Valentína Berecováab,
Martin Friák
a,
Naděžda Pizúrová
a and
Jana Pavlů
*b
aInstitute of Physics of Materials, v. v. i., Czech Academy of Sciences, Žižkova 22, Brno, 616 00, Czech Republic
bMasaryk University, Faculty of Science, Department of Chemistry, Kotlářská 2, Brno, 611 37, Czech Republic. E-mail: houserova@chemi.muni.cz
First published on 28th August 2025
Structural and magnetic properties of ultra-small tetrahedron-shaped iron oxide nanoparticles were investigated using density functional theory. Tetrahedral and truncated tetrahedral models were considered in both non-functionalized form and with surfaces passivated by pseudo-hydrogen atoms. The focus on these two morphologies reflects their experimental relevance at this size scale and the feasibility of performing fully relaxed, atomistically resolved first-principles simulations. Moreover, a novel application of pseudo-hydrogen passivation to magnetic iron oxide nanoparticles is introduced as a practical strategy to probe intrinsic surface effects on magnetism while reducing artefacts from dangling bonds. Although these terminations are simplified representations, they were found to capture essential aspects affecting nanoparticle behavior. In non-functionalized models, significant distortions due to the undercoordination were observed, including Fe–O bond shortening by up to 0.46 Å and enhanced magnetic moments on oxygen atoms. These changes disrupted ferrimagnetic ordering, with spin-flipping in both tetrahedral and octahedral sublattices leading to an almost 90% reduction in total magnetization. Upon passivation, these effects were largely mitigated: Fe–O bond lengths became more uniform and ferrimagnetic alignment was stabilized as the energetically preferred state. Averaged spin–flip energies were computed to be 58 meV and 72 meV for both geometries, which is markedly lower than values for bulk γ-Fe2O3 (407–534 meV), suggesting that magnetic disorder may emerge during synthesis at typical growth temperatures. Charge transfer analysis further showed that surface coordination strongly affects electron distribution, with surface capping restoring near bulk-like charge states.
Thanks to the ability to synthesize biocompatible IONPs with low cytotoxicity,12,13 they can be utilized in different biomedical fields, such as magnetic separation,14 carriers for drug delivery,15–17 contrast agents for magnetic resonance imaging (MRI)18–21 and heating agents in local magnetic hyperthermia treatment.22–26 For many biological applications, the size of the nanoparticles plays a crucial role in determining their effectiveness. Multiple studies have shown that monodisperse nanoparticles with cores smaller than 20 nm27–31 often yield better results. These small iron oxide nanoparticles, commonly referred to as superparamagnetic iron oxide nanoparticles (SPIONs),27,28 exhibit prolonged circulation times29 and can navigate biological barriers and tissues more efficiently, making them particularly advantageous for medical use. Some researchers have further refined this classification by defining ultra-small superparamagnetic iron oxide nanoparticles (USPIONs),30 typically with core sizes below 5 nm. For instance, Wei et al. synthesized USPIONs with an estimated diameter of 1.3 nm and demonstrated their ability to enhance longitudinal relaxation (T1), thereby brightening the signal in T1-weighted MRI,32 due to a magnetically disordered surface layer. Their analysis also revealed that these nanoparticles exhibit a structure similar to maghemite and magnetite.
Recent studies have shown that IONPs with an anisotropic shape, such as cubes,33–35 polyhedrons,33,36 nanorods37 and stars,33 often outperform their spherical counterparts in various biomedical applications. For example, Zhou et al.38 successfully synthesized tetrahedral and truncated polyhedral iron oxide nanoparticles and found that their T1 and T2 relaxation times were significantly shortened compared to the spherical NPs. This enhancement was attributed to their exposed facets and anisotropic magnetic properties, making them highly effective MRI contrast agents.
While the practical applications of magnetic iron oxide nanoparticles are well studied and established, a fundamental understanding of how microscopic parameters – such as size, shape, composition, and surface effects – influence their magnetic properties remains an area of ongoing research.
Computational studies employing density functional theory (DFT) have primarily focused on the structural, geometric and magnetic properties of small neutral stoichiometric (Fe2O3)n clusters with values n ranging from 1–539–41 up to 10, see ref. 42. Magnetic states of additional (FeO)m (m ≤ 16)43,44 and FenOm (n = 1–5)45 cluster systems were analyzed in previous publications as well. Gutsev et al.46 also investigated how functionalization affects the properties of (Fe2O3)4 clusters, showing that adding up to 18 hydrogen atoms alters both their structure and spin magnetic moments. Several authors have explored the possible adsorption behavior of the D-penicillamine47 molecule, Melphalan48 and Flutamide49 drugs on the surface of Fe6(OH)18(H2O)6 ring clusters. In addition, larger clusters50 with compositions Fe25O30 and Fe33O32 were studied using spin-polarized generalized gradient approximation (SGGA) methods, both with and without the Hubbard +U correction (SGGA + U), revealing the energetic preference for ferrimagnetic ordering. López et al.51 analyzed the ferrimagnetic configuration of an even larger spherical nanometre-sized Fe45O68 cluster and investigated the possibility of spin-flipping in the tetrahedral sublattice of the Fe17O16 cluster, which they referred to as a partial ferrimagnetic state. Beyond studying the electronic structure of iron oxide clusters at the quantum mechanical level, several theoretical studies have included simulations of small nanoparticle models (up to 5 nm), employing molecular dynamics simulations52,53 to explore their geometric structure and crystallization processes.
The aim of this study is to investigate structural, magnetic and thermodynamic properties of USPIONs with atomistic precision by using density functional theory. Our analysis captures the effects of surface atoms, local coordination environments and spin configurations. We deliberately focus on an in-depth analysis of tetrahedral nanoparticle models, particularly since this morphology was successfully obtained at the size scale studied here (Fig. 1). Notably, the authors of ref. 52, using molecular dynamics simulations, also reported that the tetrahedral morphology remains stable at sizes up to 3 nm, further supporting the relevance of our chosen morphology. To further assess the influence of surface effects, additional calculations incorporating pseudo-hydrogen atoms were performed to approximate the impact of surface functionalization. To our knowledge, pseudo-hydrogen atoms have not yet been employed to model surface capping on iron oxide nanoparticles. This approach has been mainly applied to slab models to eliminate artificial electric fields54,55 or to semiconductor NPs such as CdSe, TiO2 and GaAs to improve electronic structure predictions.56,57 However, its impact on magnetic properties remains unexplored, representing a gap in the computationally accessible modelling of magnetic nanomaterials. The resulting systems were compared with the behavior of bulk magnetite (Fe3O4) and maghemite (γ-Fe2O3).
The computational cells for magnetite and maghemite were constructed based on crystallographic data from experimental publications,71,72 while the nanoparticle morphologies were inspired by our experimental research. The analysis of the product, that we synthesized, performed using a Titan Themis 60–300 cubed transmission electron microscope, confirmed the presence of ultra-small tetrahedron-shaped IONPs enclosed by low-index {111} crystal facets (Fig. 1). This is in good agreement with the results obtained by Narnaware et al.,73 who also reported similar morphology that was easily explained due to the lower surface energy of {111} facets compared to {100} and {110} facets.
Magnetite crystallizes in a cubic inverse spinel structure containing 56 atoms per unit cell, as shown in Fig. 2(a). The composition can be further expressed as (Fe3+)A(Fe3+Fe2+)BO32. The A-site iron atoms (8 FeA per unit cell) are tetrahedrally coordinated by oxygen, forming a diamond-like lattice. The remaining iron atoms, consisting of Fe3+ and Fe2+ ions in the ratio 1:
1,74 occupy the octahedral B-sites (16 FeB per unit cell). By introducing three iron vacancies75 into the octahedral (FeB) sublattice of magnetite, the unit cell of maghemite is obtained, resulting in a stoichiometric formula of Fe2O3.05. This transformation occurs in nanoparticles at temperatures as low as 50 °C76 and is accompanied by the oxidation77 of Fe2+ to Fe3+. Cubic γ-Fe2O3 structure (Fig. 2(b)) emerges specifically when the distribution of created vacancies is disordered.
Two different tetrahedral morphologies of nanometre-sized IONPs were generated: (a) a tetrahedron-shaped nanoparticle (NPtetra) with four {111} facets, consisting of 161 atoms (5 FeA, 40 FeB and 116 O), as shown in Fig. 3(a); and (b) a nanoparticle in the shape of a truncated tetrahedron (NPtrunc-tetra) with eight {111} facets, composed of 145 atoms (5 FeA, 36 FeB and 104 O), as shown in Fig. 3(b). The NPtrunc-tetra morphology was derived from NPtetra by selective removal of FeO3 segments from all four vertices. The surface layers of NPtetra and NPtrunc-tetra are made up of oxygen atoms (Ofacet and Oedge) bonded to octahedral FeB atoms in the subsurface layer. Both nanoparticles were properly isolated within periodic boundary conditions applied by the VASP code, which was achieved by encapsulating them in cubic computational cells filled with vacuum, each with an edge length of 32 Å. Other nanoparticle sizes in the ultra-small region were not analyzed, as preserving the FeA:FeB:O stoichiometry and surface termination requires scaling of all three species simultaneously, leading to larger models with ∼400 atoms that could not be fully relaxed due to the high computational cost. Further size reduction beyond the truncated morphology will yield smaller clusters with little or no tetrahedral symmetry.
Two additional models of both nanoparticles, denoted as NPHtetra (Fig. 3(c)) and NPHtrunc-tetra (Fig. 3(d)), were created by bonding surface oxygen atoms to pseudo-hydrogen atoms (psHs). Pseudo-hydrogen atoms are a computational construct that assigns effective charges to unsaturated surfaces. In nanostructures, surface atoms often possess dangling bonds (DBs) due to their reduced coordination compared to their bulk counterparts. The use of psH atoms helps to prevent artificial charge accumulation and provides a more realistic approach to the experimental conditions.56,78 In simplified terms, pseudo-hydrogen atoms effectively mimic functionalization and surface capping without the explicit inclusion of ligands. However, they are unsuitable when chemical accuracy is required, such as in studies involving steric and electrostatic effects, binding energies, solvation or adsorption mechanisms. The assignment of effective charge values on psHs follows a simple chemical consideration known as the electron counting rule (ECR).79 For instance, Suzdalev et al.80 reported that magnetic nanoparticles assume an oxidized-like state including only Fe3+ species. In NPtetra and NPtrunc-tetra structures, these Fe3+ atoms located in the subsurface layer are coordinated by six oxygen atoms, while the inner oxygen atoms are bonded to four iron atoms. Assuming that each FeB atom donates 0.5 electrons per FeB–O bond, the DBs on surface oxygen atoms (Fig. 4(b)) are passivated with psHs assigned with different valencies:
1. psH0.5 → if one DB is present (Fig. 4(c)),
2. psH1 → if two DBs are present (Fig. 4(d)),
3. psH1.5 → if three DBs are present (Fig. 4(e)).
The overall distribution of psHs with different charge states across NPHtetra and NPHtrunc-tetra is illustrated in Fig. 5. This approach ensures charge neutrality of the surface.
All systems in this study were calculated using a spin-polarized setting for collinear magnetism, with initial atomic magnetic moments set to ferrimagnetic ordering. For the γ-Fe2O3, NPHtetra and NPHtrunc-tetra structures, supplementary calculations with different magnetic configurations were also performed, followed by an energetics analysis. The heats of formation for all structures were calculated relative to the total energies of pure elements in their ground states. In the case of γ-Fe2O3, the heat of formation −1.4080 eV at−1 was determined (with an accuracy of 0.0015 eV at−1, based on basis set convergence tests performed up to a plane wave cut-off energy of 750 eV). The effect of dispersion correction on the heat of formation is discussed in Appendix A. The local charges of ions were analyzed using the method developed by R. Bader and later implemented by Henkelman et al.81 and Yu et al.82 This approach defines and separates atoms within a system using zero-flux surfaces, where the charge density reaches a local minimum perpendicular to the surface. To gain insight into the structural properties, the radial distribution function (RDF) was computed. The RDF describes how the atomic density varies as a function of distance from a reference atom. The structural visualization of computational cells was created using the VESTA package.83
Table 1 presents the equilibrium structural characteristics of Fe3O4 and γ-Fe2O3 bulk phases, including lattice parameters and interatomic distances, along with a comparative analysis against theoretical predictions and experimental results from previous studies. The computed cubic lattice parameter of Fe3O4 exhibits excellent agreement across most theoretical methods (with deviations in volume per atom from the experimental data84 remaining below 1%), except for LDA-based ones.86 For γ-Fe2O3, calculations incorporating the empirically fitted Hubbard +U correction done by other authors75,89 show a slightly larger deviation of 2%, though this remains within an acceptable accuracy range compared to lattice parameters experimentally derived from XRD patterns87 reported by Shmakov et al. In magnetite, the interatomic distances for FeA–O and FeB–O bonds obtained from our DFT calculations show complete uniformity with fixed values of 1.88 Å and 2.06 Å, respectively. This structural consistency is also reflected in the RDF spectrum of maghemite (Fig. 6), where the vacancies within the octahedral sublattice disrupt the well-defined bonding environment, leading to lattice distortions. As a result, the RDF spectrum of γ-Fe2O3 displays visibly polydisperse peaks, indicating a broader range of Fe–O bond lengths (1.83–1.92 Å for FeA–O and 1.91–2.19 Å for FeB–O). The presence of vacancies in the γ-Fe2O3 sublattice further modifies the local coordination environment of oxygen atoms. As a result, only 43.8% of oxygen atoms are coordinated by four iron atoms (coordination number CN = 4), compared to 100% in Fe3O4. The remaining oxygen atoms are bonded to only three iron atoms (CN = 3). Despite the internal lattice deformation, both bulk structures retain a cubic arrangement with all angles being precisely 90°.
Config. | Space group | n | a | Vat | Vat/Vexp | FeA–O (Å) | FeB–O (Å) | Ref. | |||||
---|---|---|---|---|---|---|---|---|---|---|---|---|---|
Fe | O | (Å) | (Å3 at−1) | (%) | dmin | dmax | dmean | dmin | dmax | dmean | |||
a GGA with Hubbard +U correction (parameter Ueff = 3.8 eV) using Perdew–Wang 1991 (PW91) exchange–correlation functional.b Localized and rotationally invariant version of the local spin density approximation with parameter Ueff = 3.94 eV.c Spin-polarized PBE with Hubbard +U correction (parameter Ueff = 4.1 eV75 and 4.3 eV89). | |||||||||||||
Fe3O4 | Fd![]() |
24 | 32 | 8.386 | 10.53 | 99.6 | 1.88 | 1.88 | 1.88 | 2.06 | 2.06 | 2.06 | This work, PBE |
24 | 32 | 8.396 | 10.57 | 100.0 | — | — | — | — | — | — | Exp84 | ||
24 | 32 | 8.390 | 10.55 | 99.8 | — | — | 1.89 | — | — | 2.05 | GGA + U85![]() |
||
24 | 32 | 8.387 | 10.53 | 99.6 | — | — | 1.88 | — | — | 2.06 | PBE74 | ||
24 | 32 | 8.271 | 10.10 | 95.5 | — | — | 1.86 | — | — | 2.03 | LSDA + U86![]() |
||
γ-Fe2O3 | P4332 | 21 | 32 | 8.332 | 10.91 | 100.0 | 1.83 | 1.92 | 1.86 | 1.91 | 2.19 | 2.03 | This work, PBE |
21.3 | 32 | 8.347 | 10.91 | 100.0 | 1.84 | 1.89 | 1.85 | 2.01 | 2.11 | 2.08 | Exp87 | ||
21 | 32 | 8.335 | 10.93 | 100.1 | — | — | — | — | — | — | PBE88 | ||
21 | 32 | 8.395 | 11.16 | 102.3 | — | — | 1.85 | — | — | 2.04 | PBE + U75![]() |
||
21 | 32 | 8.390 | 11.14 | 102.1 | — | — | — | — | — | — | PBE + U89![]() |
Table 2 summarizes the magnetic moments obtained for bulk Fe3O4 and γ-Fe2O3 phases, along with values reported in previous theoretical studies. Both structures were confirmed to have ferrimagnetic ordering, characterized by antiparallel alignment of magnetic moments between iron atoms in the tetrahedral and octahedral sublattices. By integrating the spin density within the PAW spheres surrounding each iron cation, we obtained average magnetic moments of −3.49 μB for tetrahedral FeA and 3.58 μB for octahedral FeB iron atoms in Fe3O4. For γ-Fe2O3, the respective values were −3.35 μB and 3.66 μB. However, comparison with other methods revealed notable discrepancies. In Fe3O4, the computed moment for FeA atoms is underestimated by 0.33 μB relative to neutron diffraction93 measurements. For γ-Fe2O3, neutron diffraction94 experiments report values of −4.18 μB for FeA and 4.41 μB for FeB, showing the difference of up to 0.83 μB from our results. This deviation is primarily due to the excessive delocalization of d-electrons, a well-known feature of transition metal oxides, which challenges the accuracy of standard ab initio exchange–correlation functionals. Table 2 also highlights that previous studies incorporating the +U correction generally yield more accurate results, although discrepancies with experimentally observed properties still remain. The main challenge of this approach lies in the fact that the effective U parameter (Ueff) is empirically fitted to experimental reference data, such as band gaps. Other authors conducting calculations on bulk magnetite and maghemite have employed a wide range of U values, ranging from 3.8 eV85 to 4.3 eV.89 Notably, Grau-Crespo et al.91 and Righi et al.92 used the same U parameter for their γ-Fe2O3 calculations and yet obtained significantly different results, with the average magnetic moments differing by 0.53 μB for FeA and 0.43 μB for FeB. This underscores the inherent limitations of the semi-empirical +U approach, which are even more pronounced in systems where no clear experimental reference is available. This issue is particularly relevant for our nanoparticle models, where surface states, originating from the dangling bonds on surface atoms, modify local electronic structure, introducing in-gap states and shifting band edges. Consequently, the bulk band gap cannot serve as a reliable fitting target, and no other experimental parameters are available for direct validation. Selecting a Ueff value without a suitable reference may therefore introduce significant inconsistencies and unphysical results, rather than improve accuracy. Hybrid functionals offer a more rigorous alternative by partially correcting the self-interaction error, but their computational cost is substantial. We successfully performed a single-point HSE06 calculation on bulk γ-Fe2O3 (with PBE-relaxed structure), obtaining magnetic moments of −4.03 μB (FeA) and 4.14 μB (FeB), which more closely match the experimental values. However, due to the high computational demand, we had to restrict the k-point sampling to a single Γ-point and were unable to extend this approach to larger systems, such as our nanoparticle structures. For these reasons, we chose to rely on parameter-free computationally accessible theoretical methods, using the PBE functional, as prior research has shown that the most fundamental magnetic properties – such as spin orientations and relative magnitudes of magnetic moments – are qualitatively well captured in all approaches mentioned above (PBE with and without the +U correction, as well as HSE06). This observation is supported by a comparison of the computed magnetic moment per formula unit (f.u.) for γ-Fe2O3, which in our case is 2.24 μB f.u.−1 (PBE) and 2.25 μB f.u.−1 (HSE06), which closely agrees with 2.26 μB f.u.−1 reported by Sahu et al.,89 who applied an effective parameter Ueff = 4.3 eV. Similarly, Bentarcurt et al.75 obtained 2.28 μB f.u.−1 using the Ueff = 4.1 eV correction. In contrast, the experimental value95 is 2.5 μB f.u.−1
Config. | μmean (μB at−1) | Ref. | ||
---|---|---|---|---|
FeA | FeB | O | ||
a LSDA with Vosko–Wilk–Nusair (VWN) parametrization.b Spin-polarized PBE with Hubbard +U correction (parameter Ueff = 4.0 eV).c GGA with Hubbard +U correction (parameter Ueff = 4.0 eV). | ||||
Fe3O4 | −3.49 | +3.58 | +0.08 | This work, PBE |
−3.47 | +3.58 | +0.08 | PBE74 | |
−3.41 | +3.49 | +0.11 | VWN90![]() |
|
−4.09 | +3.98 | +0.03 | PBE + U74![]() |
|
−3.68 | +3.62 | +0.06 | LSDA + U86 | |
γ-Fe2O3 | −3.35 | +3.66 | +0.09 | This work, PBE |
−4.01 | +4.13 | +0.22 | PBE + U89 | |
−4.03 | +4.16 | +0.14 | GGA + U91![]() |
|
−3.50 | +3.73 | +0.11 | PBE + U92![]() |
Additional properties of bulk γ-Fe2O3, including Bader charge analysis and formation enthalpy, were also determined and are further presented in Table 3. The results indicate that charge transfer is also influenced by the local coordination environment. Specifically, iron atoms exhibit greater electron transfer as their coordination increases from tetrahedral to octahedral, with a charge difference of 0.0956 e. A similar trend is observed for oxygen, where charge transfer varies by 0.0798 e between the threefold- and fourfold-coordination sites. γ-Fe2O3 is classified as a charge transfer insulator,91 where the electronic structure reflects the redistribution of charge, leading to the formation of Fe3+ and O2− species. The valence band is primarily composed of occupied O 2p states, while the bottom of the conduction band is dominated by unoccupied Fe 3d levels. The resulting band gap inhibits spontaneous electron migration from oxygen back to iron atoms, thereby limiting electronic conductivity under normal conditions. Our analysis supports this behavior by confirming the distinct charge transfer from Fe to O that characterizes its insulating nature.
Config. | Site | CN | q (e) | ΔfH | |
---|---|---|---|---|---|
(eV at−1) | (kJ mol−1) | ||||
γ-Fe2O3 | FeA | 4 | −1.5631 | −1.4080 | −679.26 |
FeB | 6 | −1.6587 | |||
O | 3 | +1.0297 | |||
4 | +1.1095 |
The formation enthalpy of γ-Fe2O3 was determined to be −679.26 kJ mol−1, which is underestimated compared to results obtained from earlier theoretical92 and experimental96 studies by 3.5% and 4.5%, respectively, but still remains in reasonable agreement.
Config. | nsurf | nsurf/n | FeA–Oinner (Å) | FeB–Oinner (Å) | FeB–Ofacet (Å) | FeB–Oedge (Å) | ||||||||
---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|
(%) | dmin | dmax | dmean | dmin | dmax | dmean | dmin | dmax | dmean | dmin | dmax | dmean | ||
NPtetra | 96 | 59.6 | 1.79 | 1.88 | 1.86 | 1.94 | 2.40 | 2.09 | 1.72 | 2.04 | 1.85 | 1.58 | 1.62 | 1.60 |
NPHtetra | 37.4 | 1.89 | 1.97 | 1.92 | 1.92 | 2.24 | 1.99 | 1.90 | 2.04 | 1.94 | 1.91 | 2.00 | 1.96 | |
NPtrunc-tetra | 84 | 57.9 | 1.77 | 1.88 | 1.85 | 1.94 | 2.39 | 2.08 | 1.73 | 2.02 | 1.84 | 1.58 | 1.61 | 1.60 |
NPHtrunc-tetra | 36.7 | 1.88 | 1.95 | 1.92 | 1.91 | 2.12 | 1.96 | 1.89 | 2.13 | 1.94 | 1.93 | 2.20 | 1.98 |
The NPtetra and NPtrunc-tetra models show significant structural distortions due to the exposure to vacuum conditions. Without additional external stabilization, the surface O atoms become undercoordinated with unsaturated valencies, leading to shorter and stronger bonds. This effect is particularly evident in Oedge atoms, which are bonded to only one FeB atom, resulting in a mean FeB–Oedge bond length 0.46 Å shorter than FeB–O in bulk Fe3O4. A similar, though visibly less pronounced, trend is observed for Ofacet atoms coordinated by two or three FeB atoms. The subsurface layer also adapts to the surface-induced constraints and regularly ordered core, with bond elongation serving as a compensatory mechanism. This is reflected in FeB–Oinner bonds, which extend up to 2.40 Å in NPtetra and 2.39 Å in NPtrunc-tetra. Despite the differences between the tetrahedral and truncated tetrahedral models, which belong to the same class of polyhedral structures, the overall bond lengths remain consistent.
The stabilization of dangling bonds using psH atoms compensates surface charge in a controlled manner, reducing bond length deviations and promoting a more uniform distribution of values within the 1.89–2.24 Å range for NPHtetra and the 1.88–2.20 Å range for NPHtrunc-tetra. The RDF spectra for both structures indicate that interatomic distances approaching the upper limit of the mentioned intervals (2.24 Å and 2.20 Å) are present at an insignificant frequency compared to the reference mean values for each coordination type. Additionally, compared to bulk Fe3O4, the mean FeB–Oedge/facet/inner bond lengths show a slight contraction.
Based on these observations, it is reasonable to expect a uniformity in O–psHx bond lengths across different charge states (x = 0.5, 1 or 1.5), as this would indicate that our method, based on chemical ECR principle, effectively stabilizes the surfaces and ensures a consistent structural response to the variations in charge. Indeed, our results mostly support this expectation: the O–H1.0 bond lengths range from 0.97 to 0.98 Å, with the O–H1.0 and O–H1.5 distances being nearly identical (differing by only 0.01–0.02 Å). In contrast, the O–H0.5 bonds are noticeably elongated by 0.09 Å, which may reflect a different level of local charge compensation.
Our results highlight the critical role of surface capping and full structural relaxation, as undercoordinated surface atoms undergo substantial displacements that propagate into the subsurface layer. For this reason, static calculations constrained to bulk-like interatomic distances would easily oversimplify the system and fail to capture the true impact of surface distortions on properties such as atomic-scale magnetism.
We did not manage to stabilize ferrimagnetic ordering in the NPtetra and NPtrunc-tetra. Instead, a spin-flipping phenomenon was observed in both the FeA and FeB sublattices, leading to an alternative, more complex magnetic arrangement that deviates from the expected bulk-like state. A similar effect was reported by López et al.,51 where a spin–flip occurred in a tetrahedral FeA atom positioned at the center of an isolated cage-like Fe17O16 cluster. In both our non-functionalized models, multiple spin-flips coexist simultaneously; however, one of them also involves a FeA atom located in the center of the nanoparticle, as shown in Fig. 9 for both (C)NPtetra and (C)NPtrunc-tetra. In contrast to the additional findings of López et al.,51 where the magnitudes of FeA and FeB magnetic moments in a larger (over 100-atom) spherical Fe45O68 cluster deviated by less than 0.25 μB from those found in bulk Fe3O4, we observe an anisotropic magnetic behavior, linked to the symmetry of the tetrahedral and the truncated tetrahedral structures. Fig. 10 illustrates how the absolute value of the magnetic moment of iron atoms in NPtetra decreases in discrete steps with increasing distance from the nanoparticle center, reflecting the non-equivalent positions of Fe atoms within the structure. A similar trend is also observed for NPtrunc-tetra; however, as shown in Table 5, in this geometry, there are only five structurally and magnetically non-equivalent groups of iron atoms, compared to the six present in NPtetra. The spin-flipping processes, combined with the reduced magnetic moments, result in a decrease in the total magnetization per nanoparticle (MNP) by 87% and 89% for NPtetra and NPtrunc-tetra, respectively, relative to the calculated bulk-like reference MB values (determined for the nanoparticle stoichiometries using the average magnetic moments of FeA, FeB and O atoms from our calculations of bulk γ-Fe2O3). Factors such as reduced MNP values and spin-flipping (previously used in collinear magnetism calculations employed by Bianchetti et al.98 to mimic spin-canting processes) are significant indicators of a magnetically disordered surface layer. Eddahri et al.99 reported a reduction in surface magnetization Msurf by approximately 81% compared to the value predicted for perfect ferrimagnetic order, along with a decrease in the total magnetization MNP by around 29% for a 5 nm cubic-shaped magnetite nanoparticle under vacuum conditions, simulated by Monte Carlo methods. These findings, in agreement with our results, support the observation that magnetization MNP significantly decreases with decreasing size of the nanoparticle. Given the nanometre-scale dimensions of our models, it is more appropriate to describe the entire structure as magnetically disordered, as our MNP values are comparable to the Msurf reported in ref. 99. In fact, 59.6% and 57.9% of atoms in NPtetra and NPtrunc-tetra, respectively, are located in the surface layer, thereby enhancing surface-related magnetic effects and suppressing the emergence of a distinct magnetic core. Deviations from the expected ferrimagnetic ordering can be attributed to the cleavage of surface Fe–O bonds, which disrupts the superexchange interactions between neighboring iron atoms mediated by oxygen. Additional distortions in interatomic distances (see Fig. 7(a) for NPtetra and Fig. 8(a) for NPtrunc-tetra) and angles influence these interactions by changing the overlap between Fe 3d and O 2p orbitals, thereby directly affecting the strength of exchange coupling. As a result, the system tends to minimize its local exchange energy through magnetic frustration, which manifests as the spin-flipping phenomenon, particularly under the collinearity constraints imposed during calculations. The flow of additional charge from psHs to surface oxygen atoms helps to preserve the antiparallel alignment of the FeA and FeB sublattices, resulting in the lowest energy configuration with ferrimagnetic ordering (see the relative energies in Table 6). This additional surface bonding stabilizes the Fe–O interatomic distances close to the bulk results (shown in Fig. 7(b) for NPHtetra and in Fig. 8(b) for NPHtrunc-tetra) and enhances the ferromagnetic superexchange within the FeB sublattice, compared to the non-functionalized models. Nevertheless, MNP still remains significantly reduced by 78% and 73% for NPHtetra and NPHtrunc-tetra, respectively, relative to the estimated MB values (all MNP are reported in Table 6). This ongoing decrease in MNP, despite the preservation of bulk-like spin alignment, can be linked to variations in the magnitudes of atomic magnetic moments, with some moments substantially quenched compared to bulk. This trend is consistently observed across all nanoparticle models, as exemplified by values −0.24 μB in NPtetra and 0.64 μB in NPHtetra (Fig. 9), and likely reflects altered intra-atomic exchange interactions, although a detailed understanding of this mechanism lies beyond the scope of the present study. Importantly, surface capping with psH atoms appears to restore the tetrahedral sublattice located at the center of the nanoparticle to a magnetic state more closely resembling that of the bulk (see Fig. 9 for both (C)NPHtetra and (C)NPHtrunc-tetra).
Fe site | μmean (μB at−1) | |||||
---|---|---|---|---|---|---|
NPtetra | NPHtetra | NPtrunc-tetra | NPHtrunc-tetra | |||
Core (FeA) | (C) | ↑ | 3.498 | — | 3.495 | — |
↓ | −3.565 | −3.549 | −3.512 | −3.539 | ||
Facet (FeB) | (F) | ↑ | 1.797 | 1.001 | 1.788 | 0.999 |
3.730 | 3.388 | |||||
Edge (FeB) | (E) | ↑ | 1.512 | 0.899 | 1.243 | 0.884 |
3.760 | ||||||
↓ | −1.338 | — | −1.409 | — | ||
Vertex (FeB) | (E) | ↑ | — | 0.895 | — | — |
↓ | −0.214 | — |
Config. | Ordering | nSF | MNP/cell | Erel | |
---|---|---|---|---|---|
FeA | FeB | (μB) | (eV) | ||
NPtetra | SpinF | 1 | 16 | 18.4 | 0.00 |
NPtrunc-tetra | SpinF | 1 | 12 | 13.3 | 0.00 |
NPHtetra | FerriM | 0 | 0 | 30.3 | 0.00 |
SpinF | 1 | 13 | 6.4 | 1.02 | |
NPHtrunc-tetra | FerriM | 0 | 0 | 32.7 | 0.00 |
SpinF | 1 | 11 | 3.2 | 0.70 | |
Bulk γ-Fe2O3 | FerriM | 0 | 0 | 23.7 | 0.00 |
C1 | SpinF | 2 | 4 | 4.1 | 3.11 |
C2 | SpinF | 2 | 4 | 6.7 | 2.84 |
C3 | SpinF | 1 | 2 | 13.9 | 1.60 |
C4 | SpinF | 2 | 5 | 4.3 | 2.85 |
The difference in MNP between the lowest energy states of the functionalized and non-functionalized NP models reaches 9% for the tetrahedral and 16% for the truncated tetrahedral geometry, highlighting the notable impact of surface coating on the magnetic moments. These differences may have a limited impact on experimentally observed properties of ultra-small nanoparticles, due to the presence of a magnetically disordered layer (dependent on surface morphology100), which has been reported to reach approximately 0.9 nm101 in thickness. For NPs smaller than 3 nm, this disordered layer is associated with saturation magnetization values of zero or near zero.53,102 In contrast, it may become increasingly more relevant in larger nanoparticles, where the magnetic core contributes more substantially to the overall behavior.
Even though the ferrimagnetic configuration is identified as the lowest energy structure for the functionalized models, the final magnetic ordering obtained during synthesis may be influenced by various factors – especially when possible energetically accessible local minima lie close to the global minimum. To address this, we performed a comparative analysis of spin–flip energetics and heats of formation, as shown in Fig. 11. We began by computing additional metastable NPH structures featuring spin–flip distributions similar to those observed in the non-functionalized NP models, but exhibiting significantly lower MNP values (reduced by 12.0 μB and 10.1 μB compared to NPtetra and NPtrunc-tetra, respectively). The relative energies (Erel) of these spin-flipped configurations, calculated with respect to the ferrimagnetic ground state, incorporating both the magnetic rearrangement and structural relaxation, are presented in Table 6. To further investigate the stability of spin disorder, we also extended this approach to bulk γ-Fe2O3 by generating four distinct magnetic configurations (C1–C4) with varying distributions and numbers of spin-flipped Fe atoms in the tetrahedral and octahedral sublattices. C1 and C2 contain six flipped spins, while C3 includes only three and C4 comprises seven (the respective spin arrangements are visualized in Fig. 12). For C1–C4, the energy per spin–flip falls within the range of 407–534 meV, confirming that spin-flipping is energetically not favorable in the bulk, with ferrimagnetic ordering remaining highly stable. In contrast, in the anisotropically shaped NPs, the cost of a single flipped spin is substantially lower than in the bulk, with values of 72 meV for NPHtetra and 58 meV for NPHtrunc-tetra. This suggests increased magnetic softness and a higher probability of spin disorder. It is also important to note that our results were obtained at temperatures of 0 K, so the higher energy states with flipped spins may become accessible under synthesis conditions. At 300 °C (a commonly used temperature for polyhedral IONPs preparation), the energy contribution from thermal fluctuations is approximately 50 meV, which is comparable to the spin–flip barriers computed for these nanostructures. When combined with other factors, such as internal distortions and the kinetic effect of rapid growth, this supports the plausibility of the spin-disordered layer forming under realistic conditions. The heats of formation obtained for NP models without psHs atoms at the surfaces (NPtetra and NPtrunc-tetra) are visibly less negative than for bulk γ-Fe2O3, differing by 495 and 467 meV per atom, respectively. This extensive energy difference primarily reflects the cost of surface formation. In comparison, the contribution from spin-flipping is minimal, further supporting the possibility of spin disorder forming naturally as the nanoparticle structure evolves.
It is also important to consider that different capping groups can introduce varying effective charges, thereby influencing the energy associated with spin-flipping. In this study, we employed effective charge values designed to fully compensate for the disruptions in the Fe–O bonding network at the NP surfaces. On that note, Bianchetti et al.98 carried out calculations by applying various ligands to the corner and on the facet of a cubic magnetite IONP, analyzing the spin–flip energetics for a single octahedral iron atom. Their results showed that these energies vary widely from −15 meV/114 meV (ethanol) to 141 meV/280 meV (acetic acid) for local high coverage of ligands. The spin–flip energies for our psH-capped NPs are more consistent with their results for the non-bridging ligand, which is expected, as each surface oxygen atom in our models is capped individually by a psH species.
Fig. 14(a) shows that Oinner atoms (bound to four Fe atoms) receive a similar amount of charge from the surrounding iron atoms as in bulk γ-Fe2O3, differing on average by 0.14 e. This is also reflected in their absolute magnetic moments, which, for distances smaller than 4 Å from the center of the NP, closely match the bulk values. Four additional Oinner atoms are located farther than 5 Å from the nanoparticle center, positioned closer to the vertices of the NP where surface effects become more pronounced, and both their magnetic moments and transferred charge begin to deviate accordingly. In contrast to the values obtained for Oinner, the transferred charge to Ofacet atoms ranges from 0.63 to 0.77 e, while for Oedge atoms (coordinated by only one Fe atom), the entire range shifts to lower values, spanning from 0.36 to 0.42 e. This trend highlights the decreasing charge transfer associated with decreasing coordination number – a behavior already observed in bulk γ-Fe2O3 (Table 3), though with much smaller differences. Functionalization of surface oxygen atoms with psHs shifts the charge transfer values closer to those observed in bulk, particularly for Oinner and Ofacet (as shown in Fig. 14(b)). Ofacet atoms are capped with psH atoms carrying effective charges of 0.5 or 1.0, compared to 1.5 for Oedge. Our results show that the highest assigned charge, intended to compensate for the loss of three Fe–O bonds, does not fully restore the bulk electron distribution, as the charge transferred to Oedge atoms remains within 0.55–0.68 e, still differing on average by 0.49 e from the bulk. Additional Bader analysis revealed, that the average transferred charge from psH atoms to oxygen is 0.33 e for psH1.5, 0.60 e for psH1 and 0.50 e for psH0.5, indicating that only the latter fully transfers its nominal charge. This may also explain the elongation of O–psH0.5 bonds compared to O–psH1.5/1.0, as previously discussed in the section on the structural properties of the nanoparticle models. Even with these discrepancies in charge transfer (psH1.0 gives more than psH1.5), the functionalization still appears to stabilize magnetic interactions. None of the Ofacet or Oedge atoms exhibit absolute magnetic moments exceeding 0.105 μB, and in several cases, particularly among Oedge sites, the spin polarization is entirely quenched resulting in a zero atomic magnetic moment (see Fig. 13(b)). A comparison between the NPHtetra and NPHtetra (SpinF) configurations demonstrates that the spin-flipping within the Fe sublattices has a negligible impact on the local magnetic moments of oxygen atoms. As shown in Fig. 13(b) and (c), the absolute magnetic moments remain almost unchanged across the two configurations. This suggests that the dominant factor governing oxygen magnetization is charge transfer associated with the coordination environment, rather than the spin arrangement of iron atoms.
It is also important to note that the same trends observed in the tetrahedral configurations were consistently reproduced in the nanoparticles with truncated tetrahedral geometry, making a separate discussion of these results unnecessary.
The outcomes reveal that non-functionalized NPs undergo significant surface-induced distortions. The undercoordination influences the strength of superexchange interactions, alters charge transfer processes and ultimately disrupts ferrimagnetic ordering characteristic for bulk magnetite and maghemite. For the tetrahedral morphology, Fe–O bonds at the nanoparticle edges are shortened by up to 0.46 Å, while oxygen atoms exhibit elevated magnetic moments reaching up to 0.447 μB. These changes lead to spin-flipping processes in both the tetrahedral and octahedral iron sublattices, resulting in an almost 90% reduction in total magnetization. The surface passivation with psH atoms weakens these effects by saturating the dangling bonds formed upon cleavage, compensating for the unsatisfied valencies. In both functionalized models, Fe–O bond lengths become more uniform and closer to bulk values, as confirmed by RDF analysis. Ferrimagnetic ordering emerges as the energetically preferred configuration, in contrast to the alternative containing spin flipping. The total magnetization of the capped nanoparticle MNP increases slightly, by 9% for the tetrahedral and 16% for the truncated tetrahedral geometry. However, the overall magnetization is still substantially reduced compared to the estimated bulk-like values, primarily due to residual quenching of atomic magnetic moments in subsurface iron atoms (as low as –0.24 μB in NPtetra and 0.64 μB in NPHtetra).
Spin–flip energetics supports the emergence of magnetically disordered structures under realistic conditions. The average energy cost per single spin–flip was calculated to be 72 meV for NPHtetra and 58 meV for NPHtrunc-tetra, which is substantially lower than the 407–534 meV/spin–flip observed in metastable bulk γ-Fe2O3 configurations.
Charge transfer analysis demonstrates that oxygen atoms in different coordination environments exhibit distinct behavior. In the non-functionalized NPtetra model, oxygen atoms located at the edges receive only 0.36–0.42 e from neighboring iron atoms, while those in the facets receive 0.63–0.77 e, reflecting a trend of decreasing charge transfer with lower coordination number. Upon the functionalization, charge transfer to oxygen atoms located in the facets in NPHtetra increases to 1.05–1.25 e, more resembling bulk-like values. Although the psH1.5 group gives less than half of its nominal charge – resulting in reduced charge transfer to oxygen atoms located at the edges of NPHtetra – this capping method still helps to stabilize the local electronic environment.
Beyond the immediate significance of our results, this study also demonstrates a novel use of a pseudo-hydrogen passivation scheme to model surface capping on iron oxide nanoparticles. This method offers a computationally accessible strategy that may prove broadly useful for future simulations of magnetic nanomaterials. Looking ahead, the framework established here could be extended to larger and more complex nanoparticle morphologies by integrating machine-learned potentials trained on quantum-mechanical data. However, such extensions remain beyond the scope of the present work, as reliable machine-learning models capable of capturing complex magnetic interactions with high accuracy are still under active development.
Method | ΔfHa | ΔfH/ΔfHexp | μmean (μB at−1) | a | a/aexp | |||
---|---|---|---|---|---|---|---|---|
eV at−1 | kJ mol−1 | (%) | FeA | FeB | O | (Å) | (%) | |
a Heats of formation were calculated with respect to the equilibrium elemental reference states (bcc Fe and O2 molecule) using the same computational approach. | ||||||||
Standard PBE | −1.4080 | −679.26 | 4.5 | −3.35 | +3.66 | +0.09 | 8.3317 | 0.2 |
PBE with DFT-D3 | −1.4199 | −684.99 | 3.7 | −3.33 | +3.65 | +0.09 | 8.3013 | 0.6 |
PBE with DFT-D3 (BJ) | −1.4361 | −692.84 | 2.6 | −3.31 | +3.64 | +0.09 | 8.2789 | 0.8 |
Including dispersion corrections reduces the error in the calculated heat of formation (ΔfH) of bulk γ-Fe2O3 relative to the experimental data. While standard PBE functional yields a deviation of 4.5%, applying the DFT-D3 correction lowers this deviation to 3.7%, and the Becke–Johnson (BJ) damping variant further improves the agreement to 2.6%. However, this enhanced thermochemical accuracy is accompanied by an increased discrepancy in structural parameters – the lattice constant a becomes progressively underestimated, with the BJ-corrected value deviating by 0.8% from experiment. Notably, while the DFT-D3 (BJ) correction improves ΔfH accuracy by 1.9% relative to PBE, it also induces a 1.9% contraction in the atomic volume, reflecting a denser lattice.
Given that our study focuses primarily on relative energy differences – such as spin–flip energetics or formation enthalpy variations between different magnetic configurations – the systematic component of the dispersion energy is expected to have a very small influence. This expectation is supported by additional single-point calculations performed for the PBE-relaxed ferrimagnetic bulk γ-Fe2O3 and magnetic C1–C4 structures using the DFT-D3 (BJ) correction. The differences in formation enthalpies relative to the ferrimagnetic reference state were 60 meV at−1 for C1 (compared to 59 meV at−1 with PBE), 52 meV at−1 for C2 (PBE: 54 meV at−.1), 31 meV at−1 for C3 (PBE: 30 meV at−1) and 52 meV at−1 for C4 (PBE: 54 meV per atom). Moreover, since neither the bulk nor nanoparticle models contain features typically associated with dispersion-dominated systems (e.g., hydrogen bonding, organic ligands, or van der Waals layering), the physical justification for including dispersion corrections is limited. In this context, their role can be interpreted primarily as an empirical adjustment to improve agreement with absolute formation enthalpies, rather than reflecting physically relevant interactions in the system.
This journal is © the Owner Societies 2025 |