Next Article in Journal
A New 1D Ni (II) Coordination Polymer of s-Triazine Type Ligand and Thiocyanate as Linker via Unexpected Hydrolysis of 2,4-Bis(3,5-dimethyl-1H-pyrazol-1-yl)-6-methoxy-1,3,5-triazine
Previous Article in Journal
Photocatalytic Reduction of Cr(VI) to Cr(III) and Photocatalytic Degradation of Methylene Blue and Antifungal Activity of Ag/TiO2 Composites Synthesized via the Template Induced Route
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Multiferroics Made via Chemical Co-Precipitation That Is Synthesized and Characterized as Bi(1−x)CdxFeO3

1
Department of Physics, Faculty of Sciences, University of Gujrat, Gujrat 50700, Pakistan
2
Department of Mathematics, University of Gujrat, Gujrat 50700, Pakistan
3
Department of Mathematics, Al-Qunfudah University College, Umm Al-Qura University, Mecca 24382, Saudi Arabia
4
Center of Research, Faculty of Engineering, Future University in Egypt, New Cairo 11835, Egypt
5
Department of Mechanical Engineering, College of Engineering in Wadi Alddawasir, Prince Sattam bin Abdulaziz University, Wadi Addawaser 11991, Saudi Arabia
6
Production Engineering and Mechanical Design Department, Faculty of Engineering, Mansoura University, Mansoura 35516, Egypt
*
Authors to whom correspondence should be addressed.
Inorganics 2023, 11(3), 134; https://doi.org/10.3390/inorganics11030134
Submission received: 17 December 2022 / Revised: 9 March 2023 / Accepted: 12 March 2023 / Published: 21 March 2023
(This article belongs to the Special Issue First-Row Transition Metal-Based Catalysts for Water Oxidation)

Abstract

:
Cd-doped BiFeO3 powders, with varying doping concentrations of Cd (Bi(1−x)CdxFeO3, where x = 0–0.3), were prepared through a facile chemical co-precipitation method and calcinated at 550 °C in the air. The BiFeO3 has a rhombohedral crystal structure, which changes to an orthorhombic crystal structure with an increase in Cd doping. The presence of dopant has also altered the bandgap of material suppressing it from 2.95 eV to 2.51 eV, improving the visible light absorption. Vibrating sample magnetometry (VSM) confirmed stronger ferromagnetic character for Bi0.7Cd0.3FeO3 with a coercivity of 250 Oe, and remnant magnetization was 0.15 emu/g, which is because of the misalignment of the two sublattices of perovskite structure after doping resulting in the imbalanced magnetic moment giving rise to net nonzero magnetic behavior. The particle size reduction is observed with an increase in the doping concentration of Cd.

1. Introduction

In the recent past, multiferroics have attained tremendous attention from researchers because of their vital role in the manufacturing of memory storage devices, spintronics, and magneto-electric sensors [1,2]. Multiferroics are materials possessing both ferromagnetic and ferroelectric properties at the same time. Generally, a material should be a metal to show good ferromagnetic behavior, or it should be an insulator/dielectric (to become polarized) to exhibit the phenomenon of ferroelectricity. This makes multiferroics rare and valuable materials for the electronics industry [3,4,5]. To date, many multiferroics have been reported. The general universal formula ABX3 represents the family of materials having a perovskite structure. Here, A and B represent the cations, and X is the anion. X can vary from oxides to halides, which means there can be an oxide element or halide group element, for example, Br, Cl, or I2, which can be used as halides. A wide range of perovskite materials, including BiFeO3, BiMnO3, and ZMnO3, where Z can be a rare earth metal such as Y, Tb, Ho, Lu, etc., are multiferroic. Other examples include BaTiO3, which is dielectric ferroelectric and is used in multilayer ceramic capacitors and sensors. PbTiO3 possesses pyroelectric and piezoelectric properties and is used in transducers for marine and underwater devices and is also used as a protector. Na0.5Bi0.5 is reported for showing ferroelectricity and is used in piezoceramics.
Among the above-mentioned multiferroics, BiFeO3 is a significant compound exhibiting relatively high Neel temperature (370 °C) and Curie temperature (830 °C), which is the main point of interest for potential applications in the industry where exposure to high temperatures [6,7]. Secondly, BiFeO3 is more resistant to rust, as well. BiFeO3 is a well known ferrite, and research is performed with many different doping elements, for instance, Dzik et al. [8] examined the Dy doping in the place of Bi in BiFeO3 and observed the enhanced magnetic properties of the sample from 0.1 emu/g for the pure sample and 0.3 emu/g for the 10% doped sample when measured at room temperature (300 K). Dy also increased the measuring field frequency value for the sample at all temperatures. The obtained samples were in ceramic form when prepared with the mixed oxide method. Additionally, Ishaq et al. [9] investigated the effect of Mn-Cd co-doping in BiFeO3 and confirmed the grain size reduction of the samples with increasing doping concentration of dopants. The structural transition of pure bismuth ferrite crystal from rhombohedral to orthorhombic occurred whenever the bismuth ferrite particles were doped with a suitable metallic element in the Bi site. Moreover, D.V. Karpinsky et al. [10] reported the maximum remnant magnetization of 0.07 emu/g for Mn and Ca co-doping in BiFeO3 samples, and the antiferromagnetic behavior is observed. Arfat et al. [11] reported the specific magnetization lay around the value of 2.9 emu/g associated with 10% doping of Ni in BiFeO3. This value of specific magnetization increased significantly when increasing the value of x, which is the value of the concentration of doping.
BiFeO3 is a perovskite having a distorted rhombohedral structure lying in the R3c space group. It shows relatively higher electrical conductivity due to the presence of Bi2Fe4O9 and Bi25FeO40 secondary phases [12]. The volatilization of Fe2+ ions and Bi3+ ions at high temperatures also plays a key role in the conduction of electric current. The synthesis of single-phase bismuth ferrites is very hard due to the formation of secondary phases, which simply cannot be avoided by rapid sintering [13,14,15,16,17]. The literature suggests that the substitution of rare earth ions or alkali earth ions, such as Y3+, Nd3+, Sm3+, Sr2+, and Ba2+ can improve some properties, including magnetic and electric properties of BiFeO3 [18,19,20]. Further, the reduction in the leakage current can be achieved by suppressing the spiral structure of BiFeO3 perovskite and reducing its particle size [21,22,23].
Bismuth causes space charge defects (such as oxygen vacancies) in the perovskite structure. These free-to-move charges gather at lower energy places (such as ferroelectric domain walls) or grain boundaries when an external electric field is applied [24]. Consequently, BiFeO3 leakage current (due to secondary phases) and oxygen vacancies limit the ferroelectric loop by gathering on the domain walls. This sets up a limit for multiferroics to be efficiently used in magneto-electric devices [25,26,27]. The Fe3+ ions (partially filling the 3d orbital) are the main cause of G-type antiferromagnetic ordering, resulting from the Jahn Teller effect during structure distortion of the bulk BiFeO3 [28,29].
Generally, BiFeO3 can be synthesized by several methods, such as the solid-state method, sol-gel process, hydrothermal technique, and chemical co-precipitation method. In the solid-state method, precursors are manually mixed, and nitric acid is mostly used for leaching to avoid the impurity phases [30,31,32]. Similarly, leaching methods are also used in the sol-gel route to form ceramics of BiFeO3. BiFeO3 samples with Cd doping yield a relatively low number of undesired secondary phases (responsible for poor electrical and magnetic properties) in multiferroics. A lower quantity of undesired secondary phases of the final product is the main concern leading towards particularly selecting the chemical coprecipitation method [33,34,35,36]. Based upon the requirement of the final powder, several synthesis routes have been adopted by different researchers to produce BiFeO3. Among these methods, some are suitable to achieve small particles size, while others are useful to produce a powder with high purity levels [37,38]. The particle size of a material can have a significant influence on its magnetic behavior. In general, as the particle size decreases, the magnetic properties of the material tend to become more pronounced. This is due to the increased surface-to-volume ratio of smaller particles, which leads to a higher density of surface defects and an increased number of unpaired electrons on the surface. These factors can lead to a higher susceptibility to magnetic fields and a greater overall magnetic moment.
For ferromagnetic materials, as the particle size decreases, the Curie temperature (Tc) tends to decrease, which means that ferromagnetic ordering can occur at lower temperatures. Additionally, the coercive field (Hc) tends to increase, meaning that it takes a stronger magnetic field to switch the magnetization direction.
In addition, it is also important to consider the shape of the particles as well. Spherical particles will have different magnetic properties than elongated particles of the same size. It is worth noting that the particle size effect on magnetic behavior is not always linear, and sometimes more complex relations can be observed. Preparation of BiFeO3 particles can also be achieved using the hydrothermal method, but the resulting particle size is sometimes larger than several hundred nanometers. Preparation of BiFeO3 nano-crystals at elevated temperatures (up to 200 °C) using KOH as precipitating agent is reported by Chen et al. [39]. Conventional synthesis routes for preparing BiFeO3 particles generally involve the mixing of precursors in stoichiometric amounts, and the reaction rate largely depends upon the particle size and temperature. Chemical co-precipitation, on the other hand, is a well known method for obtaining particles with relatively smaller sizes and higher degrees of homogenization. Further, this technique does not involve very high temperatures or heavy apparatus, while many physical properties of the powder can be modified by controlling the external conditions, such as temperature, stirring period, stirring speed, amount, and type of precipitating agent. The volatilization of both Fe2+ ions and Bi3+ ions at high temperatures can play a key role in the conduction of electric current in BiFeO3. However, it is true that, in some studies, the problem of Bi volatilization during synthesis is emphasized, but the volatilization of Fe ions is not mentioned. This is likely because the volatilization of Bi ions is more pronounced and can have a greater impact on the properties of the final material. Additionally, the OZDEMIR methods [40] used to prevent Bi volatilization, such as using a reducing atmosphere or low-temperature synthesis, may also help to reduce the volatilization of Fe ions.
Therefore, in this work, a simple, facile, and efficient chemical co-precipitation technique was successfully used to synthesize BiFeO3 with different doping concentrations of Cd (Bi1−xCdxFeO3, where x = 0, 0.1, 0.3). The electrical, magnetic, optical, and structural properties of the ferrites were systematically studied through vibrating sample magnetometry, X-ray diffraction, scanning electron microscopy, EDX analysis, ultraviolet-visible spectroscopy, and Raman spectroscopy.

2. Experimental Procedure

Bi1−xCdxFeO3 (x = 0, 0.1, 0.3) samples were prepared via a simple chemical co-precipitation method using high-purity (<99%) precursors. The nitrates of metallic salts [Bi(NO3)3·5H2O, Cd(NO3)2·4H2O, Fe (NO3)3·9H2O] were used in their as-received state with no further processing. All precursors (in the stoichiometric amount) were mixed in deionized water at room temperature. The initial solution of each nitrate was stirred up until it was completely dissolved. After preparing the solutions of each reagent, they were mixed in a beaker under continuous stirring, and the temperature of the solution was slightly elevated to 60 °C. Nitric acid and polyethylene glycol (PEG) were added dropwise in a small ratio to control the morphology of particles. Further, nitric acid was used as a bleaching agent to avoid nucleation of impurity phases. Subsequently, ammonia solution was added dropwise to settle down the particles in the beaker. The whole process was carried out under constant stirring for 3 h without reducing the temperature. Finally, the prepared ferrite particles were allowed to settle down at room temperature and washed three to four times with deionized water. The pH of the resulting liquid precursor was kept between 7–9 with the help of ammonia solution and washing process. The resulting solution was then dried in an oven maintained at 80 °C. Finally, all dried powder precursors were calcined at 550 °C for 3 h in air. The final product was then ground to obtain a homogenous fluffy powder.
The co-precipitation synthesis of BiFeO3 nanoparticles can be represented by the following chemical reaction equations:
  • Formation of metal precursors:
Bismuth nitrate pentahydrate (Bi(NO3)3·5H2O) + iron nitrate nonahydrate (Fe(NO3)3·9H2O) → bismuth iron nitrate hexahydrate (BiFe(NO3)6·14H2O)
  • Precipitation of metal oxides:
BiFe(NO3)6·14H2O + sodium hydroxide (NaOH) → BiFeO3 (s) + sodium nitrate (NaNO3) + 14H2O (l)
  • Calcination of metal oxides:
BiFeO3 (s) → BiFeO3 (s) (calcined)
The calcination step is usually performed at high temperatures (e.g., 500–600 °C) to remove any residual water and to obtain crystalline BiFeO3 nanoparticles.
The structural and morphological characteristics of the prepared samples were revealed by X-ray diffraction using the GNR EXPLORER 2000 (XRD) and the scanning electron microscope TESCAN MAIA3 (SEM), respectively. The XRD was carried out using Cu-Kα radiations (with wavelength 1.54 Å) in the 2-theta range of 20–60°. To study the magnetic properties of the prepared powders, vibrating sample magnetometry (VSM) (Lake Shore 7300) was performed. Moreover, ultraviolet-visible UV1800 (UV) spectroscopy and Raman spectroscopy DPSSL laser source (with lambda = 532 nm/150 mW), as an excitation source (RS), were performed to study the absorption peaks/band gap, as well as the emission spectra of the prepared samples. Fourier transform-infrared spectroscopy FT/IR-4100 type-A, JASCO (FT-IR) was performed to analyze the bonding of oxygen with cations and the effect of doping on bonding.

3. Results and Discussion

3.1. X-ray Diffraction

Figure 1 shows the XRD patterns of BiFeO3, Bi0.9Cd0.1FeO3, and Bi0.7Cd0.3FeO3 compounds, which were calcined at 550 °C for 3 h in the powder form. The XRD pattern confirms the rhombohedral structure of the BiFeO3 sample that belongs to the R3c space group. Interestingly, no secondary phase of bismuth ferrite was found in the pure sample. The main peak is observed at 2-theta ~32°, which corresponds to 110 crystallographic planes. After the substitution of Cd in BiFeO3, a structural transition from rhombohedral to orthorhombic is observed, as shown in Figure 1, showing the characteristic dual peaks near ~32°, which are attributed to the orthorhombic phase. The peaks at 2-theta ~46° and ~53° splits into two sub-peaks, and the main peak with the highest intensity tends to broaden, which confirms the successful substitution of Cd with Bi in BiFeO3. This broadening of peaks increases with the increase in Cd doping. This might be associated with the fact that the ionic radii of Bi and Cd are different. The ionic radius of Cd+2 is 95 pm and has a smaller ionic radius than that of Bi+2, which is 116 pm under the same conditions. Therefore, a transition from rhombohedral to the orthorhombic type of structure took place. Moreover, the Bi2Fe4O9 impurity phase and Bi25FeO40 secondary phase appear in Bi0.9Cd0.1FeO3 and Bi0.7Cd0.3FeO3 samples, as shown in Figure 1. [41]. The impact of Cd-substitution on the lattice parameters and magnetic properties of BiFeO3 can vary depending on the substitution level and other factors, such as temperature and pressure. Generally, Cd-substitution can cause a decrease in the lattice parameters, which results in a decrease in the magnetic properties, as the decrease in lattice parameters can lead to a reduction in the exchange interactions between the magnetic moments in the material. The selenite-type structure is a crystallographic structure that is found in certain types of ferrites. The XRD pattern of ferrite with a silent-type phase typically exhibits characteristic diffraction peaks that can be used to identify the presence of this structure. In Figure 1, the peak between 2theta 25–30 depicts the presence of a selenite-type phase.
Debye Scherrer’s formula, mentioned in Equation (1), is used to calculate the crystallite size of fabricated materials.
D = (kλ/β cos θ)
Here, D represents the crystallite size, k is the shape constant, lambda (λ) is the wavelength of incident X-rays, theta (θ) is the angle of diffraction, and β indicates the full-width half maxima (FWHM) of the characteristic peak under consideration for calculations. The crystalline size of the prepared materials has been shown in Table 1.

3.2. Scanning Electron Microscopy

Figure 2 presents the SEM images of BiFeO3, Bi0.9Cd0.1FeO3, and Bi0.7Cd0.3FeO3 powder samples prepared with the chemical co-precipitation method. The SEM images clearly show that the pure BiFeO3 has cubic crystals with a relatively large average particle size of ~0.8 µm. As the concentration of Cd dopant in BiFeO3 is increased, the particle size is considerably decreased. Moreover, a transition from cubic crystals to spherical-shaped particles is observed with Cd doping (Figure 2). The average particle size of Bi0.9Cd0.1FeO3 was calculated to be 0.4 µm with spherical morphology. In the case of Bi0.7Cd0.3FeO3, the particle size was further decreased to ~0.25 µm, while the particles retained their spherical morphology. EDX analysis of Bi0.7Cd0.3FeO3, mentioned in Figure 3, confirms the successful doping of Cd into BiFeO3. It also provides valuable information about the percentage composition of the material. It is worth mentioning that particle size is different from crystallite size determined by XRD analysis mentioned in Table 1. Crystallite size refers to the average size of the crystalline regions in a material, while particle size refers to the size of the individual particles or grains in the material. Crystallite size affects the material’s mechanical, optical, and electrical properties, whereas particle size can affect its dispersibility and surface area.

3.3. Ultraviolet-Visible Spectroscopy

The UV-Vis spectra of all samples are shown in Figure 4. For BiFeO3, a sharp absorption peak is observed at a wavelength of 360 nm. In Cd-doped samples, the absorption peak is shifted to 395 nm and 420 nm for Bi0.9Cd0.1FeO3 and Bi0.7Cd0.3FeO3, respectively. Figure 4 shows the absorption peak of the prepared samples along with their corresponding Tauc plot. The Tauc plots reveal that the band gaps are direct and the calculated band gap values of BiFeO3, Bi0.9Cd0.1FeO3, and Bi0.7Cd0.3FeO3 are 2.95, 2.76, and 2.51 eV, respectively. It can be inferred from Figure 4 that the band gaps decrease with increasing the doping concentration of Cd in pure BiFeO3, which is evident by the absorption of light in the visible region.

3.4. Fourier Transform Infrared Spectroscopy

To study the oxygen coordination with other metal ions in both A-site and B-site cations in the perovskite structure of all bismuth ferrite samples [42], FT-IR spectroscopy results of the prepared powders are provided in Figure 5. The absorption band is slightly broadened in the range of 3200 cm−1 to 3500 cm−1, which might be associated with N-H stretching [43,44]. On some occasions, overlapping of O-H stretching and N-H stretching may also occur. The literature indicates that such types of bands (occurring at 1100 cm−1 to 1600 cm−1) are due to the presence of nitrates trapped on the surface [45]. Bands occurring at the very start of the peak, i.e., at 600 cm−1, are found to be the result of the vibrations of Fe-O and Bi-O metal-oxygen bonds [46]. Further, an octahedral FeO6 group has the characteristic of Fe-O bond bending and stretching vibrations [47,48,49].

3.5. Raman Spectroscopy

To study the structural changes concerning the different doping concentrations of Cd in BiFeO3, RS was carried out. This technique is quite sensitive to determining the local symmetrical changes in complex perovskites and ferrites [50]. Figure 6 shows RS spectra of BiFeO3, Bi0.9Cd0.1FeO3, and Bi0.7Cd0.3FeO3 obtained at room temperature. Theoretical group theory study of rhombohedral BiFeO3 (belonging to the R3c space group) suggests 13 Raman modes 4A1 and the 9E [51]. However, practically, this statement does not hold, since we obtain a relatively small number of Raman modes (due to weak polarization of ions, mode broadening, and expected degeneracy) in the materials having a large number of atoms in their unit cell [52].
For the pure BiFeO3 sample, the A1-1, A1-2, and A1-3 modes (having relatively higher scattering intensity) are observed at 180, 483, and 616 cm−1, respectively. Mode A1-4 is too weak to be observed. Similarly, all E modes are found to be at 253, 302, 378, 546, 694, 726, and 753 cm−1. As Cd is doped in pure BiFeO3, no remarkable qualitative change in Raman scattering spectra is noticed except that the second and third main A1 modes are shifted towards the right, i.e., towards the higher order of frequency. The new positions are observed at 501 and 620 cm−1, which seem to be nearly equal for both samples. This is attributed to the fact that Cd has a much lower mass as compared to Bi, indicating that the doping element (Cd) is entering at Bi sites of BiFeO3 [53,54].
The XRD patterns of the sample may show both rhombohedral and orthorhombic phases, while the Raman spectra only detect one of them. This could be because the XRD technique is more sensitive to the long-range ordering of atoms in the sample, while Raman spectroscopy is more sensitive to the bonding of the atoms. The presence of both rhombohedral and orthorhombic phases in the XRD patterns could indicate that the sample contains a mixture of both phases, while Raman spectroscopy is not sensitive enough to detect the small amount of the other phase. It is also possible that the Raman spectra only detect one of the phases because the other phase is in a low concentration and not contributing significantly to the Raman signal.

3.6. Vibrating Sample Magnetometry

To study the magnetic behavior of pure and Cd-doped BiFeO3 samples, VSM was performed with an applied field having a maximum value of 15 kOe at room temperature. Figure 7 shows the M-H curves of BiFeO3, Bi0.9Cd0.1FeO3, and Bi0.7Cd0.3FeO3 samples. It is found that at room temperature, the pure BiFeO3 and Cd-doped Bi0.9Cd0.1FeO3 show very weak ferromagnetic behavior. Interestingly, the obtained result for the bulk BiFeO3 in the present study is quite different when compared with bulk BiFeO3 from the literature, where it has been reported as antiferromagnetic at room temperature [55]. However, for the nano-sized BiFeO3, the literature supports the weak ferromagnetic behavior [56]. This difference in magnetic behavior in bulk and nano-sized BiFeO3 may be due to the reduction of particle size that resulted in no exact compensation of two magnetic sublattices with a unit cell [57]. The size of the particles can affect the magnetic properties of the material. In general, smaller particles have a higher surface-to-volume ratio, which can lead to an increase in the number of defects and impurities and can affect the magnetic properties. In the case of BiFeO3, the magnetic properties are known to be sensitive to particle size and shape, and ferromagnetism has been observed in BiFeO3 nanoparticles.
Additionally, the ferromagnetic nature of the M-H loop in BiFeO3 is also related to the presence of Fe ions, which are known to be ferromagnetic. The ferromagnetism in BiFeO3 can also be influenced by other factors, such as the crystal structure, defects, and impurities, as well as the synthesis conditions. Additionally, the result for Bi0.9Cd0.1FeO3 may arise due to the canting of distorted lattice in a previously antiferromagnetically ordered structure, causing an imbalance in antiparallel sublattice magnetization of Fe and Bi ions.
For the doped Bi0.7Cd0.3FeO3 sample, relatively stronger ferromagnetic behavior is observed, which is evident in Figure 6. Further, the measured coercivity (Hc), remnant magnetization (Mr) at maximum (15 kOe) applied external magnetic field are 250 Oe and 0.15 emu/g, respectively. The specific magnetization of Bi0.7Cd0.3FeO3 is calculated to be 0.91 emu/g. The structural transition from rhombohedral to orthorhombic may be the cause of the relatively stronger ferromagnetic behavior of the Bi0.7Cd0.3FeO3 sample. This transition can lead to structural distortion, promoting Fe–Fe atomic interactions and causing an imbalance in magnetic ordering, which otherwise was causing the cancelation of ferromagnetic behavior. This effect of imbalance in magnetic ordering was not dominant enough to suppress a considerable amount of net ferro-magneticity in Bi0.9Cd0.1FeO3 [58,59].

4. Conclusions

Cd-doped BiFeO3 (BixCd1−xFeO3, x = 0, 0.1, 0.3) samples were successfully synthesized via a simple, facile, and efficient chemical co-precipitation method. A transition from rhombohedral to orthorhombic crystal structure was noticed in the doped samples. The XRD spectra reflected the presence of secondary phases of bismuth ferrite (Bi2Fe4O9 and Bi25FeO40) in the doped samples. The SEM micrographs revealed that the particle size lies in the range of micrometers. Moreover, the pure BiFeO3 powder showed a cubic shape of crystals, which changed to a spherical shape with doping of Cd. Raman spectroscopy verified the substitution of Cd at Bi sites. FT-IR results confirmed that Cd substitution resulted in bond stretching (due to the difference in radii of Cd and Bi). Further, VSM confirmed very weak ferromagnetic behavior in pure BiFeO3 and Cd-doped Bi0.9Cd0.1FeO3, while relatively stronger ferromagnetic behavior was observed for Bi0.7Cd0.3FeO3 because the reduction in particle size enhances the magnetic ferromagnetic behavior of the material. For Bi0.7Cd0.3FeO3, the measured coercivity (Hc) remnant magnetization (Mr), at maximum, applied external magnetic field, were 250 Oe and 0.15 emu/g, respectively.

Author Contributions

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

Funding

This study is supported via funding from Prince Sattam bin Abdulaziz University project number ().

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

This study is supported via funding from Prince Sattam bin Abdulaziz University project number (PSAU/2023/R/1444).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Wang, J.B.N.J.; Neaton, J.B.; Zheng, H.; Nagarajan, V.; Ogale, S.B.; Liu, B.; Viehland, D.; Vaithyanathan, V.; Schlom, D.G.; Waghmare, U.V.; et al. Epitaxial BiFeO3 multiferroic thin film heterostructures. Science 2003, 299, 1719–1722. [Google Scholar] [CrossRef] [PubMed]
  2. Catalan, G.; Scott, J.F. Physics and applications of bismuth ferrite. Adv. Mater. 2009, 21, 2463–2485. [Google Scholar] [CrossRef]
  3. Arshad, M.; Hassan, A.; Haider, Q.; Alharbi, F.M.; Alsubaie, N.; Alhushaybari, A.; Burduhos-Nergis, D.-P.; Galal, A.M. Rotating Hybrid Nanofluid Flow with Chemical Reaction and Thermal Radiation between Parallel Plates. Nanomaterials 2022, 12, 4177. [Google Scholar] [CrossRef]
  4. Eerenstein, W.; Mathur, N.D.; Scott, J.F. Multiferroic and magnetoelectric materials. Nature 2006, 442, 759–765. [Google Scholar] [CrossRef] [PubMed]
  5. Dhanalakshmi, B.; Pratap, K.; Rao, B.P.; Rao, P.S. Effects of Mn doping on structural, dielectric, and multiferroic properties of BiFeO3 nanoceramics. J. Alloys Compd. 2016, 676, 193–201. [Google Scholar] [CrossRef]
  6. Dong, S.; Cheng, J.; Li, J.F.; Viehland, D. Enhanced magnetoelectric effects in laminate composites of Terfenol-D/Pb (Zr, Ti) O3 under resonant drive. Appl. Phys. Lett. 2003, 83, 4812–4814. [Google Scholar] [CrossRef]
  7. Fischer, P.; Polomska, M.; Sosnowska, I.; Szymanski, M. Temperature dependence of the crystal and magnetic structures of BiFeO3. J. Phys. C Solid State Phys. 1980, 13, 1931. [Google Scholar] [CrossRef]
  8. Dzik, J.; Feliksik, K.; Pikula, T.; Panek, R.; Rerak, M. Influence of Dy doping on the properties of BiFeO3. Arch. Metall. Mater. 2018, 63, 1351–1355. [Google Scholar]
  9. Ishaq, B.; Murtaza, G.; Sharif, S.; Khan, M.A.; Akhtar, N.; Will, I.; Saleem, M.; Ramay, S.M. Investigating the effect of Cd-Mn co-doped nano-sized BiFeO3 on its physical properties. Results Phys. 2016, 6, 675–682. [Google Scholar] [CrossRef] [Green Version]
  10. Karpinsky, D.; Silibin, M.; Trukhanov, A.; Zhaludkevich, A.; Maniecki, T.; Maniukiewicz, W.; Sikolenko, V.; Paixão, J.; Khomchenko, V. A correlation between crystal structure and magnetic properties in co-doped BiFeO3 ceramics. J. Phys. Chem. Solids 2019, 126, 164–169. [Google Scholar] [CrossRef]
  11. Arafat, S.S. Structural transition and magnetic properties of high Cr-doped BiFeO3 ceramic. Cerâmica 2020, 66, 114–118. [Google Scholar] [CrossRef]
  12. Zhuang, J.; Zhao, J.; Su, L.W.; Wu, H.; Bokov, A.A.; Ren, W.; Ye, Z.G. Structure and local polar domains of Dy-modified BiFeO3–PbTiO3 multiferroic solid solutions. J. Mater. Chem. C 2015, 3, 12450–12456. [Google Scholar] [CrossRef]
  13. Arshad, M.; Hussain, A.; Hassan, A.; Karamti, H.; Wroblewski, P.; Khan, I.; Andualem, M.; Galal, A.M. Scrutinization of Slip Due to Lateral Velocity on the Dynamics of Engine Oil Conveying Cupric and Alumina Nanoparticles Subject to Coriolis Force. Math. Probl. Eng. 2022, 2022, 2526951. [Google Scholar] [CrossRef]
  14. Teague, J.R.; Gerson, R.; James, W.J. Dielectric hysteresis in single crystal BiFeO3. Solid State Commun. 1970, 8, 1073–1074. [Google Scholar] [CrossRef]
  15. Carvalho, T.T.; Tavares, P.B. Synthesis and thermodynamic stability of multiferroic BiFeO3. Mater. Lett. 2008, 62, 3984–3986. [Google Scholar] [CrossRef]
  16. Zhou, W.; Deng, H.; Yu, L.; Yang, P.; Chu, J. Band-gap narrowing and magnetic behavior of Ni-doped Ba (Ti0.875Ce0.125) O3 thin films. J. Phys. D Appl. Phys. 2015, 48, 455308. [Google Scholar] [CrossRef]
  17. Arshad, M.; Karamti, H.; Awrejcewicz, J.; Grzelczyk, D.; Galal, A.M. Thermal Transmission Comparison of Nanofluids over Stretching Surface under the Influence of Magnetic Field. Micromachines 2022, 13, 1296. [Google Scholar] [CrossRef]
  18. Yuan, G.L.; Or, S.W.; Liu, J.M.; Liu, Z.G. Structural transformation and ferroelectromagnetic behavior in single-phase Bi1−xNdxFeO3 multiferroic ceramics. Appl. Phys. Lett. 2006, 89, 052905. [Google Scholar] [CrossRef] [Green Version]
  19. Yuan, G.L.; Or, S.W.; Chan, H.L.W. Structural transformation and ferroelectric–paraelectric phase transition in Bi1−xLaxFeO3 (x = 0–0.25) multiferroic ceramics. J. Phys. D Appl. Phys. 2007, 40, 1196. [Google Scholar] [CrossRef]
  20. Khomchenko, V.A.; Kiselev, D.A.; Selezneva, E.K.; Vieira, J.M.; Lopes, A.M.L.; Pogorelov, Y.G.; Araujo, J.P.; Kholkin, A.L. Weak ferromagnetism in diamagnetically-doped Bi1−xAxFeO3 (A = Ca, Sr, Pb, Ba) multiferroics. Mater. Lett. 2008, 62, 1927–1929. [Google Scholar] [CrossRef]
  21. Reddy, V.A.; Pathak, N.P.; Nath, R. Domain switching in spray pyrolysis-deposited nano-crystalline BiFeO3 films. Phys. Scr. 2012, 86, 065701. [Google Scholar] [CrossRef]
  22. Park, T.J.; Papaefthymiou, G.C.; Viescas, A.J.; Lee, Y.; Zhou, H.; Wong, S.S. Composition-dependent magnetic properties of BiFeO3-BaTiO3 solid solution nanostructures. Phys. Rev. B 2010, 82, 024431. [Google Scholar] [CrossRef]
  23. Chaudhuri, A.; Mandal, K. Enhancement of Ferromagnetic and Dielectric properties of nanostructured Barium doped Bismuth Ferrite fabricated by facile hydrothermal route. In AIP Conference Proceedings; AIP Publishing LLC: Melville, NY, USA, 2015; Volume 1665, p. 050022. [Google Scholar]
  24. Tagantsev, A.K.; Stolichnov, I.; Colla, E.L.; Setter, N. Polarization fatigue in ferroelectric films: Basic experimental findings, phenomenological scenarios, and microscopic features. J. Appl. Phys. 2001, 90, 1387–1402. [Google Scholar] [CrossRef]
  25. Kumar, M.; Yadav, K.L. Study of room temperature magnetoelectric coupling in Ti substituted bismuth ferrite system. J. Appl. Phys. 2006, 100, 074111. [Google Scholar] [CrossRef]
  26. Qi, X.; Dho, J.; Tomov, R.; Blamire, M.G.; MacManus-Driscoll, J.L. Greatly reduced leakage current and conduction mechanism in aliovalent-ion-doped BiFeO3. Appl. Phys. Lett. 2005, 86, 062903. [Google Scholar] [CrossRef]
  27. Tabares-Munoz, C.; Rivera, J.P.; Bezinges, A.; Monnier, A.; Schmid, H. Measurement of the quadratic magnetoelectric effect on single crystalline BiFeO3. Jpn. J. Appl. Phys. 1985, 24, 1051. [Google Scholar] [CrossRef] [Green Version]
  28. Przenioslo, R.; Regulski, M.; Sosnowska, I. Modulation in multiferroic BiFeO3: Cycloidal, elliptical, or SDW? J. Phys. Soc. Jpn. 2006, 75, 084718. [Google Scholar] [CrossRef]
  29. Sosnowska, I.; Neumaier, T.P.; Steichele, E. Spiral magnetic ordering in bismuth ferrite. J. Phys. C Solid State Phys. 1982, 15, 4835. [Google Scholar] [CrossRef]
  30. Maurya, D.; Thota, H.; Nalwa, K.S.; Garg, A. BiFeO3 ceramics synthesized by mechanical activation assisted versus conventional solid-state-reaction process: A comparative study. J. Alloys Compd. 2009, 477, 780–784. [Google Scholar] [CrossRef]
  31. Ghosh, S.; Dasgupta, S.; Sen, A.; Maiti, H.S. Low temperature synthesis of bismuth ferrite nanoparticles by a ferrioxalate precursor method. Mater. Res. Bull. 2005, 40, 2073–2079. [Google Scholar] [CrossRef]
  32. Cho, C.M.; Noh, J.H.; Cho, I.S.; An, J.S.; Hong, K.S.; Kim, J.Y. Low-temperature hydrothermal synthesis of pure BiFeO3 nanopowders using triethanolamine and their applications as visible-light photocatalysts. J. Am. Ceram. Soc. 2008, 91, 3753–3755. [Google Scholar] [CrossRef]
  33. Park, T.J.; Papaefthymiou, G.C.; Viescas, A.J.; Moodenbaugh, A.R.; Wong, S.S. Size-dependent magnetic properties of single-crystalline multiferroic BiFeO3 nanoparticles. Nano Lett. 2007, 7, 766–772. [Google Scholar] [CrossRef] [PubMed]
  34. Yu, B.; Li, M.; Liu, J.; Guo, D.; Pei, L.; Zhao, X. Effects of ion doping at different sites on electrical properties of multiferroic BiFeO3 ceramics. J. Phys. D Appl. Phys. 2008, 41, 065003. [Google Scholar] [CrossRef]
  35. Hassan, A.; Hussain, A.; Fernandez-Gamiz, U.; Arshad, M.; Karamti, H.; Awrejcewicz, J.; Alharbi, F.M.; Elfasakhany, A.; Galal, A.M. Computational investigation of magneto-hydrodynamic flow of newtonian fluid behavior over obstacles placed in rectangular cavity. Alex. Eng. J. 2023, 65, 163–188. [Google Scholar] [CrossRef]
  36. Kim, J.K.; Kim, S.S.; Kim, W.J. Sol–gel synthesis and properties of multiferroic BiFeO3. Mater. Lett. 2005, 59, 4006–4009. [Google Scholar] [CrossRef]
  37. Han, J.; Huang, Y.-H.; Wu, X.-J.; Wu, C.-L.; Wei, W.; Peng, B.; Huang, W.; Goodenough, J.B. Tunable synthesis of bismuth ferrites with various morphologies. Adv. Mater. 2006, 18, 2145–2148. [Google Scholar] [CrossRef]
  38. Arshad, M.; Hussain, A.; Elfasakhany, A.; Gouadria, S.; Awrejcewicz, J.; Pawłowski, W.; Elkotb, M.A.; Alharbi, F.M. Magneto-hydrodynamic flow above exponentially stretchable surface with chemical reaction. Symmetry 2022, 14, 1688. [Google Scholar] [CrossRef]
  39. Chen, C.; Cheng, J.; Yu, S.; Che, L.; Meng, Z. Hydrothermal synthesis of perovskite bismuth ferrite crystallites. J. Cryst. Growth 2006, 291, 135–139. [Google Scholar] [CrossRef]
  40. Özdemir, D.K. Temperature Susceptibility and Rheological Aging Characteristics of the Bitumen Having Different Penetration Grades. Black Sea J. Eng. Sci. 2021, 4, 209–213. [Google Scholar]
  41. Cheng, Z.; Li, A.H.; Wang, X.L.; Dou, S.X.; Ozawa, K.; Kimura, H.; Zhang, S.; Shrout, T.R. Structure, ferroelectric properties, and magnetic properties of the La-doped bismuth ferrite. J. Appl. Phys. 2008, 103, 07E507. [Google Scholar] [CrossRef]
  42. Pradeep, A.; Chandrasekaran, G. FTIR study of Ni, Cu and Zn substituted nano-particles of MgFe2O4. Mater. Lett. 2006, 60, 371–374. [Google Scholar] [CrossRef]
  43. Muneeswaran, M.; Jegatheesan, P.; Giridharan, N.V. Synthesis of nanosized BiFeO3 powders by co-precipitation method. J. Exp. Nanosci. 2013, 8, 341–346. [Google Scholar] [CrossRef]
  44. Parikh, A.; Madamwar, D. Partial characterization of extracellular polysaccharides from cyanobacteria. Bioresour. Technol. 2006, 97, 1822–1827. [Google Scholar] [CrossRef] [PubMed]
  45. Ghosh, S.; Dasgupta, S.; Sen, A.; Sekhar Maiti, H. Low-temperature synthesis of nanosized bismuth ferrite by soft chemical route. J. Am. Ceram. Soc. 2005, 88, 1349–1352. [Google Scholar] [CrossRef]
  46. Nakanishi, K.; Solomon, P.H. Infrared Absorption Spectroscopy; Holden-day: Eads, TN, USA, 1977. [Google Scholar]
  47. Anthony Raj, C.; Muneeswaran, M.; Jegatheesan, P.; Giridharan, N.V.; Sivakumar, V.; Senguttuvan, G. Effect of annealing time in the low-temperature growth of BFO thin films spin coated on glass substrates. J. Mater. Sci. Mater. Electron. 2013, 24, 4148–4154. [Google Scholar] [CrossRef]
  48. Biasotto, G.; Simões, A.Z.; Foschini, C.R.; Zaghete, M.A.; Varela, J.A.; Longo, E. Microwave-hydrothermal synthesis of perovskite bismuth ferrite nanoparticles. Mater. Res. Bull. 2011, 46, 2543–2547. [Google Scholar] [CrossRef]
  49. Yotburut, B.; Yamwong, T.; Thongbai, P.; Maensiri, S. Synthesis and characterization of coprecipitation-prepared La-doped BiFeO3 nanopowders and their bulk dielectric properties. Jpn. J. Appl. Phys. 2014, 53, 06JG13. [Google Scholar] [CrossRef]
  50. Arshad, M.; Hassan, A. A numerical study on the hybrid nanofluid flow between a permeable rotating system. Eur. Phys. J. Plus 2022, 137, 1126. [Google Scholar] [CrossRef]
  51. Haumont, R.; Kreisel, J.; Bouvier, P. Raman scattering of the model multiferroic oxide BiFeO3: Effect of temperature, pressure and stress. Phase Transit. 2006, 79, 1043–1064. [Google Scholar] [CrossRef]
  52. White, W.B. Structure of particles and the structure of crystals: Information from vibrational spectroscopy. J. Ceram. Process. Res. 2005, 6, 1–9. [Google Scholar]
  53. Hassan, A.; Hussain, A.; Arshad, M.; Awrejcewicz, J.; Pawlowski, W.; Alharbi, F.M.; Karamti, H. Heat and mass transport analysis of MHD rotating hybrid nanofluids conveying silver and molybdenum di-sulfide nano-particles under effect of linear and non-linear radiation. Energies 2022, 15, 6269. [Google Scholar] [CrossRef]
  54. Bozgeyik, M.S.; Katiyar, R.K.; Katiyar, R.S. Improved magnetic properties of bismuth ferrite ceramics by La and Gd co-substitution. J. Electroceramics 2018, 40, 247–256. [Google Scholar] [CrossRef]
  55. Zhang, S.T.; Lu, M.H.; Wu, D.; Chen, Y.F.; Ming, N.B. Larger polarization and weak ferromagnetism in quenched BiFeO3 ceramics with a distorted rhombohedral crystal structure. Appl. Phys. Lett. 2005, 87, 262907. [Google Scholar] [CrossRef]
  56. Sinha, A.; Bhushan, B.; Jagannath; Sharma, R.; Sen, S.; Mandal, B.; Meena, S.; Bhatt, P.; Prajapat, C.; Priyam, A.; et al. Enhanced dielectric, magnetic and optical properties of Cr-doped BiFeO3 multiferroic nanoparticles synthesized by sol-gel route. Results Phys. 2019, 13, 102299. [Google Scholar] [CrossRef]
  57. Dutta, D.P.; Mandal, B.P.; Naik, R.; Lawes, G.; Tyagi, A.K. Magnetic, ferroelectric, and magnetocapacitive properties of sonochemically synthesized Sc-doped BiFeO3 nanoparticles. J. Phys. Chem. C 2013, 117, 2382–2389. [Google Scholar] [CrossRef]
  58. Moriya, T. Anisotropic superexchange interaction and weak ferromagnetism. Phys. Rev. 1960, 120, 91. [Google Scholar] [CrossRef] [Green Version]
  59. Pradhan, S.K.; Roul, B.K.; Sahu, D.R. Enhancement of ferromagnetism and multiferroicity in Ho doped Fe rich BiFeO3. Solid State Commun. 2012, 152, 1176–1180. [Google Scholar] [CrossRef]
Figure 1. (a) XRD patterns of BiFeO3, Bi0.7Cd0.3FeO3, and Bi0.9Cd0.1FeO3 powders calcined at 550 °C, (b) zoomed image of the characteristic peak near 32°. The shift towards a higher angle confirms the successful substitution of Cd with Bi in the BiFeO3 crystal.
Figure 1. (a) XRD patterns of BiFeO3, Bi0.7Cd0.3FeO3, and Bi0.9Cd0.1FeO3 powders calcined at 550 °C, (b) zoomed image of the characteristic peak near 32°. The shift towards a higher angle confirms the successful substitution of Cd with Bi in the BiFeO3 crystal.
Inorganics 11 00134 g001
Figure 2. SEM images of (a) BiFeO3, (b) Bi0.9Cd0.1FeO3, (c) Bi0.7Cd0.3FeO3, and (d) particle size graph along with Gaussian fitting to evaluate average particle size for a pure sample.
Figure 2. SEM images of (a) BiFeO3, (b) Bi0.9Cd0.1FeO3, (c) Bi0.7Cd0.3FeO3, and (d) particle size graph along with Gaussian fitting to evaluate average particle size for a pure sample.
Inorganics 11 00134 g002
Figure 3. EDX spectrum of Bi0.7Cd0.3FeO3 for elemental compositional analysis confirming the successful doping of Cd into BiFeO3.
Figure 3. EDX spectrum of Bi0.7Cd0.3FeO3 for elemental compositional analysis confirming the successful doping of Cd into BiFeO3.
Inorganics 11 00134 g003
Figure 4. (a) UV visible absorption plot for BiFeO3, Bi0.9Cd0.1O3 and Bi0.7Cd0.3FeO3. The arrow in (a) shows the increasing pattern of increasing wavelength. (b) there corresponding Tauc plot for evaluating the band gap of synthesized samples.
Figure 4. (a) UV visible absorption plot for BiFeO3, Bi0.9Cd0.1O3 and Bi0.7Cd0.3FeO3. The arrow in (a) shows the increasing pattern of increasing wavelength. (b) there corresponding Tauc plot for evaluating the band gap of synthesized samples.
Inorganics 11 00134 g004
Figure 5. FT-IR spectra of BiFeO3, Bi0.9Cd0.1FeO3, and Bi0.7Cd0.3FeO3.
Figure 5. FT-IR spectra of BiFeO3, Bi0.9Cd0.1FeO3, and Bi0.7Cd0.3FeO3.
Inorganics 11 00134 g005
Figure 6. Raman scattering spectra analysis for BiFeO3, Bi0.9Cd0.1FeO3, and Bi0.7Cd0.3FeO3 at room temperature.
Figure 6. Raman scattering spectra analysis for BiFeO3, Bi0.9Cd0.1FeO3, and Bi0.7Cd0.3FeO3 at room temperature.
Inorganics 11 00134 g006
Figure 7. M-H curves of BiFeO3, Bi0.9Cd0.1FeO3, and Bi0.7Cd0.3FeO3 samples at room temperature.
Figure 7. M-H curves of BiFeO3, Bi0.9Cd0.1FeO3, and Bi0.7Cd0.3FeO3 samples at room temperature.
Inorganics 11 00134 g007
Table 1. Crystallite size calculated by the Debye-Scherrer formula.
Table 1. Crystallite size calculated by the Debye-Scherrer formula.
Material 2 θ hklCrystallite Size
BiFeO332.1711040.62 nm
Bi0.9Cd0.1FeO332.0611034.97 nm
Bi0.7Cd0.3FeO332.0111029.14 nm
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Mehmood, S.Z.; Arshad, M.; Alharbi, F.M.; Eldin, S.M.; Galal, A.M. Multiferroics Made via Chemical Co-Precipitation That Is Synthesized and Characterized as Bi(1−x)CdxFeO3. Inorganics 2023, 11, 134. https://doi.org/10.3390/inorganics11030134

AMA Style

Mehmood SZ, Arshad M, Alharbi FM, Eldin SM, Galal AM. Multiferroics Made via Chemical Co-Precipitation That Is Synthesized and Characterized as Bi(1−x)CdxFeO3. Inorganics. 2023; 11(3):134. https://doi.org/10.3390/inorganics11030134

Chicago/Turabian Style

Mehmood, Syed Zain, Mubashar Arshad, Fahad M. Alharbi, Sayed M. Eldin, and Ahmed M. Galal. 2023. "Multiferroics Made via Chemical Co-Precipitation That Is Synthesized and Characterized as Bi(1−x)CdxFeO3" Inorganics 11, no. 3: 134. https://doi.org/10.3390/inorganics11030134

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