with massive research attention invested in semiconducting 2D transition metal dichalcogenides (TMDs) in recent years, ensuing studies on doped TMDs with various nanostructure morphologies have expanded tremendously. The incorporation of various transition metal dopants such as Zn,27 Mn,28–31 Fe,32–34 Co,35–38 Cu39,40 and V41–46 into these TMD nanostructures have successfully induced localized magnetic behaviours that were experimentally verified against theoretical models. More importantly, subsequent theoretical predictions47,48 supported the thermodynamic substitution of lanthanide ions in MoS2, prompting targeted modulation of optical and magnetic properties through doping. Amidst the limited literature on lanthanide-doped MoS2, the majority revolves around Er and Eu dopants studied across a diverse range of optical (NIR-related),49–54 thermometric55 and photoelectric56,57 applications. Only a handful of studies have evaluated the ferromagnetic performances of MoS2 nanosheets arising from lanthanide doping, chiefly from Nd-,58 Dy-59 and Ho-induced60,61 BMP formation.

On account of the scarce availability of studies assessing the magnetic properties of rare-earth-doped MoS2 nanosheets, the current work presents a systematic analysis of magnetic influence from representative lanthanide ions (terbium (Tb), europium (Eu) and erbium (Er)) doped into MoS2 nanosheets. Leveraging on the expected abundance of vacancy defects in the nanosheets, we anticipate different magnetic anisotropies for the respective ions amid the crystal host lattice with varying extents of BMP interactions. Across low doping concentrations (up to 8%), the resultant ferromagnetic ordering was indeed dependent on nanosheet quality, dopant variant and doping concentrations, with exceptional fitting statistics in accordance with the BMP model. The acquired ferromagnetic transitions and Curie temperatures near room temperature further validated the potential of lanthanide-doped nanosheets as DMSs for room-temperature spintronics, in line with preceding theoretical expectations.

Furthermore, with the ever-increasing involvement of 2D materials in the development of advanced micro- and nanoelectronics owing to the diversified electronic properties of these materials,62–65 the production of solution-based TMD nanosheets has prompted the potential for mass-producing printed electronics via low-temperature processing technologies.66–69 It is imperative to evaluate the capability for inducing a magnetic response in MoS2 nanosheets through a convenient and scalable doping process to develop the technology as a magnetic or spintronic component to be integrated into multifunctional devices and monolithic integrated electronics.70–72 The advantage of functionalized nanosheet inks to produce printed layer stacks without the need for orthogonal considerations, coupled with the ease to alter the nanosheet properties flexibly, will encourage seamless integration of multifunctional and flexible printed electronics.

2. Experimental section

2.1. Chemicals

Ammonium thiomolybdate ((NH4)2MoS4, 99.97%), dimethylformamide (DMF, 99.8%), isopropanol (IPA, 99.5%) and rare-earth lanthanide chloride salts: terbium(III) chloride (TbCl3·6H2O, 99.9%), erbium(III) chloride (ErCl3·6H2O, 99.9%) and europium(III) chloride (ErCl3·6H2O, 99.9%) were purchased from Sigma Aldrich. All chemicals are used without further purification or processing steps.

2.2. Synthesis of lanthanide-doped MoS2 nanosheets

Dark red crystals of (NH4)2MoS4 (6.67 mg; 0.025 mmol) were sonicated in DMF (12 mL) until a homogeneous red solution was obtained. Aqueous stock solutions of various lanthanide salts (TbCl3, ErCl3, EuCl3) were prepared prior to use (2.5 mM), with the appropriate quantity (depending on the respective atomic doping%, detailed calculations and solution volumes utilized are presented in Table S1, ESI) withdrawn and added to the red thiomolybdate solution above. The resultant mix was sonicated before transferring into a Teflon-lined stainless-steel autoclave. The resultant suspension was heated in the oven at 200 °C for 15 hours. After the autoclave was allowed to cool naturally to room temperature, the dark precipitate obtained was isolated by centrifugation, followed by repeated procedures of redispersion and centrifugation washes in a DMF (4 mL)/IPA (11 mL) solvent pair. The resultant lanthanide-doped MoS3 nanosheets were redispersed in IPA (4 mL) and dried overnight in the oven at 80 °C. The subsequent annealing step was conducted in a tube furnace (850–900 °C, 2 hours) under different conditions (inert Ar; Ar with air) to yield crystalline lanthanide-doped MoS2 nanosheets.

2.3. Physical characterization

Raman spectra were obtained using a Renishaw inVia 2000 Micro-Raman spectrometer equipped with a 532 nm laser. SEM images with corresponding EDX spectra were acquired using a JEOL JSM-6700F field-emission scanning electron microscope (FESEM) operated at 15 kV, under Secondary Electron Imaging (SEI) with a chamber pressure of 9.63 × 10−5 Pa. The system is coupled with an Oxford Instruments X-MaxN 150 EDX detector. HRTEM and TEM measurements were acquired from a JEOL-JEM 3010 field emission transmission electron microscope operating at a 200 kV acceleration voltage under a chamber pressure of 3 × 10−8 Pa. XPS elemental analyses were conducted using a Thermo Fisher Scientific Theta Probe XPS system. Magnetic measurements were performed using a PPMS tagged to a Lakeshore model 7404 VSM.

3. Results and discussion

3.1. Characterization of lanthanide-doped nanosheets

The lanthanide-doped nanosheets herein were synthesized through a two-step hydrothermal-annealing process involving an intermediate amorphous MoS3 state. Spiked with small quantities of lanthanide chlorides (LnCl3; Ln = Tb, Eu, Er), ammonium thiomolybdate was broken down to yield Ln-doped MoS3 nanosheets, prior to a subsequent high temperature annealing to form crystalline MoS2. Fig. 1(a)–(c) presents the Raman signals of the as-produced MoS2 nanosheets characterized by the distinct E2g and A1g vibrational modes,73,74 in contrast with the relatively featureless profile of the pre-annealed amorphous MoS3 forms75 (Fig. 1(a)–(c) insets). This confirms the successful conversion under optimal annealing conditions with no alternate magnetic phase (1T-MoS2)76 or oxidized (MoO3 or MoOxSy) side products. With SEM-EDX, the nanosheet morphology of the MoS2 samples (Fig. 1(d)–(f) insets) was ascertained, with the successful incorporation of these selected f-block dopants (Tb3+, Er3+, Eu3+) verified via the presence of the corresponding elemental signals (Fig. 1(d)–(f)).
image file: d5tc01225b-f1.tif
Fig. 1 Raman and SEM-EDX characterization of Ln-doped MoS2 nanosheets. (a)–(c) Raman spectra of the respective Ln-doped MoS2 nanosheets acquired under 532 nm excitation: (a) Tb-doped MoS2, (b) Eu-doped MoS2 and (c) Er-doped MoS2. Inset: Corresponding Raman spectra of pre-formed Ln-doped MoS3 nanosheets prior to annealing: Tb-doped MoS3, Eu-doped MoS3 and Er-doped MoS3. (d)–(f) EDX elemental profile of Ln-doped MoS2 (Ln = (d) Tb, (e) Eu, (f) Er) at 5% atomic doping. Inset: Respective SEM images of the annealed nanosheets (scale bars = 10 μm).

Upon analyzing the HRTEM images (Fig. 2(a)–(c)), the nanosheet cluster was found to comprise multilayer stacks with an expected interlayer spacing of 0.7 nm, an observation coherent with the wide Raman inter-peak separation (E2g, A1g) typical of multilayer samples (Fig. 1(a)–(c)). Similar nanostructure morphologies and interlayer spacings were observed across various samples (Fig. S1, ESI), ensuring the consistency and reproducibility of the nanosheet synthesis process, along with the absence of obvious clustering of the rare-earth ions. Complementary to EDX, the indication of successful lanthanide doping was corroborated by detailed XPS analysis. As shown in Fig. 2(d) and (e), the Mo 3d and S 2p elemental signals displayed clean spectra with prominent Mo 3d5/2–Mo 3d3/2 and S 2p3/2–S 2p1/2 doublet pairs. The absence of MoS3-related peaks (S22− and bridging S peaks for S 2p; MoS3 peaks for Mo 3d; Fig. S2, ESI) affirmed a complete conversion after the annealing step, while the lack of any impurity-related signals assures the nanosheet quality. The absence of signals associated with any magnetically active Mo5+ species (spin 1/2, B.E. 234 eV) ensures no ferromagnetic interference from any Mo5+ components.77 Comparatively, the similar peak distributions and intensity ratios for the Mo 3d and S 2p spectra across all three samples (Ln = Tb, Eu, Er) affirm high sample consistency (Fig. S3, ESI).


image file: d5tc01225b-f2.tif
Fig. 2 TEM characterization and XPS analyses of Ln-doped MoS2 nanosheets. (a) Representative low-magnification TEM image of Ln-doped MoS2 nanosheet clusters. (b) Representative HRTEM of Ln-doped MoS2 nanosheets detailing the 0.7 nm interlayer stacking. (c) Corresponding SAED pattern. (d) and (e) Representative XPS spectra of (d) Mo 3d and (e) S 2p composition in the Ln-doped nanosheets. (f)–(i) Elemental signals for (f) Tb 3d, (g) Tb 4d, (h) Eu 3d and (i) Er 4d for the respective variants of Ln-doped nanosheets.

Meanwhile, the discernible lanthanide signals confirmed the presence of dopants (Tb3+, Eu3+, Er3+) within different chemical environments in the samples. Interestingly, terbium ions doped into the nanosheets adopt dual oxidation states, comprising Tb3+ and Tb4+. The XPS peaks were mapped in a similar correspondence to the mixed oxidation states in Tb4O7 (Tb3+ and Tb4+)78–80 and Tb2O3/TbO2 (Tb3+ and Tb4+)81–83 analogues (Fig. 2(f) and (g)). While it may be intuitively understandable that a redox conversion from Tb3+ ion to the Tb4+ state should allow greater thermodynamic stability for substitution at the Mo4+ site, discussions relating to such possibilities have not been typically addressed in preceding computational investigations and may warrant further studies. On the contrary, without the option of a stable +4 oxidation state, the erbium and europium dopants were unsurprisingly found to be prominent in the trivalent states (Fig. 2(h) and (i)), with the respective distinct signals of Eu 3d5/2 and Er 4d at 1136.0 eV and 168.5 eV. Furthermore, an additional peak for erbium at 172.6 eV can be assigned to the Er satellite signal (Fig. 2(i)). Coherently, the results of EDX and XPS measurements ascertained the identity of the nanosheets formed, along with the presence of the respective dopants, effectively substantiating the successful formation of the proposed nanosheet formulations.

3.2. Magnetic performance of lanthanide-doped nanosheets

The magnetic properties of MoS2 nanosheets doped with various lanthanide ions were evaluated by vibrating sample magnetometry (VSM). Standardized across 5% doping concentrations, the nanosheets doped with Tb3+, Eu3+ and Er3+ demonstrated differing magnetization dependence in response to the external magnetic field (Fig. 3(a)). The ferromagnetic signals exhibited a trend of increasing magnetization values, with the sample doped with Tb3+ having the strongest magnetization, followed by Eu3+, then Er3+. In contrast, pristine MoS2 (Fig. 3(a) inset) crafted via the same synthetic procedure presented a distinct diamagnetic signal. Herein, it is imperative to note that the diamagnetic signature in the pristine MoS2 nanosheets indicates that inherent defect-related magnetic origins (vacancy clusters, edges,84,85 and grain boundaries86,87) are significantly trivial in comparison to the effects brought about by rare-earth doping. Besides, given that most rare-earth oxides, even if present in trace quantities, are paramagnetic and not ferromagnetic at room temperature,88,89 the observed ferromagnetic ordering should signal the importance of dopants in altering the magnetic properties of the host. After brief omission of the ion paramagnetic contributions through linear extrapolation, the room temperature saturation magnetization at 10 kOe yielded an impressive 0.1963 emu g−1 for 5% Tb3+ doping, while that for 5% Eu and 5% Er attained 0.0400 and 0.0078 emu g−1, respectively.
image file: d5tc01225b-f3.tif
Fig. 3 Magnetic characterization of dopant variable Ln-doped MoS2 (Ln = Tb, Eu, and Er) nanosheets. (a) MH plot of Ln-doped MoS2 nanosheets with different dopant ions (Ln = Tb, Eu, and Er) at 5% atomic doping measured at room temperature after the correction of background paramagnetic contributions. Inset: Diamagnetic dominant signal from pristine MoS2 nanosheet reference. (b) Comparative MH plot of Ln-doped MoS2 nanosheets (Ln = Tb3+, Eu3+, and Er3+) produced under more stringent annealing conditions after correction against background paramagnetic signals. Comparison reveals a reproducible trend of the influence of dopant species and magnitude changes stemming from the stringency of annealing conditions. (c) Corresponding susceptibility (χH) plot at room temperature (300 K). (d) Enlarged low field region of the MH plot within ±300 Oe, demonstrating a minimal loop hysteresis in the measurements. (e)–(h) Proposed 4f electron distributions under crystal field (CF) splitting from a trigonal prismatic (D3h symmetry) arrangement of surrounding sulphur atoms for (e) Eu-doped MoS2, (f) Tb-doped MoS2 and (g) Er-doped MoS2 nanosheets comprising Ln substitutional doping at Mo atomic sites. (h) Proposed trend of effective magnetic moment (μeff) across 14 lanthanide ions, derived from spin–orbit interactions, under the crystal field influence proposed in (e)–(g).

Beyond dopant dependence, the magnetic capabilities of the samples were also assessed in relation to the quality of the nanosheets. In a parallel study, a separate batch of nanosheets produced under identical experimental conditions, except with greater stringency in the inert annealing process, yielded magnetic performance of uniformly reduced magnitudes (Fig. 3(b)). Under highly inert annealing conditions, the improved sample quality displayed inferior magnetic performance, indicating the critical role of potential defects, such as sulphur vacancies (VS), in contributing to the overall magnetic performance. Here, it is notable that despite the overall decrease in magnetization, the trend of dopant influence remains in the order of Tb3+ > Eu3+ > Er3+, ensuring the reliability of the results acquired. Unsurprisingly, the derived susceptibility (χH) affiliation displays an identical trend of sample magnetizability in the order of Tb3+ > Eu3+ > Er3+ (Fig. 3(c); derived from Fig. 3(a)), while a magnified examination of the MH curves under low-field (−300 Oe to 300 Oe) conditions unveils lower coercivity, asserting a soft ferromagnetic character to the nanosheets (Fig. 3(d); magnified from Fig. 3(a)).

Through a preliminary correlation of the magnetization dependence upon dopant variation and sample quality, it is possible to ascribe the ferromagnetic origin to the formation of bound magnetic polarons (BMPs). The BMPs are typically evaluated in various DMS systems,37,42,61 where polaron domains are created through ferromagnetic exchange between defect-localized carrier spins and vicinal magnetic ions. Since electron-bound VS defects have been known to be typically abundant in TMD nanostructures, the inclusion of lanthanide dopants is expected to produce BMPs from close radial proximity alignment of the lanthanide unpaired spins with the localized defect state. The hybridization of the dopant energy levels with the vacancy defect states effectively drives the ferromagnetic interactions between the Ln3+ dopant ions located within the polaron radius.7 Therefore, the overall magnetization is expected to be directly proportional to the density of defects, as well as the magnetic moment of the dopant. Since lanthanide ion substitution at the Mo4+ sites was previously determined to be the most thermodynamically favorable,48 a proposed distribution of the f-electrons under the context of a D3h trigonal prismatic crystal field90,91 of S atoms was applied for each of the dopant ions (Tb3+, Eu3+, Er3+) in Fig. 3(e)–(g). With a reasonable assignment of the crystal field splitting, the overall effective magnetic moment was determined, taking into account spin–orbit coupling from the spin and unquenched orbital momentum contributions.92–94 The resultant theoretical trend across the family of lanthanides (Fig. 3(h)) was coherent with the experimental results, yielding magnetization values in the order of MoS2:Tb3+ > MoS2:Eu3+ > MoS2:Er3+. Herein, it is notable that the magnetic moments determined from Fig. 3(h) should only adopt a qualitative relation to the actual moments measured; the theoretical derivation for magnetic moments (in μB/magnetic ion) should be less effective in quantifying non-conventional diluted magnetic systems involving coupling interactions between magnetic ions and localized defect states.

Owing to the strong ferromagnetic response of Tb-doped MoS2 nanosheets, the correlation between dopant concentration and magnetic performance was investigated based on the Tb doping concentration (at%). Across two separate batches of Tb-doped samples, similar trends were observed, with an increasing magnetic response up to an optimum at 5% Tb-doping, followed by a subsequent decline. The respective batches of nanosheets displayed magnetization trends in the order 5% > 2% > 8% (Fig. 4(a)) and 5% > 7% > 3% (Fig. 4(b)). Plotted together across a normalized scale, the comparative magnetization between the samples showed a reasonable correlation similar to doped magnetic systems—an increasing trend at low concentrations, followed by a tapering decline beyond the optimal doping concentration (Fig. 4(c)). Similar trends (with maximum magnetic performance at 5% doping) were also observed for Er- and Eu-doped samples when examined across 3%, 5% and 8% doping concentrations (Fig. S4, ESI). Naturally, with doping concentrations exceeding the optimum, the decreased inter-ionic distance between these embedded foreign ions, and/or dopant clustering beyond the solid solubility limit, promoted antiferromagnetic coupling between adjacent dopants over the BMP ferromagnetic mechanism. The ferromagnetic performance is gradually compromised by the thermodynamically favored antiferromagnetic coupling, resulting in the progressive weakening of the ferromagnetic signal. Nonetheless, with consistent assignment of elemental signals for the Tb content under XPS analyses (Fig. 4(d)–(f); full XPS data set under Fig. S5, ESI), the lack of any new chemical states or clustering signals assures doping below the host solubility limit; the decreased magnetic ordering therefore stems merely from antiferromagnetic coupling at reduced Tb–Tb inter-ionic distances.


image file: d5tc01225b-f4.tif
Fig. 4 Magnetic characterization of Tb–MoS2 nanosheets under variable atomic doping concentrations. (a) MH plot for single batch Tb-doped MoS2 nanosheets with 2%, 5% and 8% atomic doping concentrations after correction of background paramagnetic contributions. (b) MH plot for single batch Tb-doped MoS2 nanosheets with 3%, 5% and 7% atomic doping concentrations after the correction of background paramagnetic contributions. (c) Normalized comparison of saturation magnetization across samples of various doping concentrations, demonstrating a parabolic relation optimal at 5% Tb doping. Values standardized at H = 10 kOe, black and red plot points refer to sample batches presented in (a) and (b), respectively, normalized against the strongest magnetization in 5% Tb samples. (d)–(f) XPS characterization of Tb 3d signals from (d) 5% Tb-doped MoS2, (e) 7% Tb-doped MoS2, and (f) 8% Tb-doped MoS2 nanosheets.

In relation to temperature, the ferromagnetic feature of the 5% Tb-doped MoS2 sample was found to persist beyond room temperature. The unconventional shape of the MT plot reveals a rich complex interplay of multiple magnetic processes that can be deconvoluted in the subsequent analysis of individual MH plots at different temperature points (Fig. 5(a)). The processed plot of 1/χ against T revealed magnetic transitions with extrapolated x-intercepts yielding positive values (Fig. 5(b)), confirming the ferromagnetic nature of these processes. The derivative spectrum of dM/dT against T further details a Curie transition temperature (TC) at 308 K (Fig. 5(c)), reaffirming the capability of these doped nanosheets as room temperature DMSs. Through a series of MH spectra acquired at different temperature points, systematic changes in magnetic features can provide a holistic understanding of the magnetic transitions across the temperature scale. At higher temperatures (T > TC) (Fig. 5(d)), a mixed strength of paramagnetic and ferromagnetic signals exists, which increases marginally with decreasing temperature. Below TC, the system transitions into a ferromagnetic order, presenting a competitive ferromagnetic–antiferromagnetic interaction through a cusp feature between 250 K and 300 K, before stabilizing into a ferromagnetic state over a large temperature range down to 50 K (Fig. 5(a) and (e)). At 10 K, the spike in overall magnetization of the system is characterized by a paramagnetic or antiferromagnetic contribution in addition to the ferromagnetic feature, resulting in an unconventional upward curvature under both field-cooled (FC) and zero field-cooled (ZFC) conditions, with reduced bifurcation at low magnetization. Such a feature is expected to arise from an inherent low-temperature response stemming from the Tb dopants or the MoS2 host.77,85 Subtracting the background linear contribution through extrapolation, the resultant ferromagnetic signal was compared against the next acquired data point at 100 K, revealing minimal changes in the ferromagnetic feature with certainty that the overall increased magnetization stems purely from the additional linear magnetic component (Fig. 5(f)).


image file: d5tc01225b-f5.tif
Fig. 5 Temperature-dependent magnetic characterization of 5% Tb-doped MoS2 nanosheets. (a) MT relation for 5% Tb-doped MoS2 nanosheets under FC-warming, FC-cooling, FC-high field (1T) and ZFC measurement conditions. (b) 1/χ against T plot demonstrating the ferromagnetic characteristics of the sample through positive x-intercepts. (c) Derivative dM/dT against T relation in the determination of transition Curie temperature (TC) at approximately 308 K. (d) MH characteristics at high temperature domains of 325 K, 350 K and 390 K. (e) MH characteristics across the low-intermediate temperature range from 10 K to 325 K. (f) MH characteristics at low temperatures (10 K, 100 K) with processed plot (10 K) upon paramagnetic subtraction.

In comparison, the MT curves for the 2% and 7% Tb-doped samples were found to feature similar plot morphologies (Fig. 6(a), (b); red and black), with the 2% Tb-doped sample featuring a TC at approximately 319 K, while the 7% Tb-doped sample has a TC value of 300 K, along with a separate ferromagnetic transition at the onset of 259 K. (Fig. 6(a) and (b); blue). Once again, the consistent positive x-intercepts in the 1/χ plots reiterate the ferromagnetic transitions involved (Fig. 6(c) and (d)). Plotted upon a standard scale, the relation between the ascribed TC and Tb-dopant concentration presented a linear plot of inverse relation (Fig. S6, ESI).


image file: d5tc01225b-f6.tif
Fig. 6 Temperature-dependent magnetic characterization of 2% and 7% Tb-doped MoS2 nanosheets. (a) and (b) Respective MT curves (red and black) for 2% and 7% Tb-doped MoS2 nanosheets under FC-ZFC measurement conditions, overlaid with the derivative dM/dT against T (blue) to determine the transition Curie temperatures (TC) for the respective samples: 319 K for 2% Tb-doped sample (with side transition at 310 K); 300 K for the 7% Tb-doped sample (with side transition at 259 K). (c) and (d) Corresponding plots of 1/χ vs. T, demonstrating magnetic transition characteristics of the samples through x-intercepts.

3.3. Bound magnetic polaron modelling

In further evaluation of the BMP model in describing the observed ferromagnetic trends, the initial onsets of the MH plots were fitted in accordance with the following relation.
M = M0L(x) + χmH
Here, the first term models magnetic contributions from the BMPs, while the second term relates to paramagnetic contributions from the host matrix. The overall magnetization, M0 = Nms/ρ, where ms is the moment per BMP, N refers to the number of BMPs per unit volume and ρ is the density of the host matrix (ρ(MoS2) = 5.06 g cm−3). L(x) = coth(x) − 1/x is the Langevin function with x = Hmeff/kBT, where meff is the effective moment per BMP. Ignoring the interactions between the BMPs, an approximation for ms = meff allows for the determination of variables M0, meff and χm from the equation fit. As displayed in Fig. 7 and Fig. S7; ESI, the appropriate fits detailed by quality modelling statistics (Fig. S8 and S9; ESI reduced χ2 at 10−6 to 10−8) were obtained for the MH plots under the respective variations in dopant concentrations and temperature. The parameters derived from the various plots are summarized in Table 1.

image file: d5tc01225b-f7.tif
Fig. 7 BMP model fitting of the initial MH relation for the Tb-doped MoS2 sample at different atomic doping concentrations. (a) 2% Tb doping, (b) 5% Tb doping, (c) 7% Tb doping and (d) 8% Tb doping.
Table 1 Magnetization parameters (M0, meff, χm, N) derived from the BMP model fitting for Tb-doped samples under varied dopant concentrations (2%, 5%, 7%, and 8%) and the 5% Tb-doped sample under temperature variation
Dopant conc. (%Tb3+) M 0 (emu g−1) m eff (emu) χ m (emu g−1 Oe−1) N (cm−3)
2 0.056 1.0 × 10−16 3.8 × 10−6 2.8 × 1015
5 0.219 1.2 × 10−16 3.9 × 10−6 9.6 × 1015
7 0.013 2.0 × 10−16 6.4 × 10−6 3.3 × 1014
8 0.014 1.3 × 10−16 4.3 × 10−7 5.3 × 1014

Temperature (K) M 0 (emu g−1) m eff (emu) χ m (emu g−1 Oe−1) N (cm−3)
10 0.352 2.5 × 10−18 2.7 × 10−5 7.2 × 1017
100 0.277 2.8 × 10−17 1.4 × 10−6 5.0 × 1016
200 0.255 6.7 × 10−17 3.3 × 10−7 1.9 × 1016
275 0.268 1.2 × 10−16 1.6 × 10−6 1.1 × 1016
300 0.195 1.4 × 10−16 4.1 × 10−7 6.9 × 1015
325 0.065 1.0 × 10−16 2.7 × 10−6 3.3 × 1015
350 0.047 1.5 × 10−16 8.7 × 10−7 1.6 × 1015
390 0.020 9.8 × 10−17 5.4 × 10−7 1.0 × 1015


The extracted values for effective BMP moments were in the order of 10−16 emu, adopting a general trend of marginal increment with increasing dopant content. Precluding any counteracting antiferromagnetic contributions, the above-mentioned correlation is expected to present a reasonable trend, in which the increased dopant count homogeneously elevates the density of the magnetic dopant within the BMP radius, enhancing the magnetic moment of individual BMPs. Considering the batch samples, the relation of 2%, 5% and 8% doping demonstrated a linear trend against meff (Fig. 8(a)). In contrast, the sample densities of BMPs were found to be in the order 1014–1015 BMPs per cm3, with an inverse relation to the increasing dopant concentrations. The resultant trend exhibits the same curvature as the overall sample magnetization (Fig. 8(b)), with an optimal value at 5% doping, followed by a subsequent decline with higher concentrations. The matrix paramagnetic susceptibilities revealed low influence from the host material, yielding non-trending values of the order of 10−6 emu g−1 Oe.


image file: d5tc01225b-f8.tif
Fig. 8 Relation of the BMP model-derived parameters to Tb doping concentrations and temperature influence. (a) Dependence of effective magnetic moment per BMP (meff) on Tb doping concentrations, together with associated model fitting details. (b) BMP density (N) relation to Tb doping concentrations. (c) Variation in density of BMPs (N) within the sample in relation to temperature changes, together with associated model fitting details. (d) Variation in effective magnetic moment per BMP (meff) depending on the temperature measurement conditions, together with associated model fitting details.

With temperature variation, the samples accommodated a higher density of BMPs at lower temperatures, ranging from 1.0 × 1015 BMPs cm−3 at 390 K up to 7.2 × 1017 BMPs cm−3 at 10 K through a trend approximated by a (1 + x)−1.3 relation (Fig. 8(c)). It may be reasonable to accord the reduced BMP count towards greater instability when faced with increased thermal disruptions at higher temperatures. The polarons were ascribed with effective magnetic moments that generally increase with temperature for values below TC, ranging from 2.5 × 10−18 emu at 10 K and 1.4 × 10−16 emu at 300 K. The trend within this temperature range could be approximated by a cubic function (Fig. 8(d), trending statistics in Fig. S10, ESI). Beyond TC, the values do not follow a proper trend due to the interplay of additional magnetic (paramagnetic, antiferromagnetic) contributions.

4. Conclusions

Using different lanthanide dopants (Ln = Tb, Eu, Er), doped MoS2 nanosheets were synthesized to assess their various magnetic properties. The effective trend across consistently doped (5%) samples yielded magnetization in the order of Tb3+ > Eu3+ > Er3+, which is justified by the consolidation of possible mechanistic contributions stemming from crystal field influence and BMPs. Based on the Tb-doped MoS2 systems, the dependence of sample magnetic response on dopant concentration was evaluated, presenting an increasing trend at low concentration levels, followed by a subsequent decline beyond the optimal doping threshold. Such an observation could be attributed to a competing interplay between ferromagnetic and antiferromagnetic processes, resulting in acquired modulation of sample magnetization through doping proportions. The dependency of magnetization on doping ratios and annealing conditions, along with appropriate modelling of MH relations in the BMP model, served to strengthen the notion that BMP production contributes to room-temperature ferromagnetic properties. These findings are expected to provide further insights into the limited experimental studies on rare-earth-doped MoS2 systems, highlighting the potential of lanthanide-doped nanosheets for spintronic and magnetic applications.

Author contributions

Eng Tuan Poh – conceptualization, investigation, formal analysis, methodology, validation, data curation, visualization, writing – original draft; Rajasree Das – investigation, formal analysis, Manikandan Marimuthu – investigation, formal analysis; Zheng Zhang – investigation, formal analysis; Ramanathan Mahediran – funding acquisition, project administration, supervision, writing – review & editing; Chorng Haur Sow – funding acquisition, project administration, supervision, writing – review & editing.

Conflicts of interest

The authors declare no conflict of interest.

Data availability

The data supporting the article “Room Temperature Ferromagnetic Ordering from Bound Magnetic Polarons in Rare Earth Doped Ultrathin MoS2 Nanosheets” have been included as part of the ESI.

Acknowledgements

Our gratitude to Dr Chen Yun and Dr Zhao Fangfang from Bruker, along with Dr Wang Xinyun, for advice on magnetic force microscopy (MFM) as applied in this project.

References

  1. T. Dietl, H. Ohno and F. Matsukura, Phys. Rev. B: Condens. Matter Mater. Phys., 2001, 63, 195205,  DOI:10.1103/PhysRevB.63.195205.
  2. T. Dietl, H. Ohno, F. Matsukura, J. Cibert and D. Ferrand, Science, 2000, 287, 1019 CrossRef CAS PubMed.
  3. J. B. Torrance, M. W. Shafer and T. R. McGuire, Phys. Rev. Lett., 1972, 29, 1168 CrossRef CAS.
  4. M. Nawrocki, R. Planel, G. Fishman and R. Galazka, Phys. Rev. Lett., 1981, 46, 735 CrossRef CAS.
  5. T. Dietl and J. Spałek, Phys. Rev. B: Condens. Matter Mater. Phys., 1983, 28, 1548 CrossRef CAS.
  6. D. Heiman, P. A. Wolff and J. Warnock, Phys. Rev. B: Condens. Matter Mater. Phys., 1983, 27, 4848 CrossRef CAS.
  7. J. M. D. Coey, M. Venkatesan and C. B. Fitzgerald, Nat. Mater., 2005, 4, 173 CrossRef CAS PubMed.
  8. T. Dietl, Phys. Rev. Lett., 2023, 130, 086202 CrossRef CAS.
  9. M. Wu, X. Yao, Y. Hao, H. Dong, Y. Cheng, H. Liu, F. Lu, W. Wang, K. Cho and W. H. Wang, Phys. Lett. Sect. A Gen. At. Solid State Phys., 2018, 382, 111 CAS.
  10. C. Jia, B. Zhou, Q. Song, X. Zhang and Z. Jiang, RSC Adv., 2018, 8, 18837 RSC.
  11. A. A. Tedstone, D. J. Lewis and P. O’Brien, Chem. Mater., 2016, 28, 1965 CrossRef CAS.
  12. A. Yoshimura, N. Koratkar and V. Meunier, Nano Express, 2020, 1, 010008 CrossRef CAS.
  13. T. H. Nhung, R. Planel, C. Benoit, L. Guillaume and A. K. Bhattacharjee, Phys. Rev. B: Condens. Matter Mater. Phys., 1985, 31, 2388 CrossRef.
  14. P. A. Wolff and J. Warnock, J. Appl. Phys., 1984, 55, 2300 CrossRef CAS.
  15. S. Jaiswar and K. D. Mandal, J. Phys. Chem. C, 2017, 121, 19586 CrossRef CAS.
  16. P. Senthilkumar, S. Dhanuskodi, J. Karthikeyan and P. Murugan, Phys. Chem. Chem. Phys., 2019, 21, 4032 RSC.
  17. A. Bandyopadhyay, S. Sutradhar, B. J. Sarkar, A. K. Deb and P. K. Chakrabarti, Appl. Phys. Lett., 2012, 100, 252411,  DOI:10.1063/1.4729386.
  18. N. Bhakta, T. Inamori, R. Shirakami, Y. Tanioku, K. Yoshimura and P. K. Chakrabarti, Mater. Res. Bull., 2018, 104, 6 CrossRef CAS.
  19. S. Banerjee, K. Rajendran, N. Gayathri, M. Sardar, S. Senthilkumar and V. Sengodan, J. Appl. Phys., 2008, 104, 043913,  DOI:10.1063/1.2969945.
  20. W. Q. Zou, C. N. Ge, G. Venkataiah, H. L. Su, H. S. Hsu, J. C. A. Huang, X. C. Liu, F. M. Zhang and Y. W. Du, J. Appl. Phys., 2012, 111, 113704,  DOI:10.1063/1.4720381.
  21. A. Franco, H. V. S. Pessoni and M. P. Soares, J. Magn. Magn. Mater., 2014, 355, 325 CrossRef CAS.
  22. P. R. Chithira and T. Theresa John, J. Magn. Magn. Mater., 2020, 496, 165928 CrossRef CAS.
  23. L. T. Tseng, A. Suter, Y. R. Wang, F. X. Xiang, P. Bian, X. Ding, A. Tseng, H. L. Hu, H. M. Fan, R. K. Zheng, X. L. Wang, Z. Salman, T. Prokscha, K. Suzuki, R. Liu, S. Li, E. Morenzoni and J. B. Yi, Phys. Rev. B, 2017, 96, 104423 CrossRef.
  24. D. C. Agarwal, U. B. Singh, S. Gupta, R. Singhal, P. K. Kulriya, F. Singh, A. Tripathi, J. Singh, U. S. Joshi and D. K. Avasthi, Sci. Rep., 2019, 9, 6675 CrossRef CAS PubMed.
  25. P. Robkhob, I. M. Tang and S. Thongmee, Mater. Sci. Eng., B, 2020, 260, 114644 CrossRef CAS.
  26. P. Kaur, Kriti, Rahul, S. Chalotra, H. Kaur, A. Kandasami and D. P. Singh, Appl. Surf. Sci. Adv., 2021, 5, 100100 CrossRef.
  27. E. Z. Xu, H. M. Liu, K. Park, Z. Li, Y. Losovyj, M. Starr, M. Werbianskyj, H. A. Fertig and S. X. Zhang, Nanoscale, 2017, 9, 3576 RSC.
  28. R. Mishra, W. Zhou, S. J. Pennycook, S. T. Pantelides and J. C. Idrobo, Phys. Rev. B: Condens. Matter Mater. Phys., 2013, 88, 144409 CrossRef.
  29. J. Wang, F. Sun, S. Yang, Y. Li, C. Zhao, M. Xu, Y. Zhang and H. Zeng, Appl. Phys. Lett., 2016, 109, 092401 CrossRef.
  30. A. Ramasubramaniam and D. Naveh, Phys. Rev. B: Condens. Matter Mater. Phys., 2013, 87, 195201 CrossRef.
  31. X. Zhao, C. Xia, T. Wang and X. Dai, Solid State Commun., 2015, 220, 31 CrossRef CAS.
  32. S. Fu, K. Kang, K. Shayan, A. Yoshimura, S. Dadras, X. Wang, L. Zhang, S. Chen, N. Liu, A. Jindal, X. Li, A. N. Pasupathy, A. N. Vamivakas, V. Meunier, S. Strauf and E. H. Yang, Nat. Commun., 2020, 11, 6 CrossRef PubMed.
  33. H. Shu, P. Luo, P. Liang, D. Cao and X. Chen, ACS Appl. Mater. Interfaces, 2015, 7, 7534 CrossRef CAS.
  34. Z. Chen, J. He, P. Zhou, J. Na and L. Z. Sun, Comput. Mater. Sci., 2015, 110, 102 CrossRef CAS.
  35. Z. C. Xiang, Z. Zhang, X. J. Xu, Q. Zhang, Q. B. Wang and C. Yuan, Phys. Chem. Chem. Phys., 2015, 17, 15822 RSC.
  36. Y. Wang, S. Li and J. Yi, Sci. Rep., 2016, 6, 24153 CrossRef CAS PubMed.
  37. H. Duan, P. Guo, C. Wang, H. Tan, W. Hu, W. Yan, C. Ma, L. Cai, L. Song, W. Zhang, Z. Sun, L. Wang, W. Zhao, Y. Yin, X. Li and S. Wei, Nat. Commun., 2019, 10, 1584 CrossRef.
  38. J. Zhou, J. Lin, H. Sims, C. Jiang, C. Cong, J. A. Brehm, Z. Zhang, L. Niu, Y. Chen, Y. Zhou, Y. Wang, F. Liu, C. Zhu, T. Yu, K. Suenaga, R. Mishra, S. T. Pantelides, Z. G. Zhu, W. Gao, Z. Liu and W. Zhou, Adv. Mater., 2020, 32, 1906536 CrossRef CAS PubMed.
  39. W. S. Yun and J. D. Lee, Phys. Chem. Chem. Phys., 2014, 16, 8990 RSC.
  40. A. M. Hu, L. L. Wang, W. Z. Xiao and B. Meng, Phys. E, 2015, 73, 69 CrossRef CAS.
  41. X. Sun, J. Dai, Y. Guo, C. Wu, F. Hu, J. Zhao, X. Zeng and Y. Xie, Nanoscale, 2014, 6, 8359 RSC.
  42. W. Hu, H. Tan, H. Duan, G. Li, N. Li, Q. Ji, Y. Lu, Y. Wang, Z. Sun, F. Hu, C. Wang and W. Yan, ACS Appl. Mater. Interfaces, 2019, 11, 31155 CrossRef CAS.
  43. Y. T. H. Pham, M. Liu, V. O. Jimenez, Z. Yu, V. Kalappattil, F. Zhang, K. Wang, T. Williams, M. Terrones and M. H. Phan, Adv. Mater., 2020, 32, 2003607 CrossRef CAS.
  44. S. J. Yun, D. L. Duong, D. M. Ha, K. Singh, T. L. Phan, W. Choi, Y. M. Kim and Y. H. Lee, Adv. Sci., 2020, 7, 1903076 CrossRef CAS PubMed.
  45. F. Zhang, B. Zheng, A. Sebastian, D. H. Olson, M. Liu, K. Fujisawa, Y. T. H. Pham, V. O. Jimenez, V. Kalappattil, L. Miao, T. Zhang, R. Pendurthi, Y. Lei, A. L. Elias, Y. Wang, N. Alem, P. E. Hopkins, S. Das, V. H. Crespi, M. H. Phan and M. Terrones, Adv. Sci., 2020, 7, 2001174 CrossRef CAS.
  46. W. Hu, H. Tan, H. Duan, G. Li, N. Li, Q. Ji, Y. Lu, Y. Wang, Z. Sun, F. Hu, C. Wang and W. Yan, ACS Appl. Mater. Interfaces, 2019, 11, 31155 CrossRef CAS.
  47. A. Majid, A. Imtiaz and M. Yoshiya, J. Appl. Phys., 2016, 120, 142124,  DOI:10.1063/1.4963380.
  48. C. N. M. Ouma, S. Singh, K. O. Obodo, G. O. Amolo and A. H. Romero, Phys. Chem. Chem. Phys., 2017, 19, 25555 RSC.
  49. Y. Lyu, Z. Wu, W. F. Io and J. Hao, Appl. Phys. Lett., 2019, 115, 153105,  DOI:10.1063/1.5120173.
  50. G. Bai, S. Yuan, Y. Zhao, Z. Yang, S. Y. Choi, Y. Chai, S. F. Yu, S. P. Lau and J. Hao, Adv. Mater., 2016, 28, 7472 CrossRef CAS.
  51. Y. Yang, H. Pu, J. Di, Y. Zang, S. Zhang and C. Chen, Scr. Mater., 2018, 152, 64 CrossRef CAS.
  52. G. Bai, Z. Yang, H. Lin, W. Jie and J. Hao, Nanoscale, 2018, 10, 9261 RSC.
  53. G. Bai, Y. Lyu, Z. Wu, S. Xu and J. Hao, Sci. China Mater., 2020, 63, 575 CrossRef CAS.
  54. D. Xu, W. Chen, M. Zeng, H. Xue, Y. Chen, X. Sang, Y. Xiao, T. Zhang, R. R. Unocic, K. Xiao and L. Fu, Angew. Chem., Int. Ed., 2018, 130, 763 CrossRef.
  55. J. Liu, R. Van Deun and A. M. Kaczmarek, J. Mater. Chem. C, 2016, 4, 9937 RSC.
  56. M. Meng and X. Ma, Nanoscale Res. Lett., 2016, 11, 3 CrossRef.
  57. W. Shi, Z. Li, L. Wang, S. Wu, G. Zhang, M. Meng and X. Ma, Opt. Commun., 2018, 406, 50 CrossRef CAS.
  58. X. Ding, X. Cui, A. Sohail, P. P. Murmu, J. Kennedy, N. Bao, J. Ding, R. Liu, M. Peng, L. Wang, X. Chu, A. Vinu, S. P. Ringer and J. Yi, Adv. Quantum Technol., 2020, 4, 2000093 CrossRef.
  59. Q. Zhao, Q. Lu, Y. Liu and M. Zhang, Appl. Surf. Sci., 2019, 471, 118 CrossRef CAS.
  60. X. Chen and Z. Z. Lin, J. Nanopart. Res., 2018, 20, 129 CrossRef.
  61. Q. Zhao, C. Zhai, Q. Lu and M. Zhang, Phys. Chem. Chem. Phys., 2019, 21, 232 RSC.
  62. C. Liu, H. Chen, S. Wang, Q. Liu, Y. G. Jiang, D. W. Zhang, M. Liu and P. Zhou, Nat. Nanotechnol., 2020, 15, 545 CrossRef CAS.
  63. D. Akinwande, C. Huyghebaert, C. H. Wang, M. I. Serna, S. Goossens, L. J. Li, H. S. P. Wong and F. H. L. Koppens, Nature, 2019, 573, 507 CrossRef CAS PubMed.
  64. H. Wang, L. Yu, Y. H. Lee, Y. Shi, A. Hsu, M. L. Chin, L. J. Li, M. Dubey, J. Kong and T. Palacios, Nano Lett., 2012, 12, 4674 CrossRef CAS PubMed.
  65. G. Fiori, F. Bonaccorso, G. Iannaccone, T. Palacios, D. Neumaier, A. Seabaugh, S. K. Banerjee and L. Colombo, Nat. Nanotechnol., 2014, 9, 768 CrossRef CAS PubMed.
  66. T. Carey, S. Cacovich, G. Divitini, J. Ren, A. Mansouri, J. M. Kim, C. Wang, C. Ducati, R. Sordan and F. Torrisi, Nat. Commun., 2017, 8, 1202,  DOI:10.1038/s41467-017-01210-2.
  67. T. Carey, A. Arbab, L. Anzi, H. Bristow, F. Hui, S. Bohm, G. Wyatt-Moon, A. Flewitt, A. Wadsworth, N. Gasparini, J. M. Kim, M. Lanza, I. McCulloch, R. Sordan and F. Torrisi, Adv. Electron. Mater., 2021, 7, 2100112 CrossRef CAS.
  68. F. Torrisi, T. Hasan, W. Wu, Z. Sun, A. Lombardo, T. S. Kulmala, G. W. Hsieh, S. Jung, F. Bonaccorso, P. J. Paul, D. Chu and A. C. Ferrari, ACS Nano, 2012, 6, 2992 CrossRef CAS PubMed.
  69. S. Conti, L. Pimpolari, G. Calabrese, R. Worsley, S. Majee, D. K. Polyushkin, M. Paur, S. Pace, D. H. Keum, F. Fabbri, G. Iannaccone, M. Macucci, C. Coletti, T. Mueller, C. Casiraghi and G. Fiori, Nat. Commun., 2020, 11, 3566 CrossRef CAS PubMed.
  70. M. Sivan, Y. Li, H. Veluri, Y. Zhao, B. Tang, X. Wang, E. Zamburg, J. F. Leong, J. X. Niu, U. Chand and A. V. Y. Thean, Nat. Commun., 2019, 10, 5201 CrossRef PubMed.
  71. J. Jiang, K. Parto, W. Cao and K. Banerjee, IEEE J. Electron Devices Soc., 2019, 7, 878 CAS.
  72. S. Y. Seo, J. Park, J. Park, K. Song, S. Cha, S. Sim, S. Y. Choi, H. W. Yeom, H. Choi and M. H. Jo, Nat. Electron., 2018, 1, 512 CrossRef CAS.
  73. Z. Hu, Y. Bao, Z. Li, Y. Gong, R. Feng, Y. Xiao, X. Wu, Z. Zhang, X. Zhu, P. M. Ajayan and Z. Fang, Sci. Bull., 2017, 62, 16 CrossRef CAS PubMed.
  74. H. Li, Q. Zhang, C. C. R. Yap, B. K. Tay, T. H. T. Edwin, A. Olivier and D. Baillargeat, Adv. Funct. Mater., 2012, 22, 1385 CrossRef CAS.
  75. C. H. Chang and S. S. Chan, J. Catal., 1981, 72, 139 CrossRef CAS.
  76. L. Cai, J. He, Q. Liu, T. Yao, L. Chen, W. Yan, F. Hu, Y. Jiang, Y. Zhao, T. Hu, Z. Sun and S. Wei, J. Am. Chem. Soc., 2015, 137, 2622 CrossRef CAS PubMed.
  77. R. Sanikop and C. Sudakar, ACS Appl. Nano Mater., 2020, 3, 576 CrossRef CAS.
  78. C. Zhu, C. Lv, M. Jiang, J. Zhou, D. Li, X. Ma and D. Yang, Appl. Phys. Lett., 2016, 108, 051113,  DOI:10.1063/1.4941430.
  79. M. Glätzle, O. Janka, V. Svitlyk, D. Chernyshov, M. Bartsch, H. Zacharias, R. Pöttgen and H. Huppertz, Chem. – Eur. J., 2018, 24, 15236 CrossRef PubMed.
  80. D. D. Sarma and C. N. R. Rao, J. Electron Spectrosc. Relat. Phenom., 1980, 20, 25 CrossRef CAS.
  81. M. A. Farid, H. Zhang, A. Yang, G. Tian, M. Wu, G. Li, F. Liao and J. Lin, Eur. J. Inorg. Chem., 2017, 1427 CrossRef CAS.
  82. H. Li, W. Li, S. Gu, F. Wang and H. Zhou, Catal. Sci. Technol., 2016, 6, 3510 RSC.
  83. C. J. Lee, A. Sayal, S. Vashishtha and J. F. Weaver, Phys. Chem. Chem. Phys., 2019, 22, 379 RSC.
  84. W. Zhou, M. Chen, M. Guo, A. Hong, T. Yu, X. Luo, C. Yuan, W. Lei and S. Wang, Nano Lett., 2020, 20, 2923 CrossRef CAS PubMed.
  85. D. Gao, M. Si, J. Li, J. Zhang, Z. Zhang, Z. Yang and D. Xue, Nanoscale Res. Lett., 2013, 8, 129 CrossRef.
  86. N. Gao, Y. Guo, S. Zhou, Y. Bai and J. Zhao, J. Phys. Chem. C, 2017, 121, 12261 CrossRef CAS.
  87. Z. Zhang, X. Zou, V. H. Crespi and B. I. Yakobson, ACS Nano, 2013, 7, 10475 CrossRef CAS PubMed.
  88. J. B. Maccheskey, H. J. Williams, R. C. Sherwood and J. F. Potter, J. Chem. Phys., 1966, 44, 596 CrossRef.
  89. K. Trepka and Y. Tao, Appl. Phys. Lett., 2020, 117, 122410,  DOI:10.1063/5.0023466.
  90. M. Ju, X. Kuang, C. Lu, H. Li, J. Wang, C. Zhang, Y. Zhu and Y. Yeung, J. Mater. Chem. C, 2016, 4, 1988 RSC.
  91. W. W. Lukens, M. Speldrich, P. Yang, T. J. Duignan, J. Autschbach and P. Kögerler, Dalton Trans., 2016, 45, 11508 RSC.
  92. J. H. Van Vleck and A. Frank, Phys. Rev., 1929, 34, 1494 CrossRef CAS.
  93. B. Wang, S. Jiang, X. Wang and S. Gao, Sci. China, Ser. B: Chem., 2009, 52, 1739 CrossRef CAS.
  94. L. Sorace and D. Gatteschi, Lanthanides Actinides Mol. Magn., 2015, 2, 1 Search PubMed.

Footnote

Electronic supplementary information (ESI) available. See DOI: https://doi.org/10.1039/d5tc01225b

This journal is © The Royal Society of Chemistry 2025
Click here to see how this site uses Cookies. View our privacy policy here.