Van-Hien
Hoang
a,
Nam-Suk
Lee
b and
Heon-Jung
Kim
*ac
aDepartment of Physics, Graduate School, Daegu University, Gyeongbuk 38453, Republic of Korea
bNational Institute for Nanomaterials Technology (NINT), Pohang University of Science and Technology (POSTECH), Pohang, 37673, Republic of Korea. E-mail: nslee@postech.ac.kr
cDepartment of Materials-Energy Science and Engineering, College of Engineering, Daegu University, Gyeongbuk 38453, Republic of Korea. E-mail: hjkim76@daegu.ac.kr
First published on 9th June 2025
Dislocations commonly occur in thin films under large misfit strain due to the accumulation of strain energy, significantly altering the films' properties. This study investigates the microstructure of Fe2O3 polymorphs in films of various thicknesses deposited on yttria-stabilized zirconia (001) substrates. The results reveal that the ε-Fe2O3 phase is formed and stabilized at thicknesses below a critical threshold of 20 nm. Beyond this threshold, the volume fraction of the ε-Fe2O3 phase decreases, and the α-Fe2O3 phase begins to emerge. With further increases in thickness, the ε-Fe2O3 phase fully transforms into the α-Fe2O3 phase. Detailed analysis suggests that this phase transformation is driven by the formation of misfit dislocations at the film/substrate interface, which compensates for the tensile strain induced by the substrate.
To date, four crystalline polymorphs of Fe2O3 (also known as iron(III) oxide or ferric oxide) have been explored, each with significantly different structural and magnetic properties: α-Fe2O3, β-Fe2O3, γ-Fe2O3, and ε-Fe2O3.17 Among these polymorphs, ε-Fe2O3 is particularly remarkable because it exhibits a giant coercive field of approximately 2 T at room temperature. Recent studies suggest that this high room-temperature coercivity arises from the disordered structure of ε-Fe2O3.18 Additionally, ε-Fe2O3 exhibits a large magnetocrystalline anisotropy, which is attributed to the establishment of a single-domain character in ε-Fe2O3 nano-objects, as well as the nonzero orbital component of the Fe3+ magnetic moment, contributing to strong spin–orbit coupling.19,20 These unique properties make ε-Fe2O3 an attractive material for high-coercivity recording media. Moreover, its millimeter-wave ferromagnetic resonance and magnetoelectric coupling21,22 make it suitable for various applications, including electric/magnetic field-tunable devices and technologies requiring effective suppression of electromagnetic interference and stabilization of electromagnetic transmission. The ε-Fe2O3 phase could become one of the most important functional magnetic materials if synthesized in pure form and with high yield. Its practical application could potentially surpass the current material limits in technologies requiring significant magnetic hardness. Moreover, ε-Fe2O3 may open up new technological areas, benefiting from its remarkable coupled magnetoelectric properties, where an applied electric field can influence its magnetic characteristics. This property is highly valuable for applications in low-power spintronic devices, such as magnetoelectric random-access memory (MeRAM) and voltage-controlled magnetic tunnel junctions, as well as for ferromagnetic resonance capabilities, which are useful for high-frequency applications such as microwave filters, isolators, and circulators.23 These properties make it an ideal candidate for high-speed spintronic devices and radar communication systems, which are uncommon in simple iron oxides. However, synthesizing pure samples of this nanomaterial without contamination from other iron oxide phases is highly challenging due to its high surface energy. The high surface energy of ε-Fe2O3 results in a higher nucleation barrier, making it difficult to initiate and sustain its growth as the dominant phase. Instead, competing phases like α-Fe2O3 and γ-Fe2O3, which have lower nucleation barriers, tend to form more readily. Additionally, ε-Fe2O3 has significant thermal instability.19,24 Up to now, ε-Fe2O3 has been synthesized nanoparticles having a spherical (sphere-like)25–28 and nanorod (nanowire)morphology.20,21,29,30 For films, ε-Fe2O3 was successfully deposited on STO (111),31 yttrium-stabilized zirconia (YSZ) (100),32,33 and muscovite (mica)34 substrates. However, the mechanism underlying the stability of the ε-Fe2O3 film on the substrate remains under debate, and its phase transformation kinetics are still not fully understood. In this study, we have investigated the effect of film thickness on the film's structure. Based on X-ray diffraction (XRD) and high-resolution transmission electron microscopy (HRTEM) measurements, we demonstrate that the Fe2O3 phase transformation is attributed to the formation of a misfit dislocation array at the film/substrate interface when the film thickness exceeds a threshold of 20 nm. Note that our study focuses on specific PLD growth conditions, annealing, and the use of a YSZ substrate. Since these factors were crucial for obtaining single-phase ε-Fe2O3, variations in temperature, pO2, substrate type, and annealing conditions could influence phase formation. These insights extend beyond our specific experimental setup. However, previous studies on ε-Fe2O3 synthesis under diverse conditions also emphasize phase changes with thickness. These findings highlight that controlling misfit dislocations and phase transformations through thickness adjustments offers a pathway to tune the material properties of Fe2O3 films finely.
The crystal structure of the films was analyzed using X-ray diffraction (XRD) with synchrotron radiation. Measurements were conducted at the 3A beamline of the Pohang Light Source (PLS) in South Korea, utilizing an 11.17 keV photon beam with a wavelength of 1.11 Å.
To investigate the thickness, microstructure, and atomic structure of the films, high-resolution transmission electron microscopy (HRTEM) was performed using a Cs-corrected scanning transmission electron microscope (STEM, JEOL JEM-2100F) at an accelerating voltage of 200 kV. The chemical composition of the films was assessed through energy-dispersive X-ray spectroscopy (EDS) elemental mapping.
YSZ is cubic in its bulk form with a lattice parameter of a = 5.12 Å, while ε-Fe2O3 has an orthorhombic structure with aε = 5.08 Å; bε = 8.78 Å; cε = 9.47 Å. In contrast, α-Fe2O3 possesses a rhombohedral structure with aα = bα = 5.05 Å; cα = 13.74 Å. Understanding how ε-Fe2O3 and α-Fe2O3 can grow on the YSZ (001) substrate is intriguing. Luca Corbellini et al. proposed two possible mechanisms for the ε-Fe2O3 growth on YSZ.32 The first mechanism involves the matching of the in-plane lattice parameter aε of ε-Fe2O3 with that of the YSZ substrate, as illustrated in the lower left panel of Fig. 1(c). Since aε is smaller than aYSZ, the film is tensile strained by the substrate to be 0.58%. In the second case, shown in the middle of the lower panel of Fig. 1(c), the diagonal of the orthorhombic unit cell, formed by the short and long sides (aε and bε), matches twice the lattice constant of the substrate, 2aYSZ ≈ (aε2 + bε2)0.5 resulting in a tensile strain of approximately 0.97%. When the film thickness exceeds 20 nm, the α-Fe2O3 phase begins to appear alongside ε-Fe2O3. There is an edge point where two ε-Fe2O3 planes form an angle of ∼120°. There is a key point where two ε-Fe2O3 planes form an angle of ∼120°, which corresponds to the angle formed in the α-Fe2O3 structure.32 The lower right panel of Fig. 1(c) illustrates the mechanism responsible for the formation of the α-Fe2O3 phase.
Fast Fourier transform (FFT) analysis of the HRTEM images was performed to examine the crystal structure and lattice orientation of the films. Fig. 2(a) and (b) show cross-sectional HRTEM images near the interface for samples with thicknesses of 20 nm and 50 nm. The corresponding FFT images of the regions highlighted by white solid boxes are displayed at the bottom of Fig. 2. The FFT analysis reveals that the crystal structure of the thickness 20 nm corresponds to an orthorhombic phase at both [010] and [−100] zone axes, consistent with ε-Fe2O3, while the film with 50 nm exhibits a rhombohedral structure at the [110] and [010] zone axes, characterized by α-Fe2O3 phase. These results are consistent with the X-ray diffraction patterns shown in Fig. 1(a). Additionally, we have determined the lattice spacing of several planes of ε-Fe2O3 near the film/substrate interface as follows: (01−3): 0.265 nm (ref. ∼0.296 nm); (−21−3): 0.177 nm (ref. ∼0.193 nm); (−200): 0.251 nm (ref. ∼0.254 nm).
Next, we will discuss why ε-Fe2O3 can be stabilized on YSZ (001) substrate up to 20 nm in thickness. For nanoparticles, in general, two factors have been found to play an essential role in determining which nanosized Fe2O3 polymorph will be formed from a given precursor and how it can subsequently be transformed into various ferric oxide phases.35 These parameters include chemical potential (η) and surface energy (σ). In this case, ε-Fe2O3 can exist when the size of the Fe2O3 particle falls within an interval defined by −6ν(σε − σγ)/(ηε − ηγ) < d < −6ν((σε − σα)/(ηε − ηα)). Where ν is molar volume; d is the size of a nanoparticle; ηα, ηγ, ηε and σα, σγ, σε are chemical potential and surface energy of α-Fe2O3, γ-Fe2O3, and ε-Fe2O3 phases, respectively. This implies that if nanoparticles of Fe2O3 grow large enough, the existence of ε-Fe2O3 is no longer favored. In other words, reducing the sizes of the Fe2O3 particle increases the contribution of the surface (or interface) energy to the system, which stabilizes ε-Fe2O3 in the nanoscaled size. However, for thin films, in addition to the two factors mentioned, strain energy contribution can play an important role in the stability of ε-Fe2O3 film. Since the YSZ substrate induces tensile strain in the films, as shown in Fig. 1(c), the relaxation of strain in the film occurs through the formation of dislocation, leading to misfit dislocation at the film/substrate interface, and the dislocation density of partially relaxed films may depend on the layer thickness.
To confirm the existence of dislocations, the microstructure of the as-prepared thin films with thicknesses of 20 nm and 50 nm on the YSZ (001) substrate was investigated using a high-angle annular dark-field scanning transmission electron microscopy (HAADF-STEM) projected along the [100] zone axis, as shown in Fig. 3. Fig. 3(a) reveals periodic black dots at the ε-Fe2O3/YSZ interface (marked by red circles) and straight lines (indicated by red arrows) corresponding to misfit dislocations and threading dislocations, respectively.5,31,36 Since there is a large lattice mismatch between the ε-Fe2O3 thin films and the substrate, initially, the film remains coherently strained, meaning its lattice conforms to the substrate. However, as the film grows thicker, the accumulated strain energy increases. Once the critical thickness is exceeded, the film can no longer maintain coherent strain, and misfit dislocations form at the interface to relieve strain. This transition leads to a strain-relaxed state, where the film begins to adopt its bulk crystal structure rather than being constrained by the substrate lattice. As a result, the film loses its coherency with the substrate.11 In our study, the observed structural transition at increased film thickness is a consequence of this strain relaxation mechanism. As the film thickness increases to 40–45 nm (with dislocations still observed in the HAADF-STEM images, though not shown here), the density of misfit dislocations rises, generating stress fields that effectively compensate for the misfit stresses. If it is energetically favorable for misfit dislocations to form at the film/substrate interface, these defects on the ε-Fe2O3 surface may act as catalysts for the nucleation of a new phase, α-Fe2O3, within the film.37 This is consistent with the observed data in Fig. 1(a), where ε-Fe2O3 and α-Fe2O3 phases coexist when the thickness is within the range of 40–45 nm. The role of misfit dislocations as catalytic nucleation sites for the formation of a new phase has also been observed in other materials.4,14,38 The relative potency of these catalytic sites determines the spatial distribution of nucleation sites. Once the misfit dislocation network is fully formed as dislocation arrays at the film/substrate interface, the process of climb occurs, characterized by mass transportation around the dislocation cores. This relaxes the misfit strain in the film layer near the free surface, bringing the crystal lattice parameter of this layer closer to that of its ideal, unstrained state. Consequently, when the thickness reaches 50 nm, the strain impact from the substrate on the ε-Fe2O3 film becomes negligible, resulting in the instability of ε-Fe2O3 and its complete transformation into the α-Fe2O3 phase. Interestingly, in this case, dislocations completely disappear in the α-Fe2O3 film, as shown in Fig. 3(b). The behavior of the misfit dislocation becomes even clearer when viewed on a larger scale in the HAADF-STEM images shown in Fig. 3(c) and (d), where periodic black dots are observed at the ε-Fe2O3/YSZ interface [Fig. 3(c)], but are absent at the α-Fe2O3/YSZ interface [Fig. 3(d)].36 The absence of misfit dislocations in the 50 nm-thick film (α-Fe2O3) can be explained by calculating the misfit strain energy density (wf) accumulated in the film, which is induced by the substrate. According to M. Y. Gutkin et al., wf can be described as follows:4
w f = [G/(1 − v)](fa2 + fb2 + 2vfafb), which can be simply rewritten as
wf/G = (fa2 + fb2 + 2vfafb)/(1 − v), | (1) |
The contrast variation observed in the BF-TEM image at the interface provides an important indication of the presence of misfit dislocations; however, this alone is insufficient to confirm their existence. To directly visualize misfit dislocations, STEM-HAADF imaging combined with Bragg filtering of selected atomic planes was performed on the ε-Fe2O3 film near a region exhibiting dark contrast (black dot), as shown in Fig. 4. The core of a misfit dislocation was clearly revealed within the yellow box in Fig. 4(b), extracted from the red-marked region in Fig. 4(a). Based on the analysis of a Burgers circuit around the dislocation core, the dislocation was identified as a [001]-type edge dislocation. The in-plane dislocation density δ* in the ε-Fe2O3 film (with a thickness of 20 nm) was estimated based on the number of such dislocation cores (black dots) along the interface, yielding a value of approximately 5 (dots)/22 (nm) = 22.7 × 105 cm−1.
To investigate the spatial distribution of elements in the film, energy-dispersive X-ray spectroscopy (EDS) measurements were performed on samples with thicknesses of 20 nm and 50 nm as shown in Fig. 5(a) and (b), respectively. Two distinct layers with a clear interface were observed, and the elements were represented by different colors. Notably, in the 20 nm film, the oxygen signal at the interface appeared darker than in the outer layer and substrate, suggesting the presence of oxygen vacancies. These results align with the dislocations observed in Fig. 3(a), (c) and 4, as misfit dislocations can lead to a higher concentration of oxygen defects at the interface compared to regions further away from it.5
Notably, high-resolution STEM-HAADF imaging and Bragg filtering were used to investigate the interface structure of the ε-Fe2O3/YSZ system. Although the interfacial region is only a few nanometers thick, the observed atomic stacking and lattice fringes appear consistent with the ε-Fe2O3 structure. Importantly, no lattice spacings corresponding to metallic Fe (e.g., 0.203 nm for (110)) or FeO (e.g., 0.301 nm for (200)) were observed. While EDS mapping suggests a degree of oxygen deficiency near the interface, this does not correspond to a distinct Fe or FeO phase but rather points to non-stoichiometry within the ε-phase. Due to the small number of atomic layers in this region, FFT analysis is inherently limited and should be interpreted with caution. However, the continuity of the ε-Fe2O3 lattice across the interface and the absence of secondary phases support the conclusion that oxygen loss leads to a partially reduced, non-stoichiometric ε-Fe2O3 phase. This behavior may be stabilized by misfit dislocations, which are known to accommodate strain and promote the formation of oxygen vacancies. Although our case differs from that reported by Matsuzaki et al.,39 where interfacial redox phenomena and oxygen diffusion effects in Fe3O4/YSZ significantly altered the oxide phase composition and stoichiometry, the comparison highlights the broader relevance of interfacial chemistry in oxide systems.
This journal is © The Royal Society of Chemistry 2025 |