Abstract
Mn1-xZnxFe2O4 ferrofluids were produced from natural sand for magnetic sensors and radar absorbing materials. The X-ray diffraction data showed that the Zn partially substituted the Mn and Fe ions to construct a spinel structure. The increasing Zn composition decreased the lattice parameters of the structure. The transmission electron microscopy images showed that the filler Mn1-xZnxFe2O4 nanoparticles tended to agglomerate in three dimensions. Lognormal and mass fractal models were used to fit the small-angle X-ray scattering data of the ferrofluids demonstrated that the ferrofluids formed chain-like structures with a fractal dimension of 1.12–1.67 that was constructed from primary particles with sizes of 3.6–4.1 nm. The filler, surfactant, and carrier liquid of the ferrofluids were confirmed by the functional groups of the metal oxides, tetramethylammonium hydroxide, and H2O, respectively. The secondary particles contributed to the saturation magnetization of the Mn1-xZnxFe2O4 ferrofluids. The Mn1-xZnxFe2O4 ferrofluids demonstrated excellent performance as magnetic sensors with high stability, especially compared with MnFe2O4 ferrofluids. Furthermore, the ferrofluids exhibited excellent radar absorbing materials. The Mn1-xZnxFe2O4 ferrofluids prepared in this work may serve as a future platform for advancing magnetic sensors and radar absorbing materials.
Keywords: Materials science, Nanotechnology, Electromagnetism, Natural sand, Mn1-xZnxFe2O4, Ferrofluid, Magnetic sensor, Radar absorbing material
Materials science; Nanotechnology; Electromagnetism; Natural sand; Mn1-xZnxFe2O4; Ferrofluid; Magnetic sensor; Radar absorbing material
1. Introduction
In the last few years, fabrication methods for producing spinel MnZn ferrite (Mn1-xZnxFe2O4) nanoparticles with high quality structural and magnetic properties have been developed [1]. The development has also intended to improve the specific application performance of Mn1-xZnxFe2O4 nanoparticles, especially for sensors [2] and radar or microwave absorbing materials [3]. In general, the application of Mn1-xZnxFe2O4 nanoparticles has been triggered by their fascinating characteristics, such as excellent magnetic loss, good corrosion resistance, moderate saturation magnetization, and low cost [4]. The Mn1-xZnxFe2O4 nanoparticles have also been proposed for practical applications, including sensors, owing to their high magnetization sensitivity to temperature [5]. Furthermore, the Mn1-xZnxFe2O4 nanoparticles have high reflection loss and broadband absorption, providing a significant advantage as a radar absorbing material [6, 7]. However, until recently, there were at least two main challenges with producing Mn1-xZnxFe2O4 nanoparticles for magnetic sensors and radar absorbing materials with excellent performance. The first issue is producing Mn1-xZnxFe2O4 nanoparticles with a specific hierarchical structure, such as chain-like structure, with high stability and sensitivity under a small external magnetic field. The second issue is producing Mn1-xZnxFe2O4 nanoparticles in mass production using simple, inexpensive, effective, and ecofriendly synthesis.
To overcome the first issue, it is essential to produce small Mn1-xZnxFe2O4 magnetic nanoparticles in a monodisperse system by using a suitable surfactant agent to prevent agglomeration. In this study, we employed a single surfactant technique to coat magnetic particles with tetramethylammonium hydroxide (TMAH). The use of a surfactant is crucial to produce monodisperse magnetic nanoparticles with superparamagnetic characteristics [8, 9]. Fabricating magnetic nanoparticles in a ferrofluid system is an ideal experiment to address the structure issue. Practically, ferrofluids have specific advantages compared with bulk, film, and even in nanopowder forms. One such benefit is that ferrofluids can maintain magnetic properties in their liquid form and are very sensitive to external magnetic fields [10, 11], which is required for magnetic sensor applications. Interestingly, the interaction between magnetic nanoparticles and external magnetic fields is contributed to by the magnetic moment and liquid flexibility, making magnetic particles in ferrofluids easier to align under an external magnetic field [12]. In addition, ferrofluids can also maintain the properties of a fluid even in the presence of high magnetic fields, and magnetic particles as fillers do not separate from the carrier fluid to provide additional advantages for sensors and radar absorber applications.
To address the second issue, we proposed a fabrication method that uses an abundant natural material as the primary precursor to produce Mn1-xZnxFe2O4 ferrofluids. To the best our knowledge, the fabrication of Mn1-xZnxFe2O4 nanoparticles for ferrofluids tends to use relatively expensive commercial primary precursors that are less economical for mass production [5, 13, 14]. Therefore, in this study, we developed a fabrication method of Mn1-xZnxFe2O4 ferrofluids using abundant natural iron sand as the main precursor that is effective, efficient, inexpensive, and ecofriendly. Natural iron sand can be employed effectively to minimize the fabrication cost for the synthesis of high quality magnetic nanoparticles [15]. Moreover, to further reduce the fabrication cost, water was used as the polar carrier liquid in the fabrication of Mn1-xZnxFe2O4 ferrofluids with a chain-like structure. The prepared ferrofluids were investigated by their hierarchical nanostructures, molecular structures, and magnetic properties. Finally, the performance of the Mn1-xZnxFe2O4 ferrofluids as magnetic sensors and radar absorbing materials were also explored.
2. Experimental methods
In this study, natural iron sand was used as the primary source of Fe by the formation of ferric chloride and ferrous chloride. MnCl2.H2O was used as the primary source of Mn and ZnCl2 was employed as the main source of Zn. The other materials were TMAH used as the surfactant, H2O used as the carrier liquid, HCl (38%) used as the solvent, and NH4OH (25%) used as the precipitating agent. Dried natural iron sand was separated from impurities using a permanent magnet to produce Fe3O4 powder with a purity of 99.5% [15]. Twenty grams of Fe3O4 powder was reacted with 58 mL of HCl using a magnetic stirrer for 30 min at room temperature to obtain ferric chloride and ferrous chloride. The ferric chloride and ferrous chloride were reacted with MnCl2.H2O and ZnCl2 at specific compositions under stirring for 30 min, followed by a sonication process in a room temperature ultrasonic bath at a frequency of 40 kHz for 10 min. The next process was titration with 27 mL of NH4OH for 30 min to obtain a black precipitate consisting of Mn1-xZnxFe2O4 nanoparticles. The precipitate was then filtered and washed using distilled water until the pH was 7. The x value of Mn1-xZnxFe2O4 representing Zn and Mn ions was varied, resulting in compositions of Fe3O4, MnFe2O4, Mn0.75Zn0.25Fe2O4, Mn0.5Zn0.5Fe2O4, Mn0.25Zn0.75Fe2O4, and ZnFe2O4. Subsequently, 1.2 g of Mn1-xZnxFe2O4 nanoparticles were reacted with 1.2 mL of TMAH under rough stirring for 15 min to obtain a homogeneous suspension. Five milliliters of H2O was then added to the suspension as a dispersant media and stirred for 15 min to obtain the Mn1-xZnxFe2O4 ferrofluids.
X-ray diffractometry (XRD; X'Pert Pro, PANalytical) was used to determine the phase, particle size, and lattice parameters of the samples. The morphology of the samples was identified using transmission electron microscopy (TEM; JEOL 1400). Characterization of the functional group of the Mn1-xZnxFe2O4 ferrofluids was performed using Fourier-transform infrared spectroscopy (FTIR; Shimadzu). The magnetic properties were identified using vibrating magnetometer samples (VSM; Oxford). The hierarchical nanostructure of the Mn1-xZnxFe2O4 ferrofluids was investigated using a synchrotron-based small-angle X-ray scattering (SAXS). Finally, the potential application of the samples was examined by characterization of the magnetic sensor performance in a magnetic field range of 0–200 mT. The relaxation particle time was also investigated at room temperature. The radar absorption performance of the samples was evaluated using a vector network analyzer (VNA) at a frequency range of 8–12 GHz, which was the working area of the radar.
3. Results and discussion
The X-ray diffraction patterns of the Mn1-xZnxFe2O4 nanoparticles as fillers are depicted in Figure 1. The green lines represent the fitting model using AMCSD No. 0007423 and the experimental data are represented by the black circles. The Rietveld refinement method was employed using the Rietica program to investigate the crystal structure and particle size of the Mn1-xZnxFe2O4 nanoparticles [16]. Visually, the most intense diffraction peak was Miller index (311), which tended to shift to a lower 2θ value originating from the Zn ion composition. Quantitatively, the highest diffraction peak of the Fe3O4 nanoparticles was detected at 35.66° 2θ, while the highest diffraction peak of the MnFe2O4 nanoparticles shifted to a lower 2θ value at 35.48°. The shifting of the highest diffraction peak originated with the presence of Mn substitution into the Fe3O4 nanoparticles [17, 18] owing to the difference in Fe2+, Fe3+, and Mn2+ ionic radii of 0.77, 0.65, and 0.82 Å, respectively [19, 20]. Therefore, the lattice parameters increased, leading to a shift in the diffraction peak to a lower position. Furthermore, the higher Zn ion composition in the Mn0.75Zn0.25Fe2O4, Mn0.5Zn0.5Fe2O4, Mn0.25Zn0.75Fe2O4, and ZnFe2O4 samples tended to increase the highest diffraction peak position to 35.59°, 35.62°, 35.64°, and 35.71°, respectively. The increasing 2θ values with decreasing lattice parameters originated from the increasing Zn ions with an ionic radii of 0.74 Å, which is smaller than that of the Mn ions [21]. The lattice parameters, crystal volume, and particle size of the samples are presented in Table 1. Interestingly, the trend of the lattice parameters and crystal volumes tended to be similar to the particle size pattern of the samples. For the maximum Mn ion composition, the lattice parameters and crystal volume increased significantly and decreased gradually with the increasing Zn ion composition. The crystal structure of the samples was similar to a previous study [22]. In this work, all samples had a spinel cubic structure [23], where the Mn2+, Zn2+, Fe2+, and Fe3+ were randomly in octahedral and tetrahedral sites in the spinel system with lattice parameters of a = b = c.
Figure 1.
X-ray diffraction patterns of the Mn1-xZnxFe2O4 nanoparticles.
Table 1.
Lattice parameters and particle size of the Mn1-xZnxFe2O4 nanoparticles.
| Sample | Lattice parameters a = b = c (Å) | Crystal volume V = a3 (Å3) |
Particle size (nm) |
|---|---|---|---|
| Fe3O4 | 8.351 ± 0.005 | 582.4 ± 0.6 | 9.5 ± 0.3 |
| MnFe2O4 | 8.380 ± 0.003 | 588.5 ± 0.6 | 11.5 ± 0.3 |
| Mn0.75Zn0.25Fe2O4 | 8.377 ± 0.003 | 587.9 ± 0.5 | 10.4 ± 0.2 |
| Mn0.5Zn0.5Fe2O4 | 8.372 ± 0.002 | 586.8 ± 0.7 | 9.8 ± 0.2 |
| Mn0.25Zn0.75Fe2O4 | 8.365 ± 0.004 | 585.3 ± 0.5 | 9.1 ± 0.2 |
| ZnFe2O4 | 8.349 ± 0.003 | 581.9 ± 0.7 | 10.1 ± 0.3 |
With regard to the change in the crystal parameters with Zn and Mn doping, the cationic distribution of the spinel structure variate depends on the degree of inversion (i). Based on the stoichiometry, the degree of inversion can be defined as the fraction of A-sites that are filled by cations and initially attributed to the B-sites. The Mn1-xZnxFe2O4 nanoparticles (0 < x < 1) are defined by the cationic distribution shown in Eq. (1), according to the degree of inversion proposed by Klencsár et al. [24]. The samples with x = 0.25, 0.50, and 0.75 have a degree of inversion of 0.59, 0.50, and 0.41, respectively. Furthermore, the degree of inversion of samples x = 0 and 1 are 0.37 and 0.68, respectively. Using the calculation model by Klencsár et al. [24] and Liu et al. [25], including applying the degree of inversion, the cationic distribution of samples x = 0 and 1 can be expressed in Eqs. (2) and (3). Moreover, the cationic distribution of Fe3O4 is presented in Eq. (4).
| (1) |
| (2) |
| (3) |
| (4) |
Figure 2 presents the TEM characterization illustrating the morphology of the Mn1-xZnxFe2O4 nanoparticles. The characterization focused on three samples, Fe3O4, Mn0.5Zn0.5Fe2O4, and ZnFe2O4 nanoparticles. As shown in Figure 2, the fabricated samples were nanometer in size. Based on a physics perspective, small magnetic nanoparticles have high surface area, which leads to an increase in van der Waals force [26]. This force causes the Mn1-xZnxFe2O4 nanoparticles to agglomerate. The average particle size of the Fe3O4, Mn0.5Zn0.5Fe2O4, and ZnFe2O4 nanoparticles were 11.1, 13.0, and 10.3 nm, respectively. Although the TEM characterization showed that the nanoparticles were successfully fabricated in a nanometer size, the particle size and hierarchical structure of the Mn1-xZnxFe2O4 ferrofluids could not be characterized. Therefore, an in situ investigation was conducted using a synchrotron-based SAXS to investigate the real particle size and hierarchical structure of the Mn1-xZnxFe2O4 ferrofluids. The SAXS profiles of the Mn1-xZnxFe2O4 ferrofluids and its fitting analysis are presented in Figure 3. The SAXS profiles were fitted using mathematical models as well as the form factor P(q,R), structure factor S(q,R), intensity I(q), and distribution of particle size f(R), as shown in Eqs. (5), (6), (7), and (8) [18, 27]:
| (5) |
| (6) |
| (7) |
| (8) |
where q is the scattering vector, R is the particle size, σ is the polydispersity index, R1 is the primary particles, R2 is the secondary particles, Ro is the mean particle radii, N is the normalization factor, ξ is the cutoff length from fractal correlation, and D is the fractal dimension [20]. In general, the ferrofluids with a single surfactant stabilization have particle size distribution along with a polydisperse character and chain-like structure [28]. According to the theoretical approach, the particle size of the ferrofluids can be suitably fit with a lognormal model in constructing the fractal structures [29]. Specifically, the fitting analysis results of the SAXS data for the Mn1-xZnxFe2O4 ferrofluids are presented in Table 2. The primary particles of the Mn1-xZnxFe2O4 ferrofluids were spherical, with a diameter of 3.6–4.1 nm that formed secondary cluster particles, which is similar to the results of a previous work [30]. The primary particle building blocks constructed a fractal structure with a chain-like structure for all samples with a fractal dimension from 1.12 to 1.67. The primary particles built the secondary particles, constructing a fractal structure in one dimension as a chain-like-structure [31]. An illustration of the hierarchical structure of the MnxZn1-xFe2O4 ferrofluids is shown in Figure 4. Theoretically, the chain-like structure in the ferrofluids is built even in the presence of an external magnetic field.
Figure 2.
TEM images of the Mn1-xZnxFe2O4 nanoparticles.
Figure 3.
Synchrotron SAXS profiles of the (a) Fe3O4, (b) MnFe2O4, (c) Mn0.75Zn0.25Fe2O4, (d) Mn0.5Zn0.5Fe2O4, (e) Mn0.25Zn0.75Fe2O4, and (f) ZnFe2O4 ferrofluids. The black circles and green solid lines represent the experimental data and fitting model, respectively.
Table 2.
Results of the synchrotron SAXS fitting of the Mn1-xZnxFe2O4 ferrofluids.
| Sample | Primary particle (nm) | Secondary particle (nm) | Fractal aggregate (nm) | Fractal dimension |
|---|---|---|---|---|
| Fe3O4 | 3.6 ± 0.1 | 11.1 ± 0.1 | 80.0 ± 4.5 | 1.67 ± 0.17 |
| MnFe2O4 | 3.6 ± 0.2 | 10.5 ± 0.1 | 27.9 ± 1.3 | 1.49 ± 0.13 |
| Mn0.75Zn0.25Fe2O4 | 4.1 ± 0.2 | 13.9 ± 0.2 | 68.7 ± 3.6 | 1.36 ± 0.11 |
| Mn0.5Zn0.5Fe2O4 | 3.6 ± 0.1 | 16.8 ± 0.3 | 65.7 ± 3.1 | 1.13 ± 0.05 |
| Mn0.25Zn0.75Fe2O4 | 4.0 ± 0.2 | 13.0 ± 0.2 | 32.8 ± 1.0 | 1.40 ± 0.10 |
| ZnFe2O4 | 3.6 ± 0.1 | 12.8 ± 0.1 | 68.9 ± 3.4 | 1.12 ± 0.05 |
Figure 4.
Illustration of the hierarchical structure of the Mn1-xZnxFe2O4 ferrofluids.
The functional groups of the Mn1-xZnxFe2O4 ferrofluids are presented in Figure 5. The stretching band of Fe3+–O was observed at the wavenumber of 433 cm−1 and the band of Fe2+–O was in the range of 573–634 cm−1 [32]. Fe3+–O and Fe2+–O were located at the octahedral and tetrahedral sites. The functional groups of the metal–oxygen (M–O) group originated from the magnetic particles as fillers to form ferrofluids. The competition of the Mn and Zn ions at the octahedral and tetrahedral sites in the spinel system tended to change the lattice parameters of the Mn1-xZnxFe2O4 ferrofluids. Furthermore, the presence of TMAH was evaluated by the functional group of the C–N, C–H, and O–H observed at the wavenumbers of 1395, 1508, and 3377 cm−1, respectively [33]. The TMAH acted as a surfactant to cover the magnetic nanoparticles. Finally, the carrier liquid was detected by the functional group of O–H identified at the same position with the stretching band of the TMAH at the wavenumber of 3379 cm−1 [34], broadening the peak for the O–H stretching band. Therefore, all ferrofluid components were perfectly observed originating from the filler, surfactant, and carrier liquid.
Figure 5.
Infrared spectrum of the Mn1-xZnxFe2O4 ferrofluids.
The magnetic properties of the Mn1-xZnxFe2O4 ferrofluids were evaluated from the magnetization curves, as shown in Figure 6. The magnetization patterns of the ferrofluids tended to be different, specifically related to saturation magnetization. Theoretically, the smaller the particle size, the greater the magnetic moment on the surface. Under this condition, the magnetic moment on the surface contributes significantly to the magnetic properties of the materials. Kaur et al. explained that the saturation magnetization of magnetic nanoparticles in the spinel system depends on their magnetic moment and exchange interaction [22]. In the spinel structure, the magnetic moment is influenced by the atomic spin interaction mechanism at the A-site and B-site. Therefore, the net magnetic moment can be calculated using Eq. (9) [35].
| (9) |
Figure 6.
Magnetization curves of the Mn1-xZnxFe2O4 ferrofluids at room temperature.
By substituting the magnetic moment values of Fe2+ (4 μB), Mn2+ (5 μB), Zn2+ (0 μB), Fe3+ (5 μB), and Mn3+ (4 μB) into Eq. (9) and Eqs. (1), (2), (3), and (4), the net magnetic moment in the spinel system produced, as shown in Table 3. The calculated magnetic moment value has a different saturation magnetization pattern for each sample. Therefore, the magnetic properties of the ferrofluids are not only influenced by the net magnetic moment and exchange interaction, but also influenced by other factors, including particle dispersion, nanoparticle arrangement, and the energy of dipole interaction of two contacting particles [36]. Thus, the SAXS data provides an important role in explaining the correlation of the structure parameters of the ferrofluids to their magnetic properties.
Table 3.
Magnetic properties of the Mn1-xZnxFe2O4 ferrofluids.
| Sample | Magnetic moment (μB) | Saturation magnetization (emu/g) | Remanent magnetization (emu/g) | Secondary particle (nm) | Polydispersity index |
|---|---|---|---|---|---|
| Fe3O4 | 4.00 | 0.053 ± 0.004 | 3.4 ± 0.2 | 11.1 ± 0.1 | 0.13 ± 0.03 |
| MnFe2O4 | 5.00 | 0.046 ± 0.004 | - | 10.5 ± 0.1 | 0.11 ± 0.02 |
| Mn0.75Zn0.25Fe2O4 | 4.59 | 0.086 ± 0.007 | 2.4 ± 0.1 | 13.9 ± 0.2 | 0.38 ± 0.05 |
| Mn0.5Zn0.5Fe2O4 | 7.00 | 0.188 ± 0.008 | 4.2 ± 0.2 | 16.8 ± 0.3 | 0.41 ± 0.04 |
| Mn0.25Zn0.75Fe2O4 | 4.41 | 0.079 ± 0.006 | 2.0 ± 0.1 | 13.0 ± 0.2 | 0.26 ± 0.02 |
| ZnFe2O4 | 6.04 | 0.070 ± 0.005 | 4.1 ± 0.2 | 12.8 ± 0.1 | 0.29 ± 0.05 |
In the ferrofluids, strong inter-particle interactions cause large magnetic particles to form chains. Under an external magnetic field, small magnetic particles can still maintain Brownian motion or link to the end of the chain. This behavior contributes to the change in the coercivity field and remanent magnetization of the ferrofluids. Based on Table 2, the ZnFe2O4 ferrofluid sample has a large average aspect ratio (chain correlation length ξ = 68.9 nm with secondary particles = 12.8 nm), which indicates that the sample has a large average anisotropy shape and implies an increase in the remanent magnetization. In this work, ZnFe2O4 is ferrimagnetic because the magnetic moments located in the tetrahedral position are opposites within the octahedral position, resulting in unequal magnetic moments in the spinel system. The spinel system can be divided into three types: normal, inverse, and mixed spinel systems. In the previous work, Grasset et al. showed that the ZnFe2O4 with a normal spinel exhibits antiferromagnetic character [37]. The diamagnetic Zn2+ ions are only at tetrahedral sites and Fe3+ ions are only at octahedral sites and they are coupled with each other via a superexchange pathway through tetrahedral sites. Furthermore, other previous works showed that zinc ferrite nanoparticles are antiferromagnetic in a normal spinel system and with bulk materials; however, they show ferrimagnetic behavior at the nanometer scale. Other reports similarly showed that ZnFe2O4 does not construct a normal spinel system. ZnFe2O4 nanoparticles show ferrimagnetic behavior at room temperature and antiferromagnetic ordering below 9 K [38]. Other showed that the ZnFe2O4 nanoparticles have ferrimagnetic behavior at room temperature [39, 40]. Interestingly, the ZnFe2O4 ferrofluid sample has a fractal dimension of D ~1, which is interpreted as a chain arrangement resembling a rod. According to the theory, the demagnetization factor of the chain arrangement can be expressed by Eq. (10) [41].
| (10) |
The longer the chain arrangement, the smaller the Dfactor. When the Dfactor is low, the remanent magnetization increases. Furthermore, other ferrofluid samples have larger fractal dimension that interprets a denser chain arrangement with shorter chain lengths. Therefore, the coercivity field and remanent magnetization decrease.
Interestingly, the Mn0.5Zn0.5Fe2O4 ferrofluids have a superparamagnetic character with the highest saturation magnetization value. However, compared to the bulk samples, the saturation magnetization of the ferrofluids was lower because of the effect of the surfactant and carrier liquid [42]. The surfactant and carrier liquid in the ferrofluids strongly contribute to creating repulsion forces between particles, thereby overcoming aggregation. If the particle interaction force is greater than the electrostatic repulsion force, the ferrofluids will be polydisperse; otherwise, the ferrofluids will be monodisperse. Based on the SAXS data analysis, the Mn1-xZnxFe2O4 ferrofluids belong to a bidisperse system because they contain two different particle sizes, namely primary and secondary particles. In the ferrofluid system, the small particles have Brownian motion, while the large particles have a stronger coupling interaction and dominate the Brownian motion [43]. Therefore, the saturation magnetization also depends on the behavior of the large particles. High saturation magnetization occurs with samples that have a large particle size and a polydisperse index, which is consistent with previous studies [42, 43, 44, 45].
The performance of the magnetic sensor was investigated by measuring the sensitivity of the Mn1-xZnxFe2O4 ferrofluids under an external magnetic field. The external magnetic field increased until the magnetic moment of the ferrofluids achieved saturation. Furthermore, the magnetic field was then reduced to obtain an output voltage as a function of the magnetic field, as shown in Figure 7. In general, when the ferrofluids was subject to an increasing external magnetic field, the output voltage also increased and vice versa. For the Fe3O4 ferrofluids, the splitting curve was observed to start from the magnetic field of 10 mT. Therefore, the stability of the Fe3O4 ferrofluids as magnetic sensors was not optimum. However, the MnFe2O4 ferrofluids had high stability, which was identified from the relative absence of the splitting curve. This shows that MnFe2O4 ferrofluids have high potential to be used as magnetic sensor materials for a small magnetic field. The results of this work coincide with the magnetic sensor developed by Zhao and co-workers based on ferrofluids combining with the photonic crystal fiber [46]. They observed that the ferrofluids characteristics change rapidly because of the magnetic field. The nearly coinciding curves of the increasing and decreasing magnetic field proves that the sensor performance is ideal. Furthermore, for Mn0.75Zn0.25Fe2O4 ferrofluids, the magnetic sensor performance of the Mn0.75Zn0.25Fe2O4 ferrofluids is not as good as the MnFe2O4 ferrofluids for fields greater than 35 mT. This difference can be explained by the presence of Zn ions, which substitute the Mn ion to affect the magnetic moment of the sample. Moreover, the sensor performance curve of the Mn0.5Zn0.5Fe2O4 ferrofluids fluctuates above the 35 mT because of the increased molar fraction of Zn ions. The Mn0.25Zn0.75Fe2O4 and ZnFe2O4 ferrofluids had a splitting curve, although with increasing Zn ion composition, the splitting becomes relatively narrow. Thus, the best magnetic sensor performance was observed with the MnFe2O4 ferrofluids, which have excellent stability in both up and down fields.
Figure 7.
Magnetic sensor performance of the Mn1-xZnxFe2O4 ferrofluids at room temperature.
The particle relaxation time is essential to ensure sensor performance. The output voltage as a function of the particle relaxation time of the Mn1-xZnxFe2O4 ferrofluids is shown in Figure 8. In a relatively short time, the magnetic particles in the ferrofluids were able to return to the initial state. This improves the sensor performance because the ferrofluids do not require a long time to relax, indicating that the ferrofluids as a magnetic sensor material have excellent flexibility. In general, the magnetic particles of the Mn1-xZnxFe2O4 ferrofluids, except Mn0.5Zn0.5Fe2O4 ferrofluids, tended to be easily and quickly stable after a mechanical treatment less than 3 min. We predict that this phenomenon is caused by secondary particles, where the Mn0.5Zn0.5Fe2O4 ferrofluids are the largest. Because the secondary particles are large, the magnetic particles become easier and faster to precipitate. In this case, the particle relaxation time is defined as the time to of the magnetic particles of the ferrofluids to reach stability. The particle relaxation time is influenced by the particle motion characteristics of the ferrofluids. The closer the particles, the more limited the particle motion under mechanical treatment and it will quickly return to its original state after the treatment is removed [47]. This phenomenon is consistent with research conducted by Zhou and co-workers, demonstrating that ferrofluids have high flexibility [48]. Thus, if the ferrofluids are given an external treatment, then after the treatment is removed, it will quickly return to its original state.
Figure 8.
Particle relaxation time of the Mn1-xZnxFe2O4 ferrofluids at room temperature.
The radar absorption performance of the Mn1-xZnxFe2O4 nanoparticles was evaluated using a VNA at room temperature, and its result is presented in Figure 9. The experiment was conducted at a frequency range of 8–12 GHz for the working area of the radar. The reflection loss (RL) as the ability of the samples to absorb the radar was evaluated using Eq. (11):
| (11) |
where μr represents the complex permeability of the absorber, εr represents the complex permittivity of the absorber, f represents the frequency, d represents the thickness of the absorber, and c represents the velocity of light in a vacuum [49]. In principle, good radar absorbing materials have special characteristics such as lightweight, thin, and the ability to cover a broad frequency [50].
Figure 9.
Reflection loss of the Mn1-xZnxFe2O4 nanoparticles at a frequency range of 8–12 GHz.
Based on Figure 9, the RL of the samples was −14.1, −15.4, −12.9, −11.0, −12.9, and −11.7 dB for the Fe3O4, MnFe2O4, Mn0.75Zn0.25Fe2O4, Mn0.5Zn0.5Fe2O4, Mn0.25Zn0.75Fe2O4, and ZnFe2O4 samples, respectively. Such RL was observed at frequencies of 10.8, 10.9, 10.8, 10.8, 10.9, and 10.8 GHz, respectively. The maximum peak was observed for the MnFe2O4 sample, indicating that this sample exhibited the best performance as a radar absorbing material. Akinay and co-workers identified the radar absorbing performance of a polyvinyl butyral (PVB), Fe3O4, and nickel ferrite nanoparticle composite with various thicknesses [51]. The absorption peak was in the frequency range of 1.0–3.0 GHz for the PVB/Fe3O4 nanocomposites and 1.0–2.6 GHz for the PVB/NiFe2O4 nanocomposites with a minimum thickness of 3 mm. Theoretically, the absorbing phenomenon of the magnetic materials is initiated by the changing energy of the electromagnetic waves when the magnetic dipole moves to rotate in the material [52]. Based on the data analysis, increasing Zn or decreasing Mn in the Mn1-xZnxFe2O4 nanoparticles decreases the RL because Zn is a diamagnetic material, reducing the energy needed to rotate the magnetic moment [53]. The Mn0.5Zn0.5Fe2O4 nanoparticles had a minimum RL believed to be the effect of the largest secondary particle (16.8 nm) with the highest polydispersity index (0.41). With a larger particle size, the ratio of the surface area to volume is smaller, causing low absorption of the materials. Interestingly, the MnFe2O4 nanoparticles with the highest Mn composition had a maximum RL. Based on the theoretical calculation, Mn2+ with a higher magnetic moment (5 μB) replaced Fe2+ with a lower magnetic moment (4 μB). Furthermore, the MnFe2O4 nanoparticles with the smallest primary particles as building blocks contributed to the increasing surface area. Therefore, the MnFe2O4 nanoparticles more easily absorbed the microwave. The microwave directed to the nanoparticles in the Mn1-xZnxFe2O4 ferrofluids also interacts with the TMAH surfactant agent and carrier liquid that has dielectric characteristics. Therefore, a Coulomb force appears owing to the interaction between the electric field from the microwave and free radicals from the surfactant and dispersant [54]. Therefore, the microwave stimulates the charge acceleration that produces an electric current. In addition, the magnetic field of the ferrofluids generates destructive superposition with the opposite phase under the microwave [55]. Consequently, the radar that interacts with the Mn1-xZnxFe2O4 ferrofluids was difficult to detect by the radar receiver. As a result, our Mn1-xZnxFe2O4 ferrofluids are expected to form a novel platform for advancing radar absorbing materials.
4. Conclusion
In this study, the synthesis of Mn1-xZnxFe2O4 ferrofluids was performed through a coprecipitation method employing natural iron sand as the main precursor. As the Zn molar fraction increases, the lattice parameters, crystal volume, and particle size of the ferrofluids tended to decrease. In general, the Mn1-xZnxFe2O4 nanoparticles had a spherical shape that were nanometer in scale. The functional groups for the main components of the Mn1-xZnxFe2O4 ferrofluids were detected from the metal oxides, TMAH, and H2O used as the filler, surfactant, and carrier liquid, respectively. The aggregation characteristics strongly contributed to the saturation magnetization of the Mn1-xZnxFe2O4 ferrofluids. Furthermore, the potential of the ferrofluids as magnetic sensors was demonstrated by excellent performance, especially for the MnFe2O4 ferrofluids. For radar absorbing applications, the maximum Mn composition presented the best performance because it was the smallest building block of the magnetic particles. Therefore, the ecofriendly Mn1-xZnxFe2O4 nanoparticles fabricated from natural sand are potential magnetic sensors and radar absorbing materials.
Declarations
Author contribution statement
Ahmad Taufiq: Conceived and designed the experiments; Analyzed and interpreted the data; Wrote the paper.
Syamsul Bahtiar: Performed the experiments; Analyzed and interpreted the data.
Rosy Eko Saputro: Analyzed and interpreted the data; Wrote the paper.
Defi Yuliantika: Contributed reagents, materials, analysis tools or data.
Arif Hidayat: Analyzed and interpreted the data; Wrote the paper.
Sunaryono Sunaryono: Conceived and designed the experiments.
Nurul Hidayat: Analyzed and interpreted the data.
Samian Samian: Conceived and designed the experiments; Performed the experiments.
Siriwat Soontaranon: Analyzed and interpreted the data; Contributed reagents, materials, analysis tools or data.
Funding statement
This work was supported by Kementerian Riset Teknologi Dan Pendidikan Tinggi Republik Indonesia (071/SP2H/LT/DRPM/2018).
Competing interest statement
The authors declare no conflict of interest.
Additional information
No additional information is available for this paper.
References
- 1.Thakur P., Chahar D., Taneja S., Bhalla N., Thakur A. A review on MnZn ferrites: synthesis, characterization and applications. Ceram. Int. 2020;46:15740–15763. doi: 10.1016/j.ceramint.2020.03.287. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 2.Srinivas Ch., Kumar E.R., Ramesh P.N., Krishna M., Singh M.S., Prajapat C.L., Verma A., Sastry D.L. Evaluation of structural and dielectric properties of Mn2+-substituted Zn-spinel ferrite nanoparticles for gas sensor applications. Sensor. Actuator. B Chem. 2020;316:128127. [Google Scholar]
- 3.Lopes B.H.K., Portes R.C., do Amaral Junior M.A., Florez-Vergara D.E., Gama A.M., Silva V.A., Quirino S.F., Baldan M.R. X Band electromagnetic property influence of multi-walled carbon nanotube in hybrid MnZn ferrite and carbonyl iron composites. J. Mater. Res. Technol. 2020;9:2369–2375. [Google Scholar]
- 4.Ma T., Cui Y., Sha Y., Liu L., Ge J., Meng F., Wang F. Facile synthesis of hierarchically porous rGO/MnZn ferrite composites for enhanced microwave absorption performance. Synth. Met. 2020;265:116407. [Google Scholar]
- 5.Auzans E., Zins D., Blums E., Massart R. Synthesis and properties of Mn-Zn ferrite ferrofluids. J. Mater. Sci. 1999;34:1253–1260. [Google Scholar]
- 6.Haritha T., Ramji K., Murthy K.K., Nagasree P.S., Nagesh D.B. Springer; Singapore: 2020. Synthesis and Microwave Absorption Properties of MnZn Ferrite Nanocomposite, Recent Trends in Mechanical Engineering; pp. 419–429. [Google Scholar]
- 7.Teber A., Cil K., Yilmaz T., Eraslan B., Uysal D., Surucu G., Baykal A.H., Bansal R. Manganese and zinc spinel ferrites blended with multi-walled carbon nanotubes as microwave absorbing materials. Aerospace. 2017;4:2. [Google Scholar]
- 8.van Silfhout A., Erné B. Magnetic detection of nanoparticle sedimentation in magnetized ferrofluids. J. Magn. Magn Mater. 2019;472:53–58. [Google Scholar]
- 9.El Mendili Y., Grasset F., Randrianantoandro N., Nerambourg N., Greneche J.-M., Bardeau J.-F. Improvement of thermal stability of maghemite nanoparticles coated with oleic acid and oleylamine molecules: investigations under laser irradiation. J. Phys. Chem. C. 2015;119:10662–10668. [Google Scholar]
- 10.Latikka M., Backholm M., Timonen J.V.I., Ras R.H.A. Wetting of ferrofluids: phenomena and control. Curr. Opin. Colloid Interface Sci. 2018;36:118–129. [Google Scholar]
- 11.Spizzo F., Sgarbossa P., Sieni E., Semenzato A., Dughiero F., Forzan M., Bertani R., Del Bianco L. Synthesis of ferrofluids made of iron oxide nanoflowers: interplay between carrier fluid and magnetic properties. Nanomaterials. 2017;7:373. doi: 10.3390/nano7110373. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 12.Lv R.-Q., Zhao Y., Wang D., Wang Q. Magnetic fluid-filled optical fiber Fabry–Pérot sensor for magnetic field measurement. IEEE Photon. Technol. Lett. 2014;26:217–219. [Google Scholar]
- 13.Shojaeizadeh E., Veysi F., Goudarzi K., Feyzi M. Magnetoviscous effect investigation of water based Mn-Zn Fe2O4 magnetic nanofluid under the influence of magnetic field: an experimental study. J. Magn. Magn Mater. 2019;477:292–306. [Google Scholar]
- 14.Arulmurugan R., Vaidyanathan G., Sendhilnathan S., Jeyadevan B. Mn–Zn ferrite nanoparticles for ferrofluid preparation: study on thermal–magnetic properties. J. Magn. Magn. Mater. 2006;298:83–94. [Google Scholar]
- 15.Taufiq A., Nikmah A., Hidayat A., Sunaryono S., Mufti N., Hidayat N., Susanto H. Synthesis of magnetite/silica nanocomposites from natural sand to create a drug delivery vehicle. Heliyon. 2020;6 doi: 10.1016/j.heliyon.2020.e03784. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 16.Hunter B. Rietica - a visual Rietveld program. IUCr Comm. Powder Diffr. Newsl. 1998;20:21. [Google Scholar]
- 17.Lasheras X., Insausti M., de la Fuente J.M., Gil de Muro I., Castellanos-Rubio I., Marcano L., Fernández-Gubieda M.L., Serrano A., Martín-Rodríguez R., Garaio E., García J.A., Lezama L. Mn-doping level dependence on the magnetic response of MnxFe3−xO4 ferrite nanoparticles. Dalton Trans. 2019;48:11480–11491. doi: 10.1039/c9dt01620a. [DOI] [PubMed] [Google Scholar]
- 18.Taufiq A., Sunaryono, Rachman Putra E.G., Okazawa A., Watanabe I., Kojima N., Pratapa S., Darminto Nanoscale clustering and magnetic properties of MnxFe3−xO4 particles prepared from natural magnetite. J. Supercond. Nov. Magn. 2015;28:2855–2863. [Google Scholar]
- 19.Li M., Gao Q., Wang T., Gong Y.-S., Han B., Xia K.-S., Zhou C.-G. Solvothermal synthesis of MnxFe3−xO4 nanoparticles with interesting physicochemical characteristics and good catalytic degradation activity. Mater. Des. 2016;97:341–348. [Google Scholar]
- 20.Taufiq A., Sunaryono, Hidayat N., Hidayat A., Putra E.G.R., Okazawa A., Watanabe I., Kojima N., Pratapa S., Darminto Studies on nanostructure and magnetic behaviors of mn-doped black iron oxide magnetic fluids synthesized from iron sand. Nano. 2017;12:1750110. [Google Scholar]
- 21.Bezerra D.M., Rodrigues J.E.F.S., Assaf E.M. Structural, vibrational and morphological properties of layered double hydroxides containing Ni2+, Zn2+, Al3+ and Zr4+ cations. Mater. Char. 2017;125:29–36. [Google Scholar]
- 22.Kaur N., Chudasama B. Structure induced tunable magnetic properties of Zn substituted Mn1−xZnxFe2O4 (x = 0–1) NPs. Micro Nano Lett. 2017;12:151–156. [Google Scholar]
- 23.Siddeeg S.M., Amari A., Tahoon M.A., Alsaiari N.S., Rebah F.B. Removal of meloxicam, piroxicam and Cd+2 by Fe3O4/SiO2/glycidyl methacrylate-S-SH nanocomposite loaded with laccase. Alexandria Eng. J. 2020;59:905–914. [Google Scholar]
- 24.Klencsár Z., Tolnai Gy., Korecz L., Sajó I., Németh P., Osán J., Mészáros S., Kuzmann E. Cation distribution and related properties of MnxZn1−xFe2O4 spinel nanoparticles. Solid State Sci. 2013;24:90–100. [Google Scholar]
- 25.Liu J., Bin Y., Matsuo M. Magnetic behavior of Zn-doped Fe3O4 nanoparticles estimated in terms of crystal domain size. J. Phys. Chem. C. 2012;116:134–143. [Google Scholar]
- 26.Priyananda P., Sabouri H., Jain N., Hawkett B.S. Steric Stabilization of γ-Fe2O3 superparamagnetic nanoparticles in a hydrophobic ionic liquid and the magnetorheological behavior of the ferrofluid. Langmuir. 2018;34:3068–3075. doi: 10.1021/acs.langmuir.7b04291. [DOI] [PubMed] [Google Scholar]
- 27.Yuliantika D., Taufiq A., Putra E.G.R. Hierarchical structure and antibacterial activity of olive oil based MZFe2O4 ferrofluids. J. Phys. Conf. 2020;1436 [Google Scholar]
- 28.Veligzhanin A.A., Frey D.I., Shulenina A.V., Gruzinov A.Yu., Zubavichus Ya.V., Avdeev M.V. Characterization of aggregate state of polydisperse ferrofluids: some aspects of anisotropy analysis of 2D SAXS in magnetic field. J. Magn. Magn. Mater. 2018;459:285–289. [Google Scholar]
- 29.Cherny A.Y., Anitas E.M., Osipov V.A., Kuklin A.I. Scattering from surface fractals in terms of composing mass fractals. J. Appl. Crystallogr. 2017;50:919–931. [Google Scholar]
- 30.Paula F.L.D.O. SAXS Analysis of magnetic field influence on magnetic nanoparticle clusters. Condens. Matter. 2019;4:55. [Google Scholar]
- 31.Woźniak E., Špírková M., Šlouf M., Garamus V.M., Šafaříková M., Šafařík I., Štěpánek M. Stabilization of aqueous dispersions of poly(methacrylic acid)-coated iron oxide nanoparticles by double hydrophilic block polyelectrolyte poly(ethylene oxide)-block-poly(N-methyl-2-vinylpyridinium iodide) Colloids Surfaces A Physicochem. Eng. Aspects. 2017;514:32–37. [Google Scholar]
- 32.Nguyen H.D., Nguyen T.D., Nguyen D.H., Nguyen P.T. Magnetic properties of Cr doped Fe3O4 porous nanoparticles prepared through a co-precipitation method using surfactant. Adv. Nat. Sci. Nanosci. Nanotechnol. 2014;5 [Google Scholar]
- 33.Khan F., Baek S.-H., Kim J.H. One-step and controllable bipolar doping of reduced graphene oxide using TMAH as reducing agent and doping source for field effect transistors. Carbon. 2016;100:608–616. [Google Scholar]
- 34.Taufiq A., Ikasari F.N., Yuliantika D., Sunaryono S., Mufti N., Susanto H., Suarsini E., Hidayat N., Fuad A., Hidayat A., Diantoro M. Structural, magnetic, optical and antibacterial properties of magnetite ferrofluids with PEG-20000 template. Mater. Today Proc. 2019;17:1728–1735. [Google Scholar]
- 35.Callister W.D., Rethwisch D.G. eighth ed. John Wiley & Sons; Hoboken, NJ: 2010. Materials Science and Engineering: an Introduction. [Google Scholar]
- 36.Gutiérrez L., de la Cueva L., Moros M., Mazarío E., de Bernardo S., de la Fuente J.M., Morales M.P., Salas G. Aggregation effects on the magnetic properties of iron oxide colloids. Nanotechnology. 2019;30:112001. doi: 10.1088/1361-6528/aafbff. [DOI] [PubMed] [Google Scholar]
- 37.Grasset F., Labhsetwar N., Li D., Park D.C., Saito N., Haneda H., Cador O., Roisnel T., Mornet S., Duguet E., Portier J., Etourneau J. Synthesis and magnetic characterization of zinc ferrite nanoparticles with different environments: powder, colloidal solution, and zinc ferrite–silica core–shell nanoparticles. Langmuir. 2002;18:8209–8216. [Google Scholar]
- 38.Andjelković L., Šuljagić M., Lakić M., Jeremić D., Vulić P., Nikolić A.S. A study of the structural and morphological properties of Ni–ferrite, Zn–ferrite and Ni–Zn–ferrites functionalized with starch. Ceram. Int. 2018;44:14163–14168. [Google Scholar]
- 39.Raveendran S., Alam M.M., Khan Mohd.I.K., Dhayalan A., Kannan S. In situ formation, structural, mechanical and in vitro analysis of ZrO2/ZnFe2O4 composite with assorted composition ratios. Mater. Sci. Eng. C. 2020;108:110504. doi: 10.1016/j.msec.2019.110504. [DOI] [PubMed] [Google Scholar]
- 40.Rodríguez K.L.S., Quintero J.J.M., Medina Chanduví H.H., Rebaza A.V.G., Faccio R., Adeagbo W.A., Hergert W., Torres C.E.R., Errico L.A. Ab-initio approach to the stability and the structural, electronic and magnetic properties of the (001) ZnFe2O4 surface terminations. Appl. Surf. Sci. 2020;499:143859. [Google Scholar]
- 41.Wang J., Cao S., Xia S., Gan N. The assembly of chain-like Ni arrays under magnetic field and its enhanced properties. Int. J. Nanosci. 2010;9:543–547. [Google Scholar]
- 42.Ivanov A.O., Kantorovich S.S., Reznikov E.N., Holm C., Pshenichnikov A.F., Lebedev A.V., Chremos A., Camp P.J. Magnetic properties of polydisperse ferrofluids: a critical comparison between experiment, theory, and computer simulation. Phys. Rev. E. 2007;75 doi: 10.1103/PhysRevE.75.061405. [DOI] [PubMed] [Google Scholar]
- 43.Huang J.P., Holm C. Magnetization of polydisperse colloidal ferrofluids: effect of magnetostriction. Phys. Rev. E. 2004;70 doi: 10.1103/PhysRevE.70.061404. [DOI] [PubMed] [Google Scholar]
- 44.Russier V., de Montferrand C., Lalatonne Y., Motte L. Size and polydispersity effect on the magnetization of densely packed magnetic nanoparticles. J. Appl. Phys. 2012;112 [Google Scholar]
- 45.Xu-Fei W., Li-Qun S. Polydispersity effects on the magnetization of diluted ferrofluids: a lognormal analysis. Chinese Phys. B. 2010;19:107502. [Google Scholar]
- 46.Zhao Y., Wu D., Lv R.-Q. Magnetic field sensor based on photonic crystal fiber taper coated with ferrofluid. IEEE Photon. Technol. Lett. 2015;27:26–29. [Google Scholar]
- 47.Sólyom J. Springer; Berlin ; New York: 2007. Fundamentals of the Physics of Solids: Structure and Dynamics. [Google Scholar]
- 48.Zhou Y., Song L., Yu L., Xuan X. Inertially focused diamagnetic particle separation in ferrofluids. Microfluid. Nanofluidics. 2017;21:1–14. [Google Scholar]
- 49.Shen G., Xu M., Xu Z. Double-layer microwave absorber based on ferrite and short carbon fiber composites. Mater. Chem. Phys. 2007;105:268–272. [Google Scholar]
- 50.Munir A. Microwave radar absorbing properties of multiwalled carbon nanotubes polymer composites: a review. Adv. Polym. Technol. 2017;36:362–370. [Google Scholar]
- 51.Akinay Y., Hayat F., Çakir M., Akin E. Magnetic and microwave absorption properties of PVB/Fe3O4 and PVB/NiFe2O4 composites. Polym. Compos. 2018;39:3418–3423. [Google Scholar]
- 52.Praveena K., Sadhana K., Matteppanavar S., Liu H.-L. Effect of sintering temperature on the structural, dielectric and magnetic properties of Ni0.4Zn0.2Mn0.4Fe2O4 potential for radar absorbing. J. Magn. Magn Mater. 2017;423:343–352. [Google Scholar]
- 53.Smitha P., Singh I., Najim M., Panwar R., Singh D., Agarwala V., Das Varma G. Development of thin broad band radar absorbing materials using nanostructured spinel ferrites. J. Mater. Sci. Mater. Electron. 2016;27:7731–7737. [Google Scholar]
- 54.Zare Y., Shams M.H., Jazirehpour M. Tuning microwave permittivity coefficients for enhancing electromagnetic wave absorption properties of FeCo alloy particles by means of sodium stearate surfactant. J. Alloys Compd. 2017;717:294–302. [Google Scholar]
- 55.Li W., Li C., Lin L., Wang Y., Zhang J. All-dielectric radar absorbing array metamaterial based on silicon carbide/carbon foam material. J. Alloys Compd. 2019;781:883–891. [Google Scholar]









