Next Article in Journal
Graphene Growth Directly on SiO2/Si by Hot Filament Chemical Vapor Deposition
Next Article in Special Issue
Synergistic Piezo-Photocatalysis of BiOCl/NaNbO3 Heterojunction Piezoelectric Composite for High-Efficient Organic Pollutant Degradation
Previous Article in Journal
Characterization and Applications of Metal Ferrite Nanocomposites
Previous Article in Special Issue
Interface Optimization and Transport Modulation of Sm2O3/InP Metal Oxide Semiconductor Capacitors with Atomic Layer Deposition-Derived Laminated Interlayer
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Phase Structure and Electrical Properties of Sm-Doped BiFe0.98Mn0.02O3 Thin Films

1
School of Materials Science and Engineering, Shandong Jianzhu University, Jinan 250101, China
2
State Key Laboratory of Crystal Materials, Institute of Crystal Materials, Shandong University, Jinan 250100, China
3
School of Data and Computer Science, Shandong Women’s University, Jinan 250300, China
*
Authors to whom correspondence should be addressed.
Nanomaterials 2022, 12(1), 108; https://doi.org/10.3390/nano12010108
Submission received: 7 December 2021 / Revised: 22 December 2021 / Accepted: 24 December 2021 / Published: 30 December 2021

Abstract

:
Bi1−xSmxFe0.98Mn0.02O3 (x = 0, 0.02, 0.04, 0.06; named BSFMx) (BSFM) films were prepared by the sol-gel method on indium tin oxide (ITO)/glass substrate. The effects of different Sm content on the crystal structure, phase composition, oxygen vacancy content, ferroelectric property, dielectric property, leakage property, leakage mechanism, and aging property of the BSFM films were systematically analyzed. X-ray diffraction (XRD) and Raman spectral analyses revealed that the sample had both R3c and Pnma phases. Through additional XRD fitting of the films, the content of the two phases of the sample was analyzed in detail, and it was found that the Pnma phase in the BSFMx = 0 film had the lowest abundance. X-ray photoelectron spectroscopy (XPS) analysis showed that the BSFMx = 0.04 film had the lowest oxygen vacancy content, which was conducive to a decrease in leakage current density and an improvement in dielectric properties. The diffraction peak of (110) exhibited the maximum intensity when the doping amount was 4 mol%, and the minimum leakage current density and a large remanent polarization intensity were also observed at room temperature (2Pr = 91.859 μC/cm2). By doping Sm at an appropriate amount, the leakage property of the BSFM films was reduced, the dielectric property was improved, and the aging process was delayed. The performance changes in the BSFM films were further explained from different perspectives, such as phase composition and oxygen vacancy content.

1. Introduction

Only a small fraction of all magnetically polarized and electrically polarized materials are ferromagnetic or ferroelectric, and even fewer, namely multiferroic materials, have both properties [1,2]. In addition, the coupling between different properties of multiferroic materials will produce new properties, such as magnetoelectric effects. These materials have great development potential in the miniaturization and multi-functionalization of devices, as well as in a wide range of applications in the fields of magnetoelectric memory [3,4], sensors [5], and drivers [6]. Multiferroic materials are some of the most valuable multifunctional materials, and they have good application prospects in the field of multiferroic devices.
Multiferroic materials include single-phase materials and composite materials. However, few single-phase multiferroic materials have been discovered at present, and their Curie temperatures are usually low. Owing to its high Curie temperature (Tc = 1103 K) and Neal temperature (TN = 647 K), single-phase BiFeO3 (BFO) exhibits ferroelectric and G-type antiferromagnetism at room temperature [7,8]. Thus, it has attracted extensive attention from materials scholars and has become a hot topic for in-depth exploration of multiferroic materials [9,10,11,12].
In BFO, Bi ions are volatile at high temperature. To balance the charge, the valence of Fe ions may change from +3 to +2 [13]:
Bi Bi x + 3 2 O 2 Bi 2 O 3 + V Bi + 3 h ,
2 Fe 3 + + O O x 2 ( Fe Fe 3 + 2 + ) + V O + 1 2 O 2 .  
As a result, a large number of oxygen vacancies or other defects often exist in the prepared BiFeO3 samples, which increases the leakage current density of BiFeO3 materials and adversely affects its performance [14]. There are many ways to improve the properties of BiFeO3 materials, including element doping, solid solution, formation of a heterostructure, and control of film orientation [15,16,17,18]. Among them, many researchers adopt the element doping method to improve the performance of BiFeO3 materials [19,20,21,22]. Yun et al. prepared single-phase multiferroic BiFeO3 and Ho-doped BiFeO3 films [23]. The ferroelectric property was enhanced, and the leakage current decreased significantly. The ferroelectric property reached 20.69 μC/cm2 and the leakage current density was 2.89 × 10−9 A/cm2, and these effects were attributed to the transformation from a rhombohedral structure to a coexisting cubic and orthosymmetric structure after Ho doping. Moreover, the fatigue properties of the films doped with Ho also improved, as evidenced by a 0.4% reduction in the value of the switchable polarization. Liu et al. grew a Bi1−xEuxFeO3 (BEFOx, x = 0, 0.03, 0.05, 0.07, 0.1) thin film on LaNiO3-coated Si substrate by the pulse laser deposition method. As the doping amount increased, the position of the A1-1 mode of the films shifted to a higher wave number in the Raman spectrum [24]. With the increase in Eu, the refractive index of the film increased, and the extinction coefficient and band gap width decreased. Yang et al. prepared a BiFe1−xZnxO3 (BFZO) film (x = 0%, 1%, 2%, 3%) and found that when x = 2%, the film reached the maximum remanent polarization intensity and the minimum correction field [25]. At the same time, under a low electric field, Zn doping can significantly reduce the leakage current of BFO films. In addition, the leakage mechanism changes from Ohmic conduction under a low electric field to F-N tunneling under a high electric field. Zhang et al. prepared high-quality BiFe1−2xZnxTixO3 (BFZTO, x = 0, 0.01, 0.02, 0.03, 0.04, and 0.05) films [26]. The authors found that the BFZTO film with x = 0.02 had uniform fine grains and high density, which can inhibit the transformation of Fe3+ to Fe2+ and, thus, greatly reduce the oxygen vacancy concentration. This film had the lowest leakage current density and the highest remanent polarization intensity. By comparing P–E hysteresis loops in different areas of BiFe0.96Zn0.02Ti0.02O3 thin films, the films have high uniformity and stable properties. Concurrently, Zn and Ti co-doping also increased the dielectric permittivity from 24.9 to 35.3 and remnant magnetization from 0.05 to 0.80 emu/cm3 of BFZTO films. Liu et al. prepared Bi0.9Er0.1Fe1−xMnxO3 (BEFMxO, x = 0.00–0.03) thin films by the sol-gel method [27]. By co-doping Er and Mn, the coexistence of two phases (space groups are R3c:H and R3m:R) and the reduction of oxygen vacancy and Fe2+ concentration in BEFMxO were realized. Among all the samples, the BEFM0.02O film had the lowest oxygen vacancy concentration, the maximum remanent polarization value, and the maximum switching current. It also exhibited excellent ferroelectric stability, which means its low concentration of oxygen vacancies had less influence on the ferroelectric domains.
Kan et al. found that doping with Sm affected the phase structure of BFO samples [28,29]. Xue et al. prepared BFO films with different Sm content by the sol-gel method and found that the rhombohedral phase to pseudo-tetragonal phase transition occurs gradually with the increase in Sm [30]. Although there are many studies on the influence of element doping on BiFeO3 properties, there are few on the influence of Sm doping on the content change in the BiFeO3 thin film phase structure and thus on ferroelectric properties. In addition, the literature review revealed that for Sm doping, when the doping content is less than 10 mol%, BFO has better properties than heavily doped [31,32]. It is necessary to further adjust the doping content. In this experiment, the doping amounts of Sm were 2 mol%, 4 mol%, and 6 mol%, in order to understand the influence of Sm doping on BSFM films. The performance changes were analyzed in detail from the aspects of oxygen vacancy content, grain size, relative content of the R3c phase and the Pnma phase. Additionally, the effects of different Sm content on the ferroelectric, dielectric, leakage, and aging properties of the thin film samples were systematically studied.

2. Materials and Methods

Bi1-xSmxFe0.98Mn0.02O3 thin films (x = 0, 0.02, 0.04, 0.06) were prepared by the sol-gel method on ITO/glass substrate. Fe(NO3)9H2O (purity of 98.5%), Bi(NO3)3·5H2O (purity of 98.5%), Sm(NO3)3·6H2O (purity of 98.5%), and MnC4H6O4·4H2O (purity of 98.5%) were taken as solutes, according to stoichiometric ratio. Bi excess of 5% compensated for bismuth volatilization during high-temperature annealing. The solutes were successively added to a solvent mixture of CH3COOH and C2H6O, with a volume ratio of 3:1, and stirred at room temperature at a uniform speed until completely dissolved. Then, C5H8O2 was added to the solution as a chelating agent and stirred at room temperature for 12 h at a constant speed to obtain a red-brown and transparent precursor solution. Finally, the stable precursor solution of 0.3 mol/L was obtained by allowing the precursor solution to rest for 24 h. Then, the BSFM precursor solution was rotated onto ITO/glass substrate, and the film was coated at 3500 r/min. The wet film was dried on an electric heating plate at 250 °C to remove excess organic solvents and water. It was then placed in an annealing furnace and annealed at 550 °C. The coating was repeated, and the film was dried and annealed 10 times to obtain the desired samples.
Before testing the electrical properties of the sample, Au was sputtered on the surface of the sample to achieve the effect of conduction. We used a small-ion sputtering instrument (JS-1600, Beijing Hetong Venture Technology Co., Ltd., Beijing, China) to complete this process. The samples were characterized by an X-ray diffractometer (D8-Advance, Drudy, Germany) recorded in the 2-theta range of 20–60° with a step of 0.02°, and by a microconfocal Raman spectrometer (HR800, LabRAM, Horiba Co., Palaiseau, France) to measure in the shift range of 50–650 cm−1. The Fe and O elements in the samples were analyzed by a Wscabb X-ray photoelectron spectrometer. A dielectric tester (TH2828, Xintonghui Electronics Co., Ltd., Suzhou, China) was used to test the dielectric properties of the samples in the range of 1 kHz–1 MHz with an oscillation voltage of 1 V. The ferroelectric properties at 1 kHz and leakage properties of the samples were measured via a multiferroic tester (Radiant Co., Albuquerque, NM, USA).

3. Results and Discussion

Figure 1a shows the XRD patterns of Bi1-xSmxFe0.98Mn0.02O3 (x = 0, 0.02, 0.04, 0.06) films deposited on ITO/glass substrate. Figure 1b,c show the local magnified diffraction peaks of the (110) and (202) crystal planes, respectively. From the XRD pattern, the generated sample had a polycrystalline perovskite structure, and the films had good crystallinity, which matches well with the JCPDS (No. 86-1518) PDF standard card. However, the peak-splitting phenomenon of the rhombohedral structure was not observed in the BFSM thin films. At 2θ = 32°, the BSFM thin film samples do not show (110)/(104) peak splitting, but preferentially grow along (110). This may be due to the structural phase transition in the BSFM films. Figure 1b,c show that with the increase in Sm doping amount, the diffraction peak of (110) and the diffraction peak of (202) gradually shift to a large angle, which may be because the radii of doped Sm3+ (0.96 Å) and Mn2+ (0.67Å) are smaller than that of Bi3+ (1.03 Å) and Fe3+ (0.64 Å) [33,34]. Lattice distortion occurred during doping, and the crystal plane spacing decreased. In addition, with the increase in Sm content, the relative strengths I = I(110)/I(012) of BSFMx = 0.02, BSFMx = 0.04, BSFMx = 0.06 were 2.03, 3.70, 3.02, respectively. When the content was 4 mol%, the relative intensity reached the maximum value, and the intensity of diffraction peak (202) did not change significantly with doping amount.
The above data indicate that the crystal had structural distortion. To further study the changes in crystal structure, Rietveld refinement was conducted on all samples, and the results are shown in Figure 2a–d. All samples were used for two-phase refinement using R3c and Pnma cards from the International Crystallography database. The phase content and structural parameters of the samples are shown in Table 1. According to the experimental data, all samples have R3c and Pnma space groups. With the increase in Sm content, the content of R3c phase in the samples was 70.98%, 68.39%, 69.06%, and 67.06%. The addition of Sm reduced the content of R3c phase in the BSFM films. The lone electron pair 6S2 of Bi3+ in BiFeO3 is chemically active, and it is conducive to ferroelectric distortion [20]. The substitution of rare earth element Sm for Bi may reduce the chemical activity of the lone electron pair and reduce the rhombohedral distortion of the crystal. In all Sm-doped samples, the content of the R3c phase in BSFMx = 0.04 was the highest, and the content of the non-polar orthorhombic phase Pnma was the lowest [35,36].
The surface differentiation features of the BSFMx (x = 0–0.06) films are shown in Figure 3a–d. The crystal grain size of the BSFMx (x = 0–0.06) films are shown in the inset. The figure shows that the grain distribution on the surface of the film without Sm doping is not uniform, and that there are many voids, which may be the reason for the volatilization of the organic solution. Furthermore, the surfaces of the Sm-doped films were uniform, compact, and well combined with the substrate, indicating that the annealing mechanism was very suitable for the growth of the BSFM films on the ITO substrate. The average grain sizes of the BSFMx (x = 0–0.06) films were 62.29, 59.53, 48.93, and 60.98 nm, indicating that Sm doping can reduce grain size. Among them, the BSFMx = 0.04 film had the smallest grain size, suggesting that appropriate Sm doping accelerated the nucleation rate and decreased the grain size [37]. The cross-sectional image of the BSFMx = 0.06 film is shown in Figure 3e. As can be seen from the figure, the film has a clear interface with the substrate, and the cross-section thickness of the film is 585 nm.
To further analyze the structure of Sm-doped BSFM thin films, Raman spectroscopy was used. Figure 4a shows the Raman spectra of BSFMx (x = 0−0.06) films with different Sm content in the wave number range of 50 cm−1–650 cm−1. The data for the crystal structure of the BSFM films (Pnma+R3c) according to XRD-refined parameters and developed by FullProf software are shown in Figure 4b,c. Figure 4d–g are the Raman spectrum fitting diagrams of each sample. Group theory analysis shows that the vibration modes of the BFO film with the R3c space group with a rhombohedral perovskite structure are ΓRaman,R3c = 4A1 + 9E [38]. Four A1 and nine E vibration modes analyzed by group theory were observed in the Raman spectra of the BSFM films. The Raman vibration modes extracted from the Raman fitting are listed in Table 2. From the Raman fitting diagram, the strength of mode A in the BSFM films significantly increased, which may be related to the change in Bi-O bonds caused by Sm3+ replacing Bi3+. Owing to the Jahn–Teller distortion effect [39], the strength of modes E-8 and E-9 improved. This is because changes in Bi-O bonds lead to the distortion of the ferrite octahedron, which further changes the Fe-O bond. In addition, when the Sm element was added, the A1-1 vibration mode shifted to a higher wave number (from 141.76 cm−1 to 143.31 cm−1), because the frequency of the Raman vibration mode is related to the functions of ion mass and force [22]. Sm3+ (150.4 g) replaces Bi3+ (209.0 g), which leads to a blue shift in the A1-1 mode. These results show that Sm3+ doping causes lattice distortion, which is consistent with XRD analysis. In addition, all samples had vibration patterns near 200, 300, 400, 490, and 620 cm1, which were consistent with the Raman frequencies of the Pnma structure [40]. In Figure 4c–f, the vibration patterns are indicated by #. The results show that in addition to the R3c phase, the Pnma phase also existed in all samples, which was consistent with the XRD refinement results.
According to the defect equation (Formula (2)) [13], the Fe2+ content can affect the oxygen vacancy content. Thus, the Fe2p2/3 orbit of the BSFM films with different Sm doping amount was fitted, as shown in Figure 5a–d. The Fe2p2/3 peaks were fitted into two peaks corresponding to Fe2+ and Fe3+. From further calculations, the Fe3+:Fe2+ ratios of the BSFMx (x = 0–0.06) films were 2.33, 2.64, 2.94, and 2.77. The content of Fe2+ in the BSFMx = 0.04 film was the lowest, indicating that an appropriate amount of Sm doping can effectively inhibit the variation of Fe element. Generally, the content of Fe2+ will affect the generation of oxygen vacancies. The lower the content of Fe2+, the fewer oxygen vacancies will be generated, and the fewer defects will exist in the samples.
To further study the oxygen vacancy content in the BSFM films, the O1s orbital of the films was fitted, as shown in Figure 6a–d. In the figure, O1s is fitted into two peaks, which are lattice oxygen with binding energy of 529 eV (from metal) and oxygen vacancy of 531 eV (from defect) [41,42]. To calculate the oxygen vacancy ratio of the thin films, the two peaks were integrated to calculate the area, and the oxygen vacancy ratios of the four groups of samples were 0.24, 0.16, 0.14, and 0.15. According to the calculation results, the oxygen vacancy content and the Fe2p3/2 orbital Fe2+ content of the BSFMx = 0.04 film were the lowest. This further confirms the view that inhibiting the transformation of Fe3+ to Fe2+ reduces the oxygen vacancy content of the sample.
Figure 7 shows the leakage current density curve (J–E) of the BSFM film with different Sm content. The test electric field was 350 kV/cm. The asymmetry of leakage current of positive and negative electric fields can be attributed to the different work functions of the upper and lower electrodes. With an increasing doping amount, the leakage current density decreased first and then increased. Under the same electric field, when the doping amount was 4 mol%, the leakage current density reached the minimum value. Under a 350 kV/cm electric field, the leakage current density of the BSFMx = 0.04 film reached 8.84 × 10−5 A/cm2. The leakage current density was about 0.5 orders of magnitude lower than that of the pure BiFe0.98Mn0.02O3 film. Under normal circumstances, the leakage current of the sample was affected by oxygen vacancy content, which is mainly related to the high-temperature volatilization of Bi3+ and the variation in Fe3+. It can be seen from XPS that the Fe2+ content and the oxygen vacancy content in the BSFMx = 0.04 film was the lowest, resulting in the minimum leakage current density. In addition, grain boundaries hinder electron migration and reduce leakage current density. According to the SEM results, the BSFMx = 0.04 film has the smallest grain size, indicating that it has more grain boundaries and greater resistance to electron migration [26,42].
Figure 8a shows the hysteresis loops of the BSFM films with different Sm content with an electric field of 1410 kV/cm and frequency of 1 kHz. The remanent polarization intensities of BSFMx (x = 0−0.06) under the test electric field were 111.23, 83.30, 91.86, and 73.59 μC/cm2. The coercive field Ec was 835.22, 819.28, 830.95, and 987.55 kV/cm, respectively. When the content was less than or equal to 4 mol%, the change in the coercive field was small. The defects inhibit ferroelectric domain flipping. In all the Sm-doped samples, the remanent polarization value increased first and then decreased with increasing doping content. When the doping content was 4 mol%, the remanent polarization value was the largest, and the ferroelectric property was the best, which was consistent with the XRD analysis results. The possible reasons are as follows: (1) The leakage current density of BSFMx = 0.04 has a minimum value compared with other samples. A smaller leakage current density can improve the voltage resistance of the sample and facilitate electric domain inversion [28]. (2) According to XPS experimental data, the BSFMx = 0.04 sample had the lowest oxygen vacancy content, and the reduction of oxygen vacancy will improve its ferroelectric performance. (3) Owing to its centrosymmetric characteristics, the Pnma phase shows paramagnetic properties without ferroelectric properties, while the R3c phase mainly affects the ferroelectric property of samples [43,44,45].
Figure 8b,c show that the dielectric permittivity (ε) and dielectric loss (tanδ) of the BSFM films with different Sm contents vary with test frequency at room temperature in the range of 1 kHz–1 MHz. From the figure, the dielectric permittivity of the sample has little dependence on frequency, indicating that the sample has high intermediate frequency stability. The results show that the dielectric permittivity values of BSFMx (x = 0−0.06) at 10 kHz were 62, 89, 91, and 65. At the same frequency, the dielectric losses of BSFMx (x = 0−0.06) were 0.034, 0.025, 0.026, and 0.027. In the low frequency range (<40 kHz), Sm doping reduced the dielectric loss of the BSFM films. At the same time, Sm doping increased the dielectric permittivity of the BSFM films, and the maximum dielectric permittivity was obtained when the doping amount was 4 mol%. The existence of oxygen vacancies will distort the free volume used to replace Fe3+ in the Fe-O octahedron, which reduces the dielectric polarization and leads to a smaller dielectric permittivity [46]. As a result, the BSFMx = 0.04 film with the lowest oxygen vacancy content has the maximum dielectric permittivity.
BSFMx = 0.04 and BSFMx = 0.06 films were selected as representatives to explain the aging behavior of samples. Figure 9 shows the electrical hysteresis loop diagram of the BSFM films aged at room temperature for 110 d. The remanent polarization strength (2Pr) of the BSFMx = 0.04 and BSFMx = 0.06 films after aging treatment decreased by 24.9% and 41.3%, respectively, while the intensity of the coercive electric field (2Ec) decreased by 0.6% and 12.2%, respectively. This indicates that the two samples have different degrees of aging. Among them, the aging degree of the BSFMx = 0.04 film was smaller. Ren et al. proposed the symmetric short-range order principle of point defects and inferred that the aging effect could be triggered by ion doping [47,48,49,50]. Moreover, the migration of oxygen vacancies led to the formation of complex defect dipoles in the sample. The symmetrical short-range order of oxygen vacancies created conditions suitable for reversible domain switching. In addition, the internal electric field formed during the orderly arrangement of the defective dipoles will increase the offset of the coercive field, which directly explains the asymmetry of the coercive field strength of the sample shown in Figure 9. The domains within ferroelectrics can be switched under an applied electric field, exhibiting macroscopic polarization. However, when the applied electric field is removed, the domain structure gradually shifts to a random state, resulting in the degradation of properties over time, namely, sample aging. The movement of oxygen vacancies in the sample to the domain walls creates pin centers that provide resistance to the movement of the domain walls and reduce the mobility of the domain walls. This, in turn, affects the ferroelectric properties associated with the movement of the domain walls. According to the XPS data, compared with the BSFMx = 0.06 film, the oxygen vacancy content in the BSFMx = 0.04 film was less, which means that the defect dipole content of the sample was smaller. Thus, the resistance of the domain wall to movement was smaller, leading to weak aging effects and better retention of film performance. In addition, it can be seen from SEM images that the BSFMx = 0.04 film had a smaller grain size than the BSFMx = 0.06 film. For small-grain crystals, due to the small difference between lattice symmetry and defect symmetry, the thermodynamic force driving the symmetry matching between them is weak, and this hinders the kinetic migration of oxygen vacancies and affects reversible domain switching [47,51]. Therefore, the BSFMx = 0.04 film with small grains aged slowly.

4. Conclusions

In conclusion, Sm-doped BSFM films were prepared on ITO/glass substrates by the sol-gel method. The effects of different Sm content on the leakage current density, dielectric properties, and aging properties of the BSFM films were systematically studied. Detailed explanations were made in terms of phase transition and oxygen vacancy. XRD and Raman analyses show that the samples all contained R3c and Pnma phases, and the samples with different Sm content had different phase composition. Sm doping led to lattice structure distortion and decreases in the crystal plane spacing. XPS analysis showed that the BSFMx = 0.04 thin film sample had the lowest oxygen vacancy content, indicating that an appropriate amount of Sm doping can effectively inhibit the valence of the Fe element. The decrease in oxygen vacancy increased the dielectric permittivity and the leakage current density. The minimum leakage current density of the BSFMx = 0.04 film sample was 8.84 × 10−5 A/cm2 in a 350 kV/cm electric field. The remanent polarization intensity of the BSFMx = 0.04 film was 91.86 μC/cm2, owing to the formation of avnonpolar orthorhombic Pnma phase during the doping process. At the same time, the aging process of BSFMx = 0.04 sample was slow, and the performance of the sample was better preserved.

Author Contributions

Conceptualization, Y.W.; Data curation, Y.W.; Formal Analysis, Y.W.; Funding acquisition, F.Z.; Investigation, Y.W.; Methodology, Y.W., Z.M. and L.W.; Project administration, F.Z. and L.Z.; Resources, X.G. and Z.L.; Software, Y.W.; Supervision, L.Z.; Validation, Y.L., B.Y.; Visualization, Y.W.; Writing-original draft, Y.W.; Writing-review & editing, Y.W., L.W., Z.M. and F.Z. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by funding from the State Key Laboratory of Silicate Materials for Architectures (Wuhan University of Technology) (SYSJJ2020-05), and the Program of the Housing and Urban-Rural Construction Department of Shandong Province (2019-K7-10).

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Martin, L.W.; Ramesh, R. Overview No. 151 Multiferroic and magnetoelectric heterostructures. Acta Mater. 2012, 60, 2449–2470. [Google Scholar] [CrossRef] [Green Version]
  2. Eerenstein, W.; Mathur, N.D.; Scott, J.F. Multiferroic and magnetoelectric materials. Nature 2006, 442, 759–765. [Google Scholar] [CrossRef]
  3. Bibes, M.; Barthelemy, A. Multiferroics: Towards a magnetoelectric memory. Nat. Mater. 2008, 7, 425–426. [Google Scholar] [CrossRef]
  4. Yue, Z.W.; Tan, G.Q.; Yang, W.; Ren, H.J.; Xiao, A. Enhanced multiferroic properties in Pr-doped BiFe0.97Mn0.03O3 films. Ceram. Int. 2016, 42, 18692–18699. [Google Scholar] [CrossRef]
  5. Sreenivasulu, G.; Laletin, U.; Petrov, V.M.; Petrov, V.V.; Srinivasan, G. A permendur-piezoelectric multiferroic composite for low-noise ultrasensitive magnetic field sensors. Appl. Phys. Lett. 2012, 100, 173506. [Google Scholar] [CrossRef]
  6. Hasegawa, M.; Asano, T.; Hashimoto, K.; Lee, G.C.; Park, Y.C.; Okazaki, T.; Furuya, Y. Fabrication of multiferroic composite actuator material by combining superelastic TiNi filler and a magnetostrictive Ni matrix. Smar. Mate. Stru. 2006, 15, N124–N128. [Google Scholar] [CrossRef]
  7. Huang, A.; Handoko, A.D.; Goh, G.K.; Pallathadka, P.K.; Shannigrahi, S. Hydrothermal synthesis of (001) epitaxial BiFeO3 films on SrTiO3 substrate. CrystEngComm 2010, 12, 3806–3814. [Google Scholar] [CrossRef]
  8. Lebeugle, D.; Colson, D.; Forget, A.; Viret, M.; Bonville, P.; Marucco, J.F.; Fusil, S. Room-temperature coexistence of large electric polarization and magnetic order in BiFeO3 single crystals. Phys. Rev. B 2007, 76, 024116. [Google Scholar] [CrossRef] [Green Version]
  9. Prashanthi, K.; Shaibani, P.M.; Sohrabi, A.; Natarajan, T.S.; Thundat, T. Nanoscale magnetoelectric coupling in multiferroic BiFeO3 nanowires. Phys. Status Solidi RRL 2012, 6, 244–246. [Google Scholar] [CrossRef]
  10. Shami, M.Y.; Awan, M.S.; Anis-ur-Rehman, M. The effect of heat treatment on structural and multiferroic properties of phase-pure BiFeO3. J. Electron. Mater. 2012, 41, 2216–2224. [Google Scholar] [CrossRef]
  11. Singh, S.K. Structural and electrical properties of Sm-substituted BiFeO3 thin films prepared by chemical solution deposition. Thin Solid Film. 2013, 527, 126–132. [Google Scholar] [CrossRef]
  12. Zhou, J.; Trassin, M.; He, Q.; Tamura, N.; Kunz, M.; Cheng, C.; Wu, J. Directed assembly of nano-scale phase variants in highly strained BiFeO3 thin films. J. Appl. Phys. 2012, 112, 064102. [Google Scholar] [CrossRef] [Green Version]
  13. Reetu, A.; Sanghi, S. Rietveld analysis, dielectric and magnetic properties of Sr and Ti codoped BiFeO3 multiferroic. J. Appl. Phys. 2011, 110, 073909. [Google Scholar] [CrossRef]
  14. Qi, X.; Dho, J.; Tomov, R.; Blamire, M.G.; Blamire, J.L. Greatly reduced leakage current and conduction mechanism in aliovalent-ion-doped BiFeO3 Macmanus-driscoll. Appl. Phys. Lett. 2005, 86, 062903. [Google Scholar] [CrossRef]
  15. Gumiel, C.; Jardiel, T.; Calatayud, D.G.; Vranken, T.; Van Bael, M.K.; Hardy, A.; Peiteado, M. Nanostructure stabilization by low-temperature dopant pinning in multiferroic BiFeO3-based thin films produced by aqueous chemical solution deposition. J. Mater. Chem. C 2020, 8, 4234–4245. [Google Scholar] [CrossRef] [Green Version]
  16. Madolappa, S.; Kundu, S.; Bhimireddi, R.; Varma, K.B. Improved electrical characteristics of Pr-doped BiFeO3 ceramics prepared by sol-gel route. Mater. Res. Express 2016, 3, 065009. [Google Scholar] [CrossRef]
  17. Shimada, T.; Arisue, K.; Kitamura, T. Strain-induced phase transitions in multiferroic BiFeO3 (110) epitaxial film. Phys. Lett. A 2012, 376, 3368–3371. [Google Scholar] [CrossRef]
  18. Yang, J.C.; He, Q.; Suresha, S.J.; Kuo, C.Y.; Peng, C.Y.; Haislmaier, R.C.; Chu, Y.H. Orthorhombic BiFeO3. Phys. Rev. Lett. 2012, 109, 247606. [Google Scholar] [CrossRef] [Green Version]
  19. Chai, Z.; Tan, G.; Yue, Z.; Yang, W.; Guo, M.; Ren, H.; Lv, L. Ferroelectric properties of BiFeO3 thin films by Sr/Gd/Mn/Co multi-doping. J. Alloys Compd. 2018, 746, 677–687. [Google Scholar] [CrossRef]
  20. Wen, X.L.; Chen, Z.; Liu, E.H.; Lin, X.; Chen, C. Effect of Ba and Mn doping on microstructure and multiferroic properties of BiFeO3 ceramics. J. Alloys Compd. 2016, 678, 511–517. [Google Scholar] [CrossRef]
  21. Karpinsky, D.V.; Pakalniškis, A.; Niaura, G.; Zhaludkevich, D.V.; Zhaludkevich, A.L.; Latushka, S.I.; Kareiva, A. Evolution of the crystal structure and magnetic properties of Sm-doped BiFeO3 ceramics across the phase boundary region. Ceram. Int. 2021, 47, 5399–5406. [Google Scholar] [CrossRef]
  22. Sati, P.C.; Kumar, M.; Chhoker, S. Phase evolution, magnetic, optical and dielectric properties of Zr-substituted Bi0.9Gd0.1FeO3 multiferroics. J. Am. Ceram. Soc. 2015, 98, 1884–1890. [Google Scholar] [CrossRef]
  23. Yun, Q.; Bai, Y.L.; Chen, J.; Gao, W.; Bai, A.; Zhao, S. Improved ferroelectric and fatigue properties in Ho doped BiFeO3 thin films. Mater. Lett. 2014, 129, 166–169. [Google Scholar] [CrossRef]
  24. Liu, J.; Deng, H.M.; Zhai, X.; Lin, T.; Meng, X.; Zhang, Y.; Chu, J. Influence of Zn doping on structural, optical and magnetic properties of BiFeO3 films fabricated by the sol-gel technique. Mater. Lett. 2014, 133, 49–52. [Google Scholar] [CrossRef]
  25. Yang, S.J.; Zhang, F.Q.; Xie, X.; Sun, H.; Zhang, L.; Fan, S. Enhanced leakage and ferroelectric properties of Zn-doped BiFeO3 thin films grown by sol-gel method. J. Alloys Compd. 2018, 734, 243–249. [Google Scholar] [CrossRef]
  26. Zhang, C.C.; Dai, J.Q.; Liang, X.L. Enhanced ferroelectric properties of (Zn, Ti) equivalent co-doped BiFeO3 films prepared via the sol-gel method. Ceram. Int. 2021, 47, 16776–16785. [Google Scholar] [CrossRef]
  27. Liu, Y.; Tan, G.Q.; Ren, X.X.; Li, J.; Xue, M.; Ren, H.; Liu, W. Electric field dependence of ferroelectric stability in BiFeO3 thin films co-doped with Er and Mn. Ceram. Int. 2020, 46, 18690–18697. [Google Scholar] [CrossRef]
  28. Kan, D.; Pálová, L.; Anbusathaiah, V.; Cheng, C.J.; Fujino, S.; Nagarajan, V.; Takeuchi, I. Universal behavior and electric-field-induced structural transition in rare-earth-substituted BiFeO3. Adv. Funct. Mater. 2010, 20, 1108–1115. [Google Scholar] [CrossRef]
  29. Emery, S.B.; Cheng, C.J.; Kan, D.; Rueckert, F.J.; Alpay, S.P.; Nagarajan, V.; Wells, B.O. Phase coexistence near a morphotropic phase boundary in Sm-doped BiFeO3 films. Appl. Phys. Lett. 2010, 97, 152902. [Google Scholar] [CrossRef] [Green Version]
  30. Xue, X.; Tan, G.Q.; Ren, H.J. Structural, electric and multiferroic properties of Sm-doped BiFeO3 thin films prepared by the sol-gel process. Ceram Int. 2013, 39, 6223–6228. [Google Scholar] [CrossRef]
  31. Tao, H.; Lv, J.; Zhang, R.; Xiang, R.; Wu, J. Lead-free rare earth-modified BiFeO3 ceramics: Phase structure and electrical properties. Mater. Des. 2017, 120, 83–89. [Google Scholar] [CrossRef] [Green Version]
  32. Zhang, F.; Zeng, X.; Bi, D.; Guo, K.; Yao, Y.; Lu, S. Dielectric, ferroelectric, and magnetic properties of Sm-doped BiFeO3 ceramics prepared by a modified solid-state-reaction method. Materials 2018, 11, 2208. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  33. Liang, X.L.; Dai, J.Q. Prominent ferroelectric properties in Mn-doped BiFeO3 spin-coated thin films. J. Alloys Compd. 2021, 886, 161168. [Google Scholar] [CrossRef]
  34. Zhou, W.; Deng, H.; Cao, H.; He, J.; Liu, J.; Yang, P.; Chu, J. Effects of Sm and Mn co-doping on structural, optical and magnetic properties of BiFeO3 films prepared by a sol-gel technique. Mater. Lett. 2015, 144, 93–96. [Google Scholar] [CrossRef]
  35. Goian, V.; Kamba, S.; Greicius, S.; Nuzhnyy, D.; Karimi, S.; Reaney, I. Terahertz and infrared studies of antiferroelectric phase transition in multiferroic Bi0. 85Nd0.15FeO3. J. Appl. Phys. 2011, 110, 074112. [Google Scholar] [CrossRef] [Green Version]
  36. Gu, Y.; Zhou, Y.; Zhang, W. Optical and magnetic properties of Sm-doped BiFeO3 nanoparticles around the morphotropic phase boundary region. AIP Adv. 2021, 11, 045223. [Google Scholar] [CrossRef]
  37. Li, W.; Hao, J.; Du, J.; Fu, P.; Sun, W.; Chen, C.; Chu, R. Electrical properties and luminescence properties of 0.96(K0.48Na0.52)(Nb0.95Sb0.05)–0.04Bi0.5(Na0.82K0.18)0.5ZrO3-xSm lead-free ceramics. J. Adv. Ceram. 2020, 9, 72–82. [Google Scholar] [CrossRef] [Green Version]
  38. Singh, M.K.; Jang, H.M.; Ryu, S.; Jo, M.H. Polarized Raman scattering of multiferroic BiFeO3 epitaxial films with rhombohedral R3c symmetry. Appl. Phys. Lett. 2006, 88, 042907. [Google Scholar] [CrossRef]
  39. Wang, Y.; Nan, C.W. Site modification in BiFeO3 thin films studied by Raman spectroscopy and piezoelectric force microscopy. J. Appl. Phys. 2008, 103, 114104. [Google Scholar] [CrossRef]
  40. Iliev, M.N.; Abrashev, M.V.; Lee, H.G.; Popov, V.N.; Sun, Y.Y.; Thomsen, C.; Chu, C.W. Raman spectroscopy of orthorhombic perovskitelike YMnO3 and LaMnO3. Phys. Rev. B 1998, 57, 2872. [Google Scholar] [CrossRef]
  41. Singh, D.; Tabari, T.; Ebadi, M.; Trochowski, M.; Yagci, M.B.; Macyk, W. Efficient synthesis of BiFeO3 by the microwave-assisted sol-gel method: “A” site influence on the photoelectron chemical activity of perovskites. Appl. Surf. Sci. 2019, 471, 1017–1027. [Google Scholar] [CrossRef]
  42. Wang, J.; Luo, L.; Han, C.; Yun, R.; Tang, X.; Zhu, Y.; Feng, Z. The microstructure, electric, optical and photovoltaic properties of BiFeO3 thin films prepared by low temperature sol-gel method. Materials 2019, 12, 1444. [Google Scholar] [CrossRef] [Green Version]
  43. Ma, Z.B.; Liu, H.Y.; Wang, L.X.; Zhang, F.Q.; Zhang, F.; Zhu, L.; Fan, S. Phase transition and multiferroic properties of Zr-doped BiFeO3 thin films. J. Mater. Chem. C 2020, 48, 17307–17317. [Google Scholar] [CrossRef]
  44. Bai, H.; Li, J.; Hong, Y.; Zhou, Z. Enhanced ferroelectricity and magnetism of quenched (1−x) BiFeO3-xBaTiO3 ceramics. J. Adv. Ceram. 2020, 9, 511–516. [Google Scholar] [CrossRef]
  45. Hanani, Z.; Merselmiz, S.; Danine, A.; Stein, N.; Mezzane, D.; Amjoud, M.B. Enhanced dielectric and electrocaloric properties in lead-free rod-like BCZT ceramics. J. Adv. Ceram. 2020, 9, 210–219. [Google Scholar] [CrossRef] [Green Version]
  46. Lou, Y.H.; Song, G.L.; Chang, F.G. Investigation on dependence of BiFeO3 dielectric property on oxygen content. Chin. Phys. B 2010, 19, 077702. [Google Scholar]
  47. Ren, X.; Otsuka, K. Universal symmetry property of point defects in crystals. Phys. Rev. Lett. 2000, 85, 1016. [Google Scholar] [CrossRef] [Green Version]
  48. Ren, X. Large electric-field-induced strain in ferroelectric crystals by point-defect-mediated reversible domain switching. Nat. Mater. 2004, 3, 91–94. [Google Scholar] [CrossRef] [PubMed]
  49. Zhang, L.; Ren, X. Aging behavior in single-domain Mn-doped BaTiO3 crystals: Implication for a unified microscopic explanation of ferroelectric aging. Phys. Rev. B 2006, 73, 094121. [Google Scholar] [CrossRef] [Green Version]
  50. Zhang, L.X.; Ren, X. In situ observation of reversible domain switching in aged Mn-doped BaTiO3 single crystals. Phys. Rev. B 2005, 71, 174108. [Google Scholar] [CrossRef] [Green Version]
  51. Guo, Y.Y.; Yan, Z.B.; Zhang, N.; Cheng, W.W.; Liu, J.M. Ferroelectric aging behaviors of BaTi0.995Mn0.005O3 ceramics: Grain size effects. Appl. Phys. A 2012, 107, 243–248. [Google Scholar] [CrossRef]
Figure 1. (a) XRD patterns, (b) magnified patterns around 30.5–33°, and (c) magnified patterns around 39–41° of the BSFM films with different Sm content.
Figure 1. (a) XRD patterns, (b) magnified patterns around 30.5–33°, and (c) magnified patterns around 39–41° of the BSFM films with different Sm content.
Nanomaterials 12 00108 g001
Figure 2. Rietveld refined XRD patterns of the BSFM films with different Sm content: (a) x = 0 mol%; (b) x = 2 mol%; (c) x = 4 mol%; (d) x = 6 mol%.
Figure 2. Rietveld refined XRD patterns of the BSFM films with different Sm content: (a) x = 0 mol%; (b) x = 2 mol%; (c) x = 4 mol%; (d) x = 6 mol%.
Nanomaterials 12 00108 g002
Figure 3. SEM images of the surface morphological features of the BSFM films: (a) x = 0 mol%; (b) x = 2 mol%; (c) x = 4 mol%; (d) x = 6 mol%, and (e) the cross-sectional image of the BSFMx = 0.06 film. The insets show the distribution of crystal grain size and the average grain diameter (Dave).
Figure 3. SEM images of the surface morphological features of the BSFM films: (a) x = 0 mol%; (b) x = 2 mol%; (c) x = 4 mol%; (d) x = 6 mol%, and (e) the cross-sectional image of the BSFMx = 0.06 film. The insets show the distribution of crystal grain size and the average grain diameter (Dave).
Nanomaterials 12 00108 g003
Figure 4. (a) Raman spectra of the BSFM films with different Sm content and crystal structure of the BSFM films: (b) Pnma; (c) R3c, Raman fitting spectra of the BSFM films with different Sm content: (d) x = 0 mol%; (e) x = 2 mol%; (f) x = 4 mol%; (g) x = 6 mol%. # refers to the Raman frequencies of the Pnma structure.
Figure 4. (a) Raman spectra of the BSFM films with different Sm content and crystal structure of the BSFM films: (b) Pnma; (c) R3c, Raman fitting spectra of the BSFM films with different Sm content: (d) x = 0 mol%; (e) x = 2 mol%; (f) x = 4 mol%; (g) x = 6 mol%. # refers to the Raman frequencies of the Pnma structure.
Nanomaterials 12 00108 g004
Figure 5. Fe2p3/2 orbital XPS fitting diagram of the BSFM films with different Sm content: (a) x = 0 mol%; (b) x = 2 mol%; (c) x = 4 mol%; (d) x = 6 mol%. The black lines refer to the original curve, the red lines refer to the fitting curve, and the blue lines refer to the background line.
Figure 5. Fe2p3/2 orbital XPS fitting diagram of the BSFM films with different Sm content: (a) x = 0 mol%; (b) x = 2 mol%; (c) x = 4 mol%; (d) x = 6 mol%. The black lines refer to the original curve, the red lines refer to the fitting curve, and the blue lines refer to the background line.
Nanomaterials 12 00108 g005
Figure 6. O1s orbital XPS fitting diagram of the BSFM films with different Sm content: (a) x = 0 mol%; (b) x = 2 mol%; (c) x = 4 mol%; (d) x = 6 mol%. The black lines refer to the original curve, the red lines refer to the fitting curve, and the blue lines refer to the background line.
Figure 6. O1s orbital XPS fitting diagram of the BSFM films with different Sm content: (a) x = 0 mol%; (b) x = 2 mol%; (c) x = 4 mol%; (d) x = 6 mol%. The black lines refer to the original curve, the red lines refer to the fitting curve, and the blue lines refer to the background line.
Nanomaterials 12 00108 g006
Figure 7. Leakage current density curves of the BSFM films with different Sm content.
Figure 7. Leakage current density curves of the BSFM films with different Sm content.
Nanomaterials 12 00108 g007
Figure 8. (a) P−E hysteresis loops, (b) dielectric permittivity, and (c) dielectric loss of the BSFM films with different Sm content.
Figure 8. (a) P−E hysteresis loops, (b) dielectric permittivity, and (c) dielectric loss of the BSFM films with different Sm content.
Nanomaterials 12 00108 g008
Figure 9. Comparison of hysteresis loops of the BSFM films with different Sm content measured after an interval of 110 days: (a) x = 4 mol%; (b) x = 6 mol%.
Figure 9. Comparison of hysteresis loops of the BSFM films with different Sm content measured after an interval of 110 days: (a) x = 4 mol%; (b) x = 6 mol%.
Nanomaterials 12 00108 g009
Table 1. Rietveld refinement parameters of the BSFM films with different Sm content.
Table 1. Rietveld refinement parameters of the BSFM films with different Sm content.
SampleSpace GroupFraction (%)Lattice ParameterRw (%)
a (Å)b (Å)c (Å)α (°)β (°)γ (°)Volume
BSFMx = 0R3c70.985.595.5913.709090120370.417.90
Pnma29.025.6016.0111.289090901010.60
BSFMx = 0.02R3c68.395.585.5813.709090120369.566.83
Pnma31.615.6215.9811.269090901010.75
BSFMx = 0.04R3c69.065.585.5813.669090120368.677.40
Pnma30.945.5816.0611.309090901012.96
BSFMx = 0.06R3c67.065.585.5813.679090120368.657.17
Pnma32.945.5616.0611.289090901007.00
Table 2. The Raman modes of the BSFM films with different Sm content.
Table 2. The Raman modes of the BSFM films with different Sm content.
Raman Modes (cm−1)BSFMx = 0BSFMx = 0.02BSFMx = 0.04BSFMx = 0.06
E76.6776.9977.39278.053
E112.61112.45112.03112.80
A1-1141.76141.88142.64143.31
A1-2171.43171.05172.11172.89
pnma200.74201.38203.34205.78
A1-3226.61227.47230.80234.49
E251.23253.53259.69264.92
E273.72275.8286.47---
pnma298.05299.78313.95294.87
E348.13353.32343.83---
E372.68380.91374.02361.13
pnma397.33411.95406.72394.90
A1-4428.90442.21440.26432.93
E459.09469.92470.06465.64
pnma484.76496.11497.29495.72
E535.62525.24527.48527.32
E600.15587.81593.70595.79
pnma624.89616.55622.52621.96
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Wang, Y.; Li, Z.; Ma, Z.; Wang, L.; Guo, X.; Liu, Y.; Yao, B.; Zhang, F.; Zhu, L. Phase Structure and Electrical Properties of Sm-Doped BiFe0.98Mn0.02O3 Thin Films. Nanomaterials 2022, 12, 108. https://doi.org/10.3390/nano12010108

AMA Style

Wang Y, Li Z, Ma Z, Wang L, Guo X, Liu Y, Yao B, Zhang F, Zhu L. Phase Structure and Electrical Properties of Sm-Doped BiFe0.98Mn0.02O3 Thin Films. Nanomaterials. 2022; 12(1):108. https://doi.org/10.3390/nano12010108

Chicago/Turabian Style

Wang, Yangyang, Zhaoyang Li, Zhibiao Ma, Lingxu Wang, Xiaodong Guo, Yan Liu, Bingdong Yao, Fengqing Zhang, and Luyi Zhu. 2022. "Phase Structure and Electrical Properties of Sm-Doped BiFe0.98Mn0.02O3 Thin Films" Nanomaterials 12, no. 1: 108. https://doi.org/10.3390/nano12010108

APA Style

Wang, Y., Li, Z., Ma, Z., Wang, L., Guo, X., Liu, Y., Yao, B., Zhang, F., & Zhu, L. (2022). Phase Structure and Electrical Properties of Sm-Doped BiFe0.98Mn0.02O3 Thin Films. Nanomaterials, 12(1), 108. https://doi.org/10.3390/nano12010108

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