Next Article in Journal
Crystal Structure and Spectroscopic Analysis of 3-Diethoxyphosphoryl-28-[1-(1-deoxy-β-D-glucopyranosyl)-1H-1,2,3-triazol-4-yl]carbonylbetulin
Previous Article in Journal
Interfacial Characterization of Selective Laser Melting of a SS316L/NiTi Multi-Material with a High-Entropy Alloy Interlayer
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Ab Initio Magnetic Properties Simulation of Nanoparticles Based on Rare Earth Trifluorides REF3 (RE = Tb, Dy, Ho)

1
M.Auezov South Kazakhstan University, Shymkent 160012, Kazakhstan
2
Institute of Physics, Kazan Federal University, Kazan 420008, Russia
*
Author to whom correspondence should be addressed.
Crystals 2023, 13(10), 1487; https://doi.org/10.3390/cryst13101487
Submission received: 28 September 2023 / Revised: 8 October 2023 / Accepted: 10 October 2023 / Published: 12 October 2023

Abstract

:
Recently, nanoparticles based on rare earth fluorides have been widely investigated due to their possible medical and spintronic applications, where it is desirable to know the magnetic moment volume distribution. However, this feature is hard to find out from the experiments, so we employ ab initio spin-polarized calculations for this purpose in order to find out the tendencies and common features in three different compounds. In the present work, the nanoparticles of dipole magnets with different sizes, namely TbF3, DyF3 and HoF3, were simulated, and optimized structures were found. We present the optimized structures for the particles with various sizes, as well as for slabs constructed from same compounds. Lastly, for optimized geometries, the analysis of magnetic moments’ distribution over the nanoparticles and slabs volume was performed.

1. Introduction

In the recent years, great attention has been paid to various materials in the nanoparticle (NP) structure due to their unique properties, which arise thanks to their constrained geometry and which have many outstanding prospects both in science and technology [1,2,3,4]. Various materials were created and investigated in such a nano form and the study of rare earth fluoride NPs is gaining momentum thanks to the recent progress in technologies. The rare earth fluorides are known first of all for their magnetic properties. Apart from this, many of their practical applications are known today; for example, nanocrystalline fluoride powders are used as catalysts due to their significant specific area. Another important field is their use in medicine and biotechnology, namely the creation of nanoprobes with high biocompatibility [1], the development of radiation [2] and photodynamic [3] therapy, and nanothermometry [4].
More specifically, the trifluorides of rare earth elements are a class of materials with outstanding electronic, optical, and magnetic characteristics. For this reason, they are actively used in laser technology [5], as contrast agents in magnetic resonance imaging [6], and biological labels [7]. Nanocrystals of these compounds are used as model objects for both theoretical and practical research in various fields. The magnetic properties of rare earth trifluorides are of great importance when moving to the nanoscale due to the fact that their NPs exhibit a unique variety of magnetic properties, especially at low temperatures. In recent years, these properties have been actively studied by various methods, including computer simulation.
In order to underline the actuality of research objects, each material under investigation will be briefly discussed further separately in the introduction section.

1.1. Dipole Ferromagnet TbF3

Terbium trifluoride has a structure of the β-YF3 type, i.e., recrystallization is absent in it up to the melting point. The compound belongs to the orthorhombic symmetry space group Pnma ( D 2 h 16 ) [8] with lattice constants a = 0.6513 nm, b = 0.6949 nm and c = 0.4384 nm [9]. The corresponding crystal structure is shown in Figure 1.
According to the magnetization measurements and the behavior of the susceptibility, TbF3 is a dipole ferromagnet [10], the phase transition from the paramagnetic to the ferromagnetic state occurs at temperatures below TC = 3.95 K [10,11]. At temperatures T < TC, the magnetic moments of Tb3+ ions (≈9 μ B [12,13]) are ordered in two magnetically nonequivalent sublattices in the ac crystallographic plane at angles ϕ = ±28° to the a axis. Due to the large values of the magnetic moments, the magnetic ordering is induced mainly by the classical dipole–dipole interaction between terbium ions [8,10].
In Ref. [14], TbF3 NPs were used as contrast agents for magnetic resonance imaging under the conditions of an ultrahigh magnetic field (>3 T). The compound demonstrated an ability to attenuate X-rays and low toxicity compared to more popular contrast agents, which opens up great prospects for popularizing its use. In addition, TbF3 crystals can be used as a material for the manufacture of magneto-optical instruments and devices. The magnetic properties of the TbF3 compound are less studied than those of DyF3.

1.2. Dipole Ferromagnet DyF3

Dysprosium trifluoride DyF3 has a orthorhombic symmetry space group Pnma ( D 2 h 16 ) and crystal lattice parameters a = 0.6456 nm, b = 0.6909 nm and c = 0.4380 nm [15]. From the comparison of the calculated and measured magnetization values it was obtained that the main contribution to the magnetization is due to the dipole–dipole interaction [15]. Its crystal structure is same as for TbF3 and shown in Figure 1. The temperature of the phase transformation from paramagnetic to ferromagnetic state is TC = 2.55 K [16]. Magnetic ordering is induced mainly by dipole–dipole interactions between Dy3+ ions due to large values of magnetic moments [15].
Dysprosium fluoride DyF3 has unique magnetic properties at low temperatures. For instance, it was demonstrated that it can be used to improve the properties of Nd-Fe-B magnets [17]. Furthermore, dysprosium fluoride NPs have also been proposed for use as a contrast agent for high-field magnetic resonance imaging [6]. In a recent article [18], DyF3 NPs with sizes from 16 to 225 nm were synthesized by hydrothermal treatment in an autoclave, and it was also found that the Curie temperature of NP shifts towards lower temperatures with decreasing the size. Moreover, when the NP size is less than 5 nm, the transition to the magnetically ordered state is not observed. Additionally, the size effect was attentively investigated in another recent paper [19], where it was found that saturation of magnetization decreases with decreasing the particle size (16 nm–7 µm). The authors emphasised the need to consider both the nanoparticle core and shell, as well as the clustering process to describe the magnetic properties. Thus, experimentally the size effect onto the magnetic properties was clearly established for DyF3.

1.3. Dipole Ferromagnet HoF3

HoF3 belongs to the orthorhombic symmetry space group Pnma ( D 2 h 16 ) with lattice parameters a = 0.6404 nm, b = 0.6875 nm and c = 0.4379 nm [20]. The crystal structure is same as for previously described compounds and shown in Figure 1. It contains four Ho3+ and twelve F ions per unit cell. The compound exhibits a strong anisotropic susceptibility at low temperatures and it is close to the Ising system with two easy directions. The two lowest electronic states of HoF3 ions are singlets, clearly separated from the other energy levels [21]. The primary interaction between ions is the classical dipole interaction, which induces magnetic ordering at the Curie temperature TC = 0.53 K [22].
As for example of possible application, a facile strategy was proposed for preparing HoF3 NPs with good water solubility, low cytotoxicity and biocompatibility as a dual-modal contrast agent [23]. Authors discussed the prospects of those NPs for cancer diagnosis. They claimed that, grown by a facile one-pot solvothermal approach, these NPs could be used as both computer tomography (CT) and magnetic resonance imaging (MRI) which is also confirms the potential of Ho-based nanomaterials for bioapplication research.
To sum up, one of the problems in the physics of rare earth compounds is the effect of the size of fluoride NPs on their magnetic properties. In particular, the magnetic moments distribution within the NP has not yet been fully understood. Moreover, surface and geometry effects in magnetic core–shell NPs are attention-grabbing and a great challenge for the scientific community [24]. Also, one of the fundamental questions of the physics of nanosized objects is the question, “what size of a NP is sufficient for a formation of a magnetic ordering inside the particle?”. Furthermore, how are domains formed inside a NP? The main difficulty in the theoretical point of view is related, on the one side, with the restrictions of the precise ab initio calculations for a systems with a large number of atoms, and on the other side with inability of methods based on force-fields to conduct the magnetic calculations. Therefore, the purpose of this work is to make an attempt for ab initio calculations of magnetic NPs based on TbF3, DyF3 and HoF3 and establish the distribution of magnetic moments depending on the NP size and get close to understanding the issues raised.

2. Materials and Methods

Structural and magnetic properties calculations were realized as based on the density functional theory [25]. Exchange and correlation effects were accounted for using generalized gradient approximation (GGA-PBE) [26]. The Kohn–Sham equations [27] were solved using projectively extended wave potentials and wave functions [28]. All calculations were carried out using the VASP-6.3 (Vienna Ab-initio Simulation Package) program [29] built into the MedeA computational software [30]. The cutoff of the plane wave was taken to be 400 eV, the convergence criterion for atomic relaxation was 0.02 eV/Å, and the convergence condition for self-consistent calculations was the invariance of the total energy of the system with an accuracy of 10 5 eV. The Brillouin zones were sampled using Monkhorst–Pack grids [31,32,33], including 1 × 1 × 1 k-points for all NPs, and 5 × 1 × 7 for the slab. The k-mesh for bulk geometries was higher up to 5 × 5 × 7. The Gaussian smearing was 0.05 eV.
The following procedure was used for the calculation simplification: the geometry optimization were performed within the non-magnetic approach in order to simplify and fasten the calculations, and then the electronic and magnetic properties were extracted within the spin-polarized approach. Additionally, the optimization was performed with semi-core f electrons being treated as core states despite being higher in energy than other valence states. Even though the procedure is a brutal approximation, we checked for smaller particles to ensure that involving spin multiplicities does not affect the final result significantly.
Finally, in this work the modelling cells were created by spherical-shaped NPs based on an existing periodic model (shown in Figure 1) and surrounded by a thick-enough vacuum region to prevent the interactions with periodic copies. The analyzed surfaces were y-cut three and four unit cell slabs also surrounded by vacuum region in order to simulate the boundary of real particles with ≈2.0, 2.7 and 3.4 nm size.

3. Results

3.1. Bulk TbF3, DyF3 and HoF3

The magnetic calculations were performed within the spin-polarized approach and will be presented for each considered size separately.But, at the fist step, the bulk components were fully optimized, lattice parameters and magnetic moments were extracted and collected in Table 1 along with experimental data. The obtained lattice parameters agree well with reported experimental works even though the last were obtained at finite temperatures and at various conditions. However, the comparison of DFT and experimental magnetic moments is tricky, since the DFT assumes zero temperature. Additionally, DFT usually underestimates the magnetization values, since only the spin component is taken into account. That is why +U correction is usually used, or taking the spin-orbit coupling into account might be required. The first is associated with peculiarities while choosing the U term, the last one is time-consuming. However, the aim of the present research is a qualitative description of the magnetization distribution over the NPs, not quantitative so far. The spin-orbit calculations and the search for the best U term will be conducted in future.
In the further narration, the obtained magnetic moments of NPs will be compared with the bulk values.

3.2. Nanoparticle Geometry

In the following section the structural and magnetic properties will be presented for constructed stoichometric fully relaxed NPs. After the optimization, the RE-F distances were checked to ensure the reasonableness of the geometry relaxation. We found that RE-F distances do not change significantly while increasing the NP size for all investigated trifluorides.

3.2.1. 0.7 nm Particles

In order to keep the structures stoichiometric, the TbF3, DyF3 and HoF3 NPs were constructed as based on the bulk structures being 0.7 nm spheres while maintaining the stoichiometry of the material. The resulting structures have a chemical formula, (REF3)4, and are presented in Figure 2. Such NP size corresponds to very small structures containing only four rare earth ions as shown. As a result of optimization, structures of 0.7 nm nanoparticles take on the most symmetrical shape possible for such an initial configuration. The shape reminds the ring for all investigated materials with the difference being negligible, only in hundredths of ion-to-ion distances. In Figure 2, various sides of particles are shown.
As for the magnetic structure, the distribution is rather uniform and similar for all particles. However, the magnitude is different compared with the bulk materials. Furthermore, since the systems have symmetrical shapes with respect to the center, it would expected for them to have the same magnetic moments for all rare earth ions. On the contrary, each system has a difference in the neighboring rare earth ions of 0.656, 0.434 and 0.684 μ B for TbF3, DyF3 and HoF3, respectively. At the same time, the mean values are close to the bulk magnetic moment (collected in Table 2), being ≈0.02 μ B lower.

3.2.2. 0.8 nm Particles

The structures of 0.8 nm NPs of TbF3, DyF3 and HoF3 are shown in Figure 2. The resulting structures have a chemical formula, (REF3)6. The final nanostructures exhibit more atoms with a complex and notable structure with six rare earth ions. In particular, the rare earth ions locate the octachedral positions by poking out the F ions. All the particles are slightly elongated in one direction; the shape is ellipsoidal.
The complex geometry led to a redistribution of magnetization in accordance with the optimized, energetically favorable position of atoms. For all structures the distribution of magnetization is symmetrical with respect to the center of the NP. For TbF3, the distribution is more uniform (all ions have similar moments) and the mean value is close to the bulk. Moreover, two-side Tb ions have 6.0 μ B as in the bulk. On the contrary, for DyF3, the deviations for each ion from the bulk value are more prominent and equal up to 1.2 μ B . The magnetization ordering maintained as ferromagnetic for all structures here.

3.2.3. 1 nm Particles

The optimized structures of 1 nm NPs have a chemical formula, (REF3) 1 0 , and are collected in Figure 3. Obviously, it was found that the character of distribution is more complicated than in the previous case for smaller particles. The bigger particles contain 10 rare earth ions and have more complex structures and a more inhomogeneous distribution of magnetization.
Indeed, as depicted in Figure 4, the shape for all particles looks like an ellipsoid, but the TbF3 is least elongated, the rare earth ions are uniformly distributed and approximately equidistant from each other. On the contrary, the DyF3 particle has a distinct flattened grain shape. Still, as a result of optimization both Dy and F ions are uniformly distributed. The last HoF3 is similar to the TbF3 NP, also elongated but more flattened.
As for the magnetic moments distribution over the volume of TbF3 we observed rather homogeneous magnetization with a mean value for Tb of 5.916 μ B being close to the bulk value. However, the most distant-from-the-center ions received ≈4.95 μ B , whereas the center ions were more magnetized, up to ≈6.89 μ B .
The DyF3 nanostructure was found to be most elongated as a result of relaxation. The distribution of magnetic moments is complex and not symmetrical even though the structure was fully optimized. The magnetic moments vary from negative values of −6.76 μ B up to 6.82 μ B making the mean value smaller than in bulk. That means that there is one ion being antiferromagnetically coupled to other moments. Lastly, a similarity to the TbF3 case is that the center ions are most magnetized.
The last considered structure of HoF3 is ferromagnetic as well with more homogeneous magnetization than in the case of DyF3, but the values still vary significantly from 0 up to 6.83 μ B , however the mean magnetic moment is close to the bulk one. Again, it could be concluded that the most distant-from-the-center ions are less magnetized.

3.2.4. 1.6 nm Particles

The structures of constructed NP of TbF3, DyF3 and HoF3 with 1.6 nm size contain 176 atoms, so the optimization of such large systems is complicated within the DFT approach and it does not converge. Nevertheless, we decided to calculate the magnetization without optimization for a overview. Indeed, for TbF3, even for a non-optimized cell, we found the mean magnetic moment being close to the bulk value; in DyF3, Dy ions received the mean magnetic moment of 0.3 μ B which is less than the bulk value; whereas in the HoF3, the mean magnetic moment of Ho is 1 μ B is smaller than the bulk. Obviously, the optimization matters, so in order to overcome the restriction the slab geometries will be constructed and analyzed in the following section.

3.3. Slab Geometry

3.3.1. 2 nm Thick Slabs

As was pointed out in the previous subsection, the optimization process of particles with 1.6 nm size is associated with computational problems. However, in experiments, the synthesized NPs have an average dimension of ≈100–200 nm, and in order to approximate the real sizes we constructed the infinite slabs in order to investigate the distribution of magnetic moment within the thickness of it and to check the surface impact. Here, the crystals of TbF3, DyF3 and HoF3 were cut along xz-planes. The other cuts might be checked as well but we expect that, qualitatively, the situation would be the same.
The constructed and optimized ≈2 nm thick slabs are collected in Figure 5. In terms of the structural properties as a result of optimization, the atom positions did not change significantly in comparison with bulk and all three TbF3, DyF3 and HoF3 slabs look similar in terms of structure. At the same time the distribution of magnetic moments differs significantly. The TbF3 slab exhibited a core-type distribution with higher-magnetized ions surrounding the less-magnetized inner ions. The DyF3 slab is rather similar to TbF3 in character, with ≈5.9 μ B /Dy at the surface and 3.9 and 4.9 μ B for inner ions, respectively. Oppositely, the HoF3 slab had a negligible surface magnetization, while the inner part was quite highly magnetized at the second atomic layer up to 6.6 μ B , and the middle layers had 4.9 μ B /Ho, so there was an alternation.
At the same time, the calculated mean magnetic value was close to the bulk one for all three structures. Since the conclusions concerning the character of magnetization are not obvious here and no similarities are found, the cell was increased in thickness.

3.3.2. 2.7 nm Thick Slabs

At the next step, the computational cell of the slabs were increased by one unit cell of trifluoride. In this way, each cell represents four unit cells in the y-direction surrounded by a vacuum region, as depicted in Figure 6.
Here, for the TbF3 slab, the distribution slightly changed in comparison with the 2 nm thick slabs; no core-like distribution was observed here, it was rather uniform. On the contrary, the DyF3 slab possessed the least-magnetized middle atomic layers and up to 5.9 μ B /Dy surface layers. The last HoF3 slab, as in the previous case of 2 nm thick, had a very big alternation of up to 6.8 μ B /Ho and negligibly small ≈1 μ B /Ho atomic layers.

3.3.3. 3.4 nm Thick Slabs

Lastly, the cells were increased by one more unit cell, making the thickness of the slab 3.4 nm (Figure 7). This was a relatively thick slab, and the middle layers were supposed to be close to the bulk conditions. However, for the TbF3 slab, the distribution was similar to the previous case, but there were two atomic layers with 2.9 μ B /Tb and the middle layers were more magnetized than bulk. Overall the layers’ magnetization alternated with values being both lower and higher than bulk.
The DyF3 slab also possessed this alternation with slightly different character. As in the case of TbF3 slab, the surface had a high magnetization, but the third layer was higher, so all considered thicknesses of DyF3 slab demonstrated the alternation of higher and lower magnetization.
The HoF3 slab’s distribution was similar to the previous case of 2 nm, when the surface layers were ≈1 μ B /Ho and then the alternation was present. Similarly to DyF3, each investigated thickness of HoF3 possessed the alternation of more- and less-magnetized atomic layers but in the opposite way, with a low-magnetized surface.

4. Discussion

In this section we provide some common features observed along with corresponding discussions.
In terms of a shape the following was observed:
For 0.7–0.8 nm size NPs, all three investigated materials (TbF3, DyF3 and HoF3) have very similar atomic distributions with the only difference being in the atom-to-atom distance. When increasing the size of a NP, the shape changes from a sphere to a flattened elongated grain shape. This change becomes visible for a 1 nm size. The elongation was observed for all three compounds, but the DyF3 is the most stretched. This finding agrees with the experimental work of Ref. [6], where the morphology and crystal structure of HoF3 and DyF3 were analyzed by microscopy imaging. There, the size of synthesized NP was ≈70–110 nm, which is much higher than considered in our study, but the conclusions concerning the shape are similar. The authors found that the NPs have elongated shapes. The difference with our findings is that in the experimental research, the HoF3 and DyF3 NPs are nearly same shape and size of ≈70–110 nm length and ≈30–50 nm width depending on the initial compound used for synthesis, whereas the calculations revealed a slight difference between them, i.e., DyF3 is the more elongated. Finally, for the slabs, no structural difference was observed; the atom positions are similar to bulk ones.
In terms of magnetization the following features were found:
The distribution of magnetic moments within the NPs is mostly uniform and symmetrical with respect to the center of the particle. For the most symmetrical NP in terms of geometry, as for 0.8 nm HoF3, all rare earth ions receive similar magnetic moments.
The same is true for 1 nm, with HoF3 and TbF3 being less stretched than DyF3, and they have more homogeneous magnetization than DyF3. Indeed, the DyF3 NP being the most elongated has one Dy with zero magnetic moments and one with −6.76 μ B . That fact makes the average magnetization much less than in bulk, except for that case all the other mean magnetic moments for NPs of 0.7–1 nm size are close to the bulk value. The 1.6 nm NPs were not geometrically optimized, so we cannot take that into consideration. The summary of the obtained magnetic moments in comparison with bulk values is collected in Table 2.
As for the slab geometry, the distribution of magnetization is complex as well. Moreover, each of trifluoride slab demonstrates different behavior. The TbF3 is mostly homogeneous among the considered structures. TbF3 and DyF3 slabs have the surface with the most magnetized atomic layers. Unlike this, the HoF3 slabs of 2 and 3.4 nm thickness have negligible surface magnetization. Overall, each of the considered slab structures has an alternation of higher and lower magnetic moments. The biggest difference in the magnetization of neighboring layers was found for a DyF3 slab of 2.7 nm thickness.

5. Conclusions

Using the VASP package and DFT, calculations of the magnetic moments of TbF3, DyF3 and HoF3 NPs with sizes of 0.7–1.6 nm and 2–3.4 nm slabs were performed. An inhomogeneous distribution of magnetic moments and a different magnetic arrangements were observed depending on the NP size and slab thickness while maintaining the average value of the magnetic moment close to the bulk. Overall, each considered structure possesses the ferromagnetic arrangement. The slabs have a complex alternation of magnetized atomic layers.
The practical use of the rare earth trifluoride nanoparticles as contrast agents is limited by the toxicity of fluoride compounds to the human body. In this regard, they are coated with non-toxic polymers. The nature of the resulting coating depends on the distribution of magnetic moments and electric fields on the surface of nanoparticles. As a rule, the distribution of magnetic moments on the surface of nanoparticles is uneven, which affects the thickness of the coating and its quality. Simulations of the magnetic moments’ distribution over the nanoparticles volume and surface will allow experimentalists to assess the uniformity of the applied coating, and to propose control methods for uniform application of the polymer.
Further calculations might be conducted for the thicker slabs in order to understand the patterns of magnetization arrangement and approach the experimental sizes. Additionally, the spin–orbit approach might be used in order to investigate the spatial arrangement of each magnetic moment. Finally, other cuts of the crystals might be considered as well.

Author Contributions

Conceptualization, I.R.; methodology, O.N.; software, I.P.; validation, I.R., A.F. and O.N.; formal analysis, I.P., G.A. and P.S.; investigation, A.F.; resources, I.P.; data curation, A.F.; writing—original draft preparation, I.P.; writing—review and editing, O.N., P.S. and Z.S.; visualization, A.F. and G.B.; supervision, I.R.; project administration, P.S. and I.R.; funding acquisition, I.R. All authors have read and agreed to the published version of the manuscript.

Funding

This paper has been supported by the Kazan Federal University Strategic Academic Leadership Program (PRIORITY-2030).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data presented in this study are available on request from the corresponding author.

Acknowledgments

Computational resources were provided by the Laboratory for computer design of new materials and machine learning of Kazan Federal University.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Dong, H.; Du, S.-R.; Zheng, X.-Y.; Lyu, G.-M.; Sun, L.-D.; Li, L.-D.; Zhang, P.-Z.; Zhang, C.; Yan, C.-H. Lanthanide nanoparticles: From design toward bioimaging and therapy. Chem. Rev. 2015, 115, 10725–10815. [Google Scholar] [CrossRef] [PubMed]
  2. Du, Z.; Zhang, X.; Gou, Z.; Xie, J.; Dong, X.; Zhu, S.; Du, J.; Gu, Z. X-ray-Controlled Generation of Peroxynitrite Based on Nanosized LiLuF4:Ce3+ Scintillators and their Applications for Radiosensitization. J. Adv. Mater. 2018, 30, 1804046. [Google Scholar] [CrossRef]
  3. Cheng, T.; Riccardo, M.; Skripka, A.; Fiorenzo, V. Small and Bright Lithium-Based Upconverting Nanoparticles. J. Am. Chem. Soc. 2018, 140, 12890–12899. [Google Scholar] [CrossRef]
  4. Skripka, A.; Morinvil, A.; Matulionyte, M.; Cheng, T.; Vetrone, F. Advancing neodymium single-band nanothermometry. Nanoscale 2019, 11, 11322–11330. [Google Scholar] [CrossRef]
  5. Chen, G.; Qiu, H.; Fan, R.; Hao, S.; Tan, S.; Yang, C.; Han, G. Lanthanide-doped ultrasmall yttrium fluoride nanoparticles with enhanced multicolor upconversion photoluminescence. J. Mater. Chem. 2012, 22, 20190–20196. [Google Scholar] [CrossRef]
  6. González-Mancebo, D.; Becerro, A.I.; Rojas, T.C.; García-Martín, M.L.; de la Fuente, J.M.; Ocana, M. HoF3 and DyF3 Nanoparticles as Contrast Agents for High Field Magnetic Resonance Imaging. Part. Syst. Charact. 2017, 34, 1700116. [Google Scholar] [CrossRef]
  7. Wang, F.; Zhang, Y.; Fan, X.; Wang, M. One-pot synthesis of chitosan LaF3:Eu3+ nanocrystals for bio-applications. Nanotechnology 2006, 17, 1527–1532. [Google Scholar] [CrossRef]
  8. Holmes, L.; Guggenheim, H.J.; Hull, G.W. Spin-flip behavior in a ferromagnet, TbF3. Solid State Commun. 1970, 8, 2005–2007. [Google Scholar] [CrossRef]
  9. Wilmarth, W.R.; Begun, G.M.; Nave, S.E.; Peterson, J.R. The Raman Spectra of polycrystalline orthorhombic YF3, SmF3, HoF3, YbF3, and single crystal TbF3. J. Chem. Phys. 1988, 89, 711–715. [Google Scholar] [CrossRef]
  10. Holmes, L.M.; Guggenheim, H.J. Ferromagnetism in TbF3. J. Phys. 1971, 32, 501–502. [Google Scholar] [CrossRef]
  11. Brinkmann, J.; Courths, R.; Hufner, S.; Guggenheim, H.J. Magnetic phase diagram and critical and tricritical behaviour of TbF3. J. Magn. Magn. Mater. 1977, 6, 279–282. [Google Scholar] [CrossRef]
  12. Alakshin, E.M.; Klochkov, A.V.; Korableva, S.L.; Kuzmin, V.V.; Nuzhina, D.S.; Romanova, I.V.; Savinkov, A.V.; Tagirov, M.S. Magnetic properties of powders LiTbF4 and TbF3. Magn. Reson. Solids Electron. J. 2016, 18, 16204. [Google Scholar]
  13. Nishibu, S.; Nishio, T.; Yonezawa, S.; Kim, J.H.; Takashima, M.; Kikuchi, H.; Yamamoto, H. Preparation and magnetic properties of TbF3 containing oxide fluoride glasses. J. Fluor. Chem. 2006, 127, 821–823. [Google Scholar] [CrossRef]
  14. Zheng, X.; Wang, Y.; Sun, L.; Chen, N.; Li, L.; Shi, S.; Malaisamy, S.; Yan, C. TbF3 nanoparticles as dual-mode contrast agents for ultrahigh field magnetic resonance imaging and X-ray computed tomography. Nano Res. 2016, 9, 1135–1147. [Google Scholar] [CrossRef]
  15. Savinkov, A.V.; Korableva, S.L.; Rodionov, A.A.; Kurkin, I.N.; Malkin, B.Z.; Tagirov, M.S.; Suzuki, H.; Matsumoto, K.; Abe, S. Magnetic properties of Dy3+ ions and crystal field characterization in YF3:Dy3+ and DyF3 single crystals. J. Phys. Condens. Matter 2008, 20, 485220. [Google Scholar] [CrossRef]
  16. Holmes, L.M.; Hulliger, F.; Guggenheim, H.J.; Maita, J.P. Specific heat of a dipolar-coupled Ising ferromagnet LiTbF4. Phys. Lett. A 1974, 50, 163–164. [Google Scholar] [CrossRef]
  17. Sawatzki, S.; Dirba, I.; Schultz, L.; Gutfleisch, O. Electrical and magnetic properties of hot-deformed Nd-Fe-B magnets with diferent DyF3 additions. J. Appl. Phys. 2013, 114, 133902. [Google Scholar] [CrossRef]
  18. Alakshin, E.M.; Kondratyeva, E.I.; Garaeva, A.M.; Sakhatskii, A.; Likholetova, M.; Romanova, I.V.; Tagirov, M.S. Size effect of DyF3 nanoparticles on Curie temperature. Nanoscale 2022, 14, 11353–11358. [Google Scholar] [CrossRef]
  19. Garaeva, A.; Romanova, I.; Boltenkova, E.; Likholetova, M.; Tagirov, M.; Alakshin, E. A sudy of the size and clustering effects of DyF3 samples. Nanoscale 2023, 15, 12366–12374. [Google Scholar] [CrossRef]
  20. Zalkin, A.; Templeton, D.H. The crystal structures of YF3 and related compounds. J. Am. Chem. Soc. 1953, 75, 2453–2458. [Google Scholar] [CrossRef]
  21. Sharma, K.; Spedding, F.; Blinde, D. Optical and magnetic studies of HoF3. Phys. Rev. B 1981, 24, 82. [Google Scholar] [CrossRef]
  22. Leask, M.J.M.; Wells, M.R.; Ward, R.C.C.; Hayden, S.M.; Jensen, J. Magnetic excitations in the dipole-coupled singlet-singlet system HoF3. J. Phys. Condens. Matter 1994, 6, 505. [Google Scholar] [CrossRef]
  23. Zhang, T.; Deng, M.; Zhang, L.; Liu, Z.; Liu, Y.; Song, S.; Gong, T.; Yuan, Q. Facile Synthesis of Holmium-Based Nanoparticles as a CT and MRI Dual-Modal Imaging for Cancer Diagnosis. Front. Oncol. 2021, 11, 741383. [Google Scholar] [CrossRef] [PubMed]
  24. Evans, R.; Chantrell, R.; Chubykalo-Fesenko, O. Surface and interface effects in magnetic core–shell nanoparticles. MRS Bull. 2013, 38, 909–914. [Google Scholar] [CrossRef]
  25. Kresse, G.; Furthmüller, J. Efficiency of ab-initio total energy calculations for metals and semiconductors using a plane-wave basis set. Comput. Mater. Sci. 1996, 6, 15–50. [Google Scholar] [CrossRef]
  26. Perdew, J.P.; Burke, K.; Ernzerhof, M. Generalized gradient approximation made simple. Phys. Rev. Lett. 1996, 77, 3865. [Google Scholar] [CrossRef] [PubMed]
  27. Kohn, W.; Sham, L.J. Self-consistent equations including exchange and correlation effects. Phys. Rev. 1965, 140, A1133. [Google Scholar] [CrossRef]
  28. Blöchl, P.E. Projector augmented-wave method. Phys. Rev. 1994, 50, 17953. [Google Scholar] [CrossRef]
  29. Kresse, G.; Furthmüller, J. Efficient iterative schemes for ab initio total-energy calculations using a plane-wave basis set. Phys. Rev. 1996, 54, 11169. [Google Scholar] [CrossRef]
  30. MedeA, Version 3.7; MedeA Is a Registered Trademark of Materials Design, Inc.: San Diego, CA, USA. Available online: https://www.materialsdesign.com/ (accessed on 28 September 2023).
  31. Monkhorst, H.J.; Pack, J.D. Special points for Brillouin-zone integrations. Phys. Rev. 1976, 13, 125188. [Google Scholar] [CrossRef]
  32. Blöchl, P.E.; Jepsen, O.; Andersen, O.K. Improved tetrahedron method for Brillouin-zone integrations. Phys. Rev. 1994, 49, 16223. [Google Scholar] [CrossRef]
  33. Methfessel, M.; Paxton, A.T. High-precision sampling for Brillouin-zone integration in metals. Phys. Rev. 1989, 40, 3616. [Google Scholar] [CrossRef]
  34. Brown, P.J.; Forsyth, J.B.; Hansen, P.C.; Leask, M.J.; Ward, R.C.; Wells, M.R. Neutron diffraction determination of magnetic order in holmium trifluoride, HoF3. J. Phys. Condens. Matter 1990, 2, 4471. [Google Scholar] [CrossRef]
  35. Bleaney, B.; Gregg, J.F.; Hill, R.W.; Lazzouni, M.; Leask, M.J.M.; Wells, M.R. The magnetic properties of holmium trifluoride HoF3. J. Phys. C Solid State Phys. 1988, 21, 2721. [Google Scholar] [CrossRef]
Figure 1. The crystal structure of TbF 3 , DyF 3 and DyF 3 . Small light green spheres correspond to F ions, blue to Tb/Dy/Ho.
Figure 1. The crystal structure of TbF 3 , DyF 3 and DyF 3 . Small light green spheres correspond to F ions, blue to Tb/Dy/Ho.
Crystals 13 01487 g001
Figure 2. The crystal structure of NPs made of all considered materials with 0.7 nm size along with magnetic moments of all rare earth ions (Tb, Dy and Ho) in μ B . The connection between ions does not correspond to the chemical bonds and is used only for a better visualization. The shapes are the same for all compounds, and the pictures only show the various sides of the optimized NP.
Figure 2. The crystal structure of NPs made of all considered materials with 0.7 nm size along with magnetic moments of all rare earth ions (Tb, Dy and Ho) in μ B . The connection between ions does not correspond to the chemical bonds and is used only for a better visualization. The shapes are the same for all compounds, and the pictures only show the various sides of the optimized NP.
Crystals 13 01487 g002
Figure 3. The crystal structure of NPs made of all considered materials with 0.8 nm size along with magnetic moments of all rare earth ions (Tb, Dy and Ho) in μ B . The connection between ions does not correspond to the chemical bonds and is used only for a better visualization. The shapes are the same for all compounds; the pictures only show the various sides of the optimized NP.
Figure 3. The crystal structure of NPs made of all considered materials with 0.8 nm size along with magnetic moments of all rare earth ions (Tb, Dy and Ho) in μ B . The connection between ions does not correspond to the chemical bonds and is used only for a better visualization. The shapes are the same for all compounds; the pictures only show the various sides of the optimized NP.
Crystals 13 01487 g003
Figure 4. The crystal structure of NPs made of all considered materials with 1 nm size. The values on white balls correspond to the rare earth magnetic moments in μ B . The connection between ions does not correspond to the chemical bonds and used only for a better visualization.
Figure 4. The crystal structure of NPs made of all considered materials with 1 nm size. The values on white balls correspond to the rare earth magnetic moments in μ B . The connection between ions does not correspond to the chemical bonds and used only for a better visualization.
Crystals 13 01487 g004
Figure 5. The crystal structure of slabs made of TbF 3 , DyF 3 and HoF 3 with 2 nm size. The values on light blue balls correspond to the rare earth magnetic moments in μ B .
Figure 5. The crystal structure of slabs made of TbF 3 , DyF 3 and HoF 3 with 2 nm size. The values on light blue balls correspond to the rare earth magnetic moments in μ B .
Crystals 13 01487 g005
Figure 6. The crystal structure of slabs made of TbF 3 , DyF 3 and HoF 3 with 2.7 nm size. The values on light blue balls correspond to the rare earth magnetic moments in μ B .
Figure 6. The crystal structure of slabs made of TbF 3 , DyF 3 and HoF 3 with 2.7 nm size. The values on light blue balls correspond to the rare earth magnetic moments in μ B .
Crystals 13 01487 g006
Figure 7. The crystal structure of slabs made of TbF 3 , DyF 3 and HoF 3 with 3.4 nm size. The values on light blue balls correspond to the rare earth magnetic moments in μ B .
Figure 7. The crystal structure of slabs made of TbF 3 , DyF 3 and HoF 3 with 3.4 nm size. The values on light blue balls correspond to the rare earth magnetic moments in μ B .
Crystals 13 01487 g007
Table 1. The calculated lattice parameters and magnetic moments for studied bulk TbF3, DyF3 and HoF3 compounds along with available experimental data.
Table 1. The calculated lattice parameters and magnetic moments for studied bulk TbF3, DyF3 and HoF3 compounds along with available experimental data.
StructureTbF3DyF3HoF3
Lattice parameter
a b c, nm
Exp. lattice parameter
0.651 0.698 0.441
0.651 0.695 0.438 [9]
0.646 0.691 0.438
0.646 0.691 0.438 [15]
0.641 0.687 0.437
0.640 0.687 0.438 [34]
Magnetic moment
per Tb/Dy/Ho ion, μ B
Exp. magnetic moment
5.954
9.96 [12,13]
4.950
8 [15]
3.943
5.7 [35]
Table 2. The summarized collection of investigated structures and obtained magnetization. N is a number of rare earth ions, m is a magnetic moment. The N for the slabs correspond to the number of rare earth ions per computational unit cell (/u.c.).
Table 2. The summarized collection of investigated structures and obtained magnetization. N is a number of rare earth ions, m is a magnetic moment. The N for the slabs correspond to the number of rare earth ions per computational unit cell (/u.c.).
StructureSize, nmNMean m, μ B
TbF 3 bulk 5.95
0.745.94
0.865.92
1.0105.92
1.6445.90
2.012/u.c.5.94
2.716/u.c.5.93
3.420/u.c.5.91
DyF 3 bulk 4.95
0.744.93
0.864.93
1.0102.78
1.6444.6
2.012/u.c.4.94
2.716/u.c.4.95
3.420/u.c.4.93
HoF 3 bulk 3.94
0.743.92
0.863.9
1.0103.91
1.6442.85
2.012/u.c.3.95
2.716/u.c.3.92
3.420/u.c.3.94
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Saidakhmetov, P.; Piyanzina, I.; Faskhutdinova, A.; Nedopekin, O.; Adyrbekova, G.; Baiman, G.; Suyerkulova, Z.; Romanova, I. Ab Initio Magnetic Properties Simulation of Nanoparticles Based on Rare Earth Trifluorides REF3 (RE = Tb, Dy, Ho). Crystals 2023, 13, 1487. https://doi.org/10.3390/cryst13101487

AMA Style

Saidakhmetov P, Piyanzina I, Faskhutdinova A, Nedopekin O, Adyrbekova G, Baiman G, Suyerkulova Z, Romanova I. Ab Initio Magnetic Properties Simulation of Nanoparticles Based on Rare Earth Trifluorides REF3 (RE = Tb, Dy, Ho). Crystals. 2023; 13(10):1487. https://doi.org/10.3390/cryst13101487

Chicago/Turabian Style

Saidakhmetov, Pulat, Irina Piyanzina, Amina Faskhutdinova, Oleg Nedopekin, Gulmira Adyrbekova, Gulzagira Baiman, Zhamilya Suyerkulova, and Irina Romanova. 2023. "Ab Initio Magnetic Properties Simulation of Nanoparticles Based on Rare Earth Trifluorides REF3 (RE = Tb, Dy, Ho)" Crystals 13, no. 10: 1487. https://doi.org/10.3390/cryst13101487

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop