Next Article in Journal
Novel Nitride Materials Synthesized at High Pressure
Next Article in Special Issue
Valorization of Powder Obtained from Marble Sludge Waste and Its Suitability as a Mineral Filler
Previous Article in Journal
A Review of the Latest Developments in the Field of Refractory High-Entropy Alloys
Previous Article in Special Issue
π-Extended Boron Difluoride [NՈNBF2] Complex, Crystal Structure, Liquid NMR, Spectral, XRD/HSA Interactions: A DFT and TD-DFT Study
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Structural and Magnetic Properties Study of Fe2O3/NiO/Ni2FeO4 Nanocomposites

1
Department of Physics, College of Science, King Faisal University, P.O. Box 400, Al-Ahsa 31982, Saudi Arabia
2
Semiconductors Technology Lab, Physics Department, Faculty of Science, Ain Shams University, Cairo 11566, Egypt
*
Authors to whom correspondence should be addressed.
Crystals 2021, 11(6), 613; https://doi.org/10.3390/cryst11060613
Submission received: 1 April 2021 / Revised: 19 May 2021 / Accepted: 26 May 2021 / Published: 29 May 2021
(This article belongs to the Special Issue New Trends in Crystals at Saudi Arabia)

Abstract

:
In the current work, the nanocomposites that consist chiefly of three components—α-Fe2O3, NiO and Ni2FeO4, in two different ratios 2:2:1 (FNN-221) and 2:1:1 (FNN-211), respectively—were produced. The synthesis was done in two steps by following the chemical co-precipitation and mechanical ball-milling route. The presence of individual phase was identified from the XRD data without the detection of any additional impurities. The phase fraction of each component estimated from the profile fitting of XRD patterns were found to be 41.2%, 39.7%, 19.1% in FNN-221 sample and 49.5%, 26.4%, 24.1% for FNN-211 sample, respectively, which were consistent with the experimental values. The total magnetization at 300 K was obtained to be 13.41 emu/g and 10.95 emu/g for FNN-221 and FNN-211 samples, respectively. In FNN-211 compound the zero field coercivity (HC) expanded towards the higher field values thereby signifying the exchange bias behavior. Furthermore, the exchange bias field (Hex) for FNN-211 was obtained as 35.1 Oe.

1. Introduction

Recently, the multi-component magnetic nanoparticles captivated an enormous deal of research and technological interests among scientific communities. An important advantage of such materials over single phased magnetic nanostructures is that the magnetic properties of the former are deeply correlated to the interfacial interactions and contact between the components. Thus, their physical properties can be drastically tailored/induced simply by adjusting the chemical compositions, geometrical shapes and ratios of different components [1,2,3,4,5]. Because of these characteristics, the magnetic nanocomposites can be directly exploited in several new technological fields like spintronic devices, hyperthermia treatment, permanent magnets, photocatalysts and high-density magnetic storage media [6,7,8,9]. Nanocomposites involving the ferrites (oxides of iron) are the most extensively investigated materials in this context. Lately, the methods for the production of ferrite nanocomposites have been centered around the fabrication of metal doped ferrites into a non-magnetic and/or matrix with weak magnetic polarity (SiO2, ZnO, titanates) [2,10,11,12]. The contrast in the magnetic polarity between the host matrix and the foreign particle facilitates a variation in the degree interactions among the charge carriers. Thereby the magnetic transition and anisotropic nature can be affected. The ferrites of Cobalt (Co) and Nickel (Ni) have shown synergistic improvement in their chemical and thermal stability upon dispersion in a porous silica matrix [13,14,15]. Moreover, the low temperatures anomalies in the magnetic hysteresis and modification in the internal coercive field in the case of binary ferrites-ZnO nanocomposites, have also been mentioned in the literature [10,16,17]. Not long ago Galizia et al. investigated a “wasp waist” hysteresis loop in spinel ferrites-composites. It has been noticed that the wasp-waist shape of hysteresis in the spinel ferrites composites is sensitive toward the phase fraction/ratios and particle size of the forming components [18]. Similarly, an exchange spring like physical phenomenon in the ferrites-graphene oxide nanocomposites has been unfolded by El-Khawas et al. [19]. The principal origin of exchange bias in such systems is the staggered surface spins at the interface between the two components [20]. In addition, the influence of particle size and synthesis routes on the structural and magnetic properties in ferrites has been frequently studied by numerous researchers [21,22,23]. Nevertheless, the mechanism of interaction between dual and triple magnetic phases in the ferrites, especially at the low temperatures is still not well understood from the available literature.
Thus, in the present work, we have fabricated triple phase nanocomposites chiefly including three components: spinel Ni-ferrite (Ni2FeO4); hematite (α-Fe2O3) and nickel oxide (NiO). In order to investigate the influence of each phase on the structural and magnetic properties, the different ratios of each component have been selected and synthesized using chemical Co-precipitation route. In general, both NiO and Ni2FeO4 are antiferromagnetic below their Neel temperatures (TN) [24]. On the other hand, depending on the synthesis conditions and particle geometries, α-Fe2O3 can exhibit anti-ferromagnetism, weak ferromagnetism and even super-paramagnetism [2,6,25]. The magnetic response of the nanocomposites had been compared with the parent α-Fe2O3 component in order to highlight the involvement of particle interactions. The results obtained here, signifies that due to the involvement of spin functionalities the total magnetization of the nanocomposites has been enhanced by several folds compared to the of parent α-Fe2O3 compound. Meanwhile, the current work is also important in the sense that the interactions among multiple magnetic polarities have been investigated under the influence of several magnetic fields and temperature conditions in the single study. Based on these measurements the appearance of exchange bias effect in the case of composite structures has been identified. Such type of study is important to the literature in order to understand the magnetic behavior in three component systems.

2. Experimental Details

Initially, the nanoparticles of all the three components namely: Ni2FeO4, α-Fe2O3 and NiO were individually synthesized through the chemical co-precipitation route. The highly pure (≥99.5%; Sigma Aldrich, MI, USA) analytical grade regents of FeCl3.6H2O, Ni(NO3)2.6H2O and NiCl2.6H2O were used as starting precursors during the synthesis procedure. NaOH and aqueous NH4OH solutions were employed as precipitator for the synthesis of Ni2FeO4/NiO and α-Fe2O3, respectively. The molar amounts of each salt have been taken as such displayed in Table 1. First, the solutions of individual precursors were prepared separately by dissolving the respective molar amounts of salts in the mentioned amount of deionized (D.I) water. The solutions were subjected to continuous stirring (500 rpm) for 40 min at 60 °C to ensure the homogenous mixing of each salts in the D.I water. Once the salts were dissolved into the D.I water the reaction kinetics and the particle size were controlled through the dropwise (1 mL/min) addition of precipitator into the solutions. The reaction has been stopped upon the formation of precipitates and the reaction mixture was stirred for 1 h under ambient temperature and pressure. The obtained slurry was then washed several times with acetone and the dried in the oven at 70 °C overnight. In the second step the nanocomposites with two different ratios were prepared and labelled as FNN-221 for 2:2:1 ratio and FNN-211 for 2:1:1 ratio of α-Fe2O3, NiO and Ni2FeO4 powders, respectively (see Table 1). The powders of as synthesized α-Fe2O3, NiO and Ni2FeO4 compounds were weighted in the corresponding amounts and added in ball-milling vials with balls (diameter ~7 mm) to powder ratio of 1:20 for each sample. A small amount of ethanol was added into the vials as a buffer media along with a few drops of polyvinyl-alcohol (PVA; 2 wt.%). The samples were ball-milled for 6 h with a milling speed of 180 rpm. The obtained mixtures were dried into the oven at 70 °C. The obtained powders were heat treated at 300 °C to dissolve PVA and fuse the particles.
The crystalline structure, phase fraction and crystallite sizes of the samples were characterized through the X-ray diffraction (XRD) patterns. The XRD patterns were obtained from Bruker D8 advance diffractometer (Cu-Kα radiation). For Rietveld refinement of XRD profiles, the pseudo-Voigt function has been used. The background has been corrected through the linear interpolation between a set of background points. Scanning electron microscope (FE-SEM; Merlin compact) was employed for analyzing the particle shapes and sizes. Elemental mapping was done with Energy Dispersive X-Ray Analysis (EDX; Oxford). The magnetic measurements were carried out using vibrating sample magnetometer (VSM) equipped in physical properties measurement system (PPMS EverCool-II, Quantum Design, San Diego, CA, USA).

3. Results and Discussion

Figure 1a,b shows the XRD pattern of the prepared nanocomposites. For both the samples (FNN-221 and FNN-211) the diffraction peaks are well indexed according to the ICSD-reference pattern codes 98-002-2505, 98-002-8834 and 98-010-9150 for α-Fe2O3 (space group; R-3c), NiO (space group; Fm-3m) and Ni2FeO4 (space group; Fd-3m), respectively. The absence of any additional peak apart from the desirable three phases suggest that the all three phases were unreacted during the heat treatment, rather expresses their individual crystalline nature. The phase fractions of each component in the XRD profiles were calculated by performing the Rietveld analysis in HighScore Plus (Malvern Panalytical) software and the derived parameters are displayed in Table 2. For FNN-221 sample, the phase fractions of α-Fe2O3, NiO and Ni2FeO4 were found to be 41.2%, 39.7% and 19.1%, respectively. Meanwhile, the three phases contribute 49.5%, 26.4% and 24.1% for FNN-211 samples, respectively. It should be noted that the phase fraction of each compound in both the samples are in perfect agreement with the powder ratios weighted during the synthesis of these composites. Additionally, we observed that the lattice parameters were unaffected towards the change in the ratio of either compounds (shown in Table 2), which is obvious for composites structures. We calculated the crystallite size of the components by employing the Scherrer’s equation and Caglioti function in the HighScore plus. However, for this task the reflections of individual phases were difficult to resolve from the XRD patterns of both samples. Thus, the principal XRD peaks at 2θ = 33.20°, 43.3°, 35.67° corresponding to α-Fe2O3, NiO and Ni2FeO4 compounds, respectively, were accounted. The obtained crystallite sizes are supplied in Table 2. From both methods the crystallite sizes were identified as ~58 nm, ~285 nm and ~49 nm for α-Fe2O3, NiO and Ni2FeO4 components, respectively.
Although, the possibility that these particles might have agglomerated during the formation of composites cannot be neglected. Therefore, for further insights into the particle shapes and arrangements we analyzed the SEM images. The SEM micrographs of both the composite samples are displayed in Figure 2a,b. For both samples the particles assemble in a granular structure with irregular morphology. The granular structure is due to the post synthesis heat treatment of the samples which causes the agglomeration among the particles. These attributions are in excellent consent with the nanocomposites synthesized through chemical route by various authors [26,27,28]. Another finding from the SEM images is the irregular shapes of grains, the irregularity can be described by considering the variation in particles sizes of all the three components. Because of the variable crystallite sizes (see Table 2) of individual phase the degree of agglomeration, also differ for each component, thereby exhibiting the irregularity in granular shape. Additionally, the elemental analysis could be of importance for establishing relation between ratios of each component present in the samples, for this purpose we conducted the EDX analysis. The concentration (wt.%) of each element is presented in Table 2 and the distribution of elements are shown in Figure 2c–n. The existence of all the corresponding elements has been recognized, however their proportions are dissimilar between both the samples. For instance, in FNN-221 the Ni content (21.25%) is considerable higher than the FNN-211 (17.83%) samples. It indicates that about two folds amount of NiO successfully embedded in FNN-221 sample compared to the FNN-211 compound. It is interesting to note that these results are satisfactory with the ratios of phase fractions calculated from the XRD profiles and also consistent with the experimental values.
The phase composition of the synthesized nanocomposites has also been verified from the Raman spectra and the deconvoluted Raman peaks of both the samples are presented in Figure 3a,b. As expected, the phonon modes corresponding to the three components α-Fe2O3, NiO and Ni2FeO4 are present in both the samples. The strong Raman signals located at 405 cm−1, 494 cm−1, 612 cm−1 indicates the existence of α-Fe2O3 phase in both the samples [29]. Similarly, the phonons situated at 571, ~920 cm−1 and 1034 cm−1 allied to the NiO phase [30]. Finally, the Ni2FeO4 phase has been identified from 675 cm−1, 457 cm−1, ~567 cm−1 Raman peaks of these samples [31]. Additionally, it is fascinating to note from the Raman peaks is that in the case of FNN-221 sample the Raman signals representing NiO constitute slightly high signal density compared to the NiO phonons of FNN-211 sample. The relatively higher concentration of NiO Raman signals in FNN-221 samples are in compliance with the experiment and XRD analysis. Therefore, we rigorously analyzed the magnetic properties in the following text.
In order to identify the interaction between particles we performed magnetic measurements under different sets of temperature and magnetic field conditions. First, the variation in magnetization with respect to the temperature (M-T) was analyzed in field cooled (FC) and zero field cool (ZFC) condition. The FC-ZFC curves from 50 K to 400 K under an applied field of 500 Oe for both the nanocomposites are presented in Figure 4a,b.
It has been observed that the ZFC magnetization attain a maximum value at 229 K, 240 K and 269 K for FNN-221 and FNN-211 and α-Fe2O3 compounds, respectively. At the downturn of the onset of this blocking temperature (TB), the ZFC magnetization abruptly declined with decreasing the temperature values and separates from the FC magnetization. This distinction amplified at the low temperatures thereby indicating the thermomagnetic irreversibility between FC-ZFC magnetization. In general, the irreversible ZFC-FC magnetization signifies some sort of anisotropy and/or spin glass state in a magnetic material [32]. In the present work, the appearance of glassy states is obvious because of the co-existence of multiple magnetic components having contrasting magnetic polarities. Further, we see that at 50 K the height of magnetic moment corresponding to Fe2O3 particles is far lower than that of composite structures. This was expected because the α-Fe2O3 particles are weakly ferromagnetic only above 270 K. Below this temperature the antiferromagnetic coupling becomes dominant [6,25]. Additionally, both FC and ZFC magnetization of parent Fe2O3 particles dramatically rises with increasing the temperature and reaches a maximum at 270 K. Thereby, signifying some sort of magnetic coupling below this temperature, we connect this behavior with the Morin transition [25,33]. Moreover, such transition has been completely vanished upon the fabrication of composite structures. In the literature it has been suggested that nature of Morin transition depends on the particle size and/or interparticle interactions. For particles smaller than 50 nm the Morin transition shifts towards the lower values and vanishes for particles smaller than 20 nm [34]. In the present case, we note that the particle sizes of each components hardly vary upon the formation of composite structures (see Table 2) thereby negating any involvement of particle diameters on the Morin transition. Hence, the absence of Morin transition in both the composite samples has been assigned primarily to the interactions between the particles due to their variable spin functionalities. These attributions can be supported by the fact that the temperature dependency of magnetization for the composite structures is significantly higher than parent Fe2O3 compound. Now, our interest is to collect more details on the inter-particle interactions among all the three components, therefore we carried out further magnetic measurements by applying external stimulus i.e., in the presence of applied magnetic field (M-H) at various temperatures (50 K, 100 K, 200 K, 300 K). In Figure 5a,b we compared the M-H magnetization curves of FNN-221, FNN-211 composites with the parent Fe2O3 sample, under a maximum applied field of 3T. In both samples regardless of the isotherms the magnetization abruptly rises with increasing the magnetic field strength, approached a maximum and the magnetization did not saturate up to 3T. It indicates that the magnetic states co-exist in all samples and as a function of applied field strength the effective magnetic moment of each components reverses their polarity towards the applied field thereby inhibiting the total magnetization to reach at the saturation. Moreover, the total magnetization (at 3T) decreases with increasing the temperature (see Figure 5(a1,b1)) which can be expected for a polycrystalline compound. The rise in the temperature promotes thermal disorders among the ordered spins and intensifies the spin degree of motion, thereby causing the reduction in the total magnetization. The effect of temperature on the magnetization and magnetic hysteresis loop can be clearly seen from Figure 5(a2,b2). The parameters obtained from the hysteresis loops of the samples are presented in Table 2. It can be seen that for all the samples the total magnetization (Mt), coercive field (HC) and remanent magnetization (Mr) collectively increases with decreasing the measurement temperature. Here, the increase in coercive field can be described such that at far low temperatures the anisotropy energy increases and more significant than the magnetostatic parameter thereby producing large values of the coercivity [2,35]. Similarly, form Table 2 it should be noted that at all temperatures the Mt values for FNN-221 sample is larger than the FNN-211 sample, the high magnetization value of FNN-221 sample has been ascribed to increase in interparticle interaction due to incorporation of higher number of NiO particles in this sample. On the other hand, the coercivity of FNN-211 sample slightly higher than the FNN-221 sample which has been attributed to competing coupling interactions between the three components. To support this explanation, we have presented the magnetization curve of parent α-Fe2O3 compound separately in Figure 5c. It can be seen that the maximum value of isothermal magnetization for this sample is roughly 10-fold lower and the coercive field is barely significant compared to the FNN-221 and FNN-211 composites. The low magnetization value has been allied to the weak ferromagnetic exchange coupling in α-Fe2O3 compound. So, in FNN-211 sample because of the lower number of total antiferromagnetic domains (NiO:Ni2FeO4 = 1:1) than the FNN-221 (NiO:Ni2FeO4 = 2:1), the weakly ferromagnetic α-Fe2O3 particles compete with the antiferromagnetic domains. Thus, the ferromagnetic coercivity for this sample is larger than the FNN-221 samples. The schematic diagram representing the possible interactions among these particles has been given in Figure 6. In conclusion, it is deemed important to consider that the local competition between the magnetic states exists because of their contrasting ratios in both of the samples. Therefore, to gain more insight on the exchange coupling at the interface of these magnetic states we measured the MH magnetization with ZFC-FC conditions.
The field dependent ZFC-FC magnetization curves of both the samples are displayed in Figure 7a and the local magnification of these curves covering the coercive field and total magnetization are displayed separately in Figure 7b,c, respectively. We observed that compared to the FNN-221 nanocomposite the hysteresis loop of FNN-211 sample slightly shift towards the high field regions. Indeed, the total ZFC magnetization for FNN-211 sample is slightly lower than the FC magnetization, similarly, the coercivity expanded towards high field values for the same sample. Meanwhile, in distinction between ZFC and FC magnetization in case of the FNN-221 sample is relatively insignificant. The loop shift phenomenon in nanocomposites has been noticed by several authors and attributed to the interfacial exchange interactions (exchange bias) between the two or more magnetic states having contrasting polarities [36,37,38]. The relative shift in the magnetic hysteresis due to the exchange bias can be estimated as H e x = ( H C r i g h t + H C l e f t ) / 2 ; where H e x is the exchange bias field. The H E X value for FNN-211 compound is found to be 35.1 Oe, this value however is relatively smaller than those reported in the literature for α-Fe2O3 and α-Fe2O3: NiO composites [39,40]. The weak exchange bias in our sample is probably due to the existence of an additional magnetic state. Even so, the exchange bias in the second sample (FNN-221) is negligible, which suggests that the exchange bias effect is mainly due to the competing interactions of weak ferromagnetic and antiferromagnetic interfaces. In FNN-221 the antiferromagnetic content is relatively higher than the FNN-211 sample thereby the competition with the weakly ferromagnetic states has been prevailed. This attribution satisfies our previous assumptions and it could be the main reason for comparatively small values of zero field exchange bias in the present work.

4. Conclusions

In summary, we have synthesized ferrite nanocomposites and confirmed the phase fractions of each three α-Fe2O3, NiO and Ni2FeO4 components from XRD and SEM analysis. The temperature dependent and field dependent magnetization of FNN-221 nanocomposite are relatively higher than the FNN-211 sample. The high values of magnetization are because of the increase in interparticle interaction due to incorporation of higher number of NiO particles in this sample. Further in FNN-211 compound the zero field coercivity expands towards the higher field region (hysteresis shift) which points to the existence of exchange bias. The hysteresis loop shift in FNN-221 was completely negligible, which relates that the weak exchange bias effect in FNN-211 sample is certainly due to the local competition between the weak ferromagnetic and antiferromagnetic interfaces.

Author Contributions

Conceptualization, Z.A.; data curation, F.A.; formal analysis, C.A. and A.H.F.; funding acquisition, Z.A.; methodology, Z.A.; resources, Z.A.; visualization, C.A.; writing—original draft, Z.A., C.A., and F.A.; writing—review and editing, C.A., F.A. and A.H.F. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Deanship of Scientific Research at King Faisal University (Saudi Arabia), grant number 1807004 and The APC was funded by the same grant number 1807004.

Acknowledgments

The authors acknowledge the Deanship of Scientific Research at King Faisal University for the financial support under Ra’ed Track (Grant No. 1807004).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Wen, X.; Wang, S.; Ding, Y.; Wang, Z.L.; Yang, S. Controlled Growth of Large-Area, Uniform, Vertically Aligned Arrays of α-Fe2O3Nanobelts and Nanowires. J. Phys. Chem. B 2005, 109, 215–220. [Google Scholar] [CrossRef]
  2. Tadic, M.; Kusigerski, V.; Markovic, D.; Milosevic, I.; Spasojevic, V. High concentration of hematite nanoparticles in a silica ma-trix: Structural and magnetic properties. J. Magn. Magn. Mater. 2009, 321, 12–16. [Google Scholar] [CrossRef]
  3. Singh, H.; Yadav, K. Synthesis and study of structural, dielectric, magnetic and magnetoelectric characterization of BiFeO3–NiFe2O4 nanocomposites prepared by chemical solution method. J. Alloys Compd. 2014, 585, 805–810. [Google Scholar] [CrossRef]
  4. Vuini-Vasi, M.; Antic, B.; Kremenovic, A.; Nikolic, A.S.; Stoiljkovic, M.; Bibic, N.; Spasojevic, V.; Colomban, P. Zn,Ni fer-rite/NiO nanocomposite powder obtained from acetylacetonato complexes. Nanotechnology 2006, 17, 4877–4884. [Google Scholar]
  5. Remya, K.P.; Rajalakshmi, R.; Ponpandian, N. Development of BiFeO3/MnFe2O4 ferrite nanocomposites with enhanced magnetic and electrical properties. Nanoscale Adv. 2020, 2, 2968–2976. [Google Scholar] [CrossRef]
  6. Teja, A.S.; Koh, P.-Y. Synthesis, properties, and applications of magnetic iron oxide nanoparticles. Prog. Cryst. Growth Charact. Mater. 2009, 55, 22–25. [Google Scholar] [CrossRef]
  7. Bansal, M.; Aghamkar, P.; Ahlawat, D.S. Structural properties of CoFe2O4:SiO2 nanocomposites prepared by sol-gel method and co-precipitation method: A comparative study. Int. J. Sci. Technol. Manag. 2015, 4, 871–877. [Google Scholar]
  8. Mélinon, P.; Begin-Colin, S.; Duvail, J.L.; Gauffre, F.; Boime, N.H.; Ledoux, G.; Plain, J.; Reiss, P.; Silly, F.; Warot-Fonrose, B. Engineered inorganic core/shell nanoparticles. Phys. Rep. 2014, 543, 163–197. [Google Scholar] [CrossRef] [Green Version]
  9. Hao, P.; Du, J. Mechanical properties of bio-mimetic energy-absorbing materials under impact loading. J. Mater. Sci. 2018, 53, 3189–3197. [Google Scholar] [CrossRef]
  10. Castro, T.J.; da Silva, S.W.; Nakagomi, F.; Moura, N.S.; Franco, A., Jr.; Morais, P.C. Structural and magnetic properties of ZnO-CoFe2O4 nanocomposites. J. Magn. Magn. Mater. 2015, 389, 27–33. [Google Scholar] [CrossRef]
  11. Gharagozlou, M. Study on the influence of annealing temperature and ferrite content on the structural and magnetic proper-ties of x(NiFe2O4)/(100 − x)SiO2 nanocomposites. J. Alloys Compd. 2010, 495, 217–223. [Google Scholar] [CrossRef]
  12. Kumar, A.; Hussain, I.; Kumar, S.; Koo, B.H. Structural, optical properties and the origin of spin functionality in the Co modi-fied TiO2 nanoparticles. Vacuum 2021, 183, 109870. [Google Scholar] [CrossRef]
  13. Chaudhuri, A.; Mandal, M.; Mandal, K. Preparation and study of NiFe2O4/SiO2 core–shell nanocomposites. J. Alloys Compd. 2009, 487, 698–702. [Google Scholar] [CrossRef]
  14. Praveena, K.; Sadhana, K.; Murthy, S.R. Structural and magnetic properties of NiCuZn ferrite/SiO2 nanocomposites. J. Magn. Magn. Mater. 2011, 323, 2122–2128. [Google Scholar] [CrossRef]
  15. Dippong, T.; Deac, I.G.; Cadar, O.; Levei, E.A.; Diamandescu, L.; Borodi, G. Effect of Zn content on structural, morphological and magnetic behavior of ZnxCo1-xFe2O4/SiO2 nanocomposites. J. Alloys Compd. 2019, 792, 432–443. [Google Scholar] [CrossRef]
  16. Shekofteh-Gohari, M.; Habibi-Yangjeh, A.; Abitorabi, M.; Rouhi, A. Magnetically separable nanocomposites based on ZnO and their applications in photocatalytic processes: A review. Crit. Rev. Environ. Sci. Technol. 2018, 48, 806–857. [Google Scholar] [CrossRef]
  17. Hamed, A.S.; Ali, I.; El Ghazaly, M.; Al-Abyad, M.; Hassan, H. Nanocomposites of ZnO mixed with different Ni-ferrite contents: Structural and magnetic properties. Phys. B Condens. Matter 2021, 607, 412861. [Google Scholar] [CrossRef]
  18. Galizia, P.; Algueró, M.; Bernier, N.; Gambacorti, N.; Aza, E.; Lappas, A.; Venet, M.; Galassi, C. Magnetoelectric dual-particulate composites with wasp-waisted magnetic response for broadband energy harvesting. J. Alloys Compd. 2019, 783, 237–245. [Google Scholar] [CrossRef]
  19. El-Khawas, E.H.; Azab, A.A.; Mansour, A.M. Structural, magnetic and dielectric properties of reduced graphene ox-ide/La0.9Bi0.1FeO3 nanocomposites. Mater. Chem. Phys. 2020, 241, 122335. [Google Scholar] [CrossRef]
  20. Mishra, A.; Sharma, V.; Mohanty, T.; Kuanr, B.K. Microstructural and magnetic properties of rGO/MnFe2O4 nanocomposites; relaxation dynamics. J. Alloys Compd. 2019, 790, 983–991. [Google Scholar] [CrossRef]
  21. Parsons, P.; Duncan, K.; Giri, A.K.; Xiao, J.Q.; Karna, S.P. Electromagnetic properties of NiZn ferrite nanoparticles and their polymer composites. J. Appl. Phys. 2014, 115, 173905. [Google Scholar] [CrossRef]
  22. Huang, X.; Zhang, J.; Wang, W.; Sang, T.; Song, B.; Zhu, H.; Rao, W.; Wong, C. Effect of pH value on electromagnetic loss proper-ties of Co-Zn ferrite prepared via coprecipitation method. J. Magn. Magn. Mater. 2016, 405, 36–41. [Google Scholar] [CrossRef]
  23. Peddis, D.; Cannas, C.; Musinu, A.; Ardu, A.; Orrù, F.; Fiorani, D.; Laureti, S.; Rinaldi, D.; Muscas, G.; Concas, G.; et al. Beyond the Effect of Particle Size: Influence of CoFe2O4 Nanoparticle Arrangements on Magnetic Properties. Chem. Mater. 2013, 25, 2005–2013. [Google Scholar] [CrossRef]
  24. Rinaldi-Montes, N.; Gorria, P.; Martínez-Blanco, D.; Fuertes, A.B.; Puente-Orench, I.; Olivi, L.; Blanco, J.A. Size effects on the Néel temperature of antiferromagnetic NiO nanoparticles. AIP Adv. 2016, 6, 056104. [Google Scholar] [CrossRef] [Green Version]
  25. Tadic, M.; Panjan, M.; Damnjanovic, V.; Milosevic, I. Magnetic properties of hematite (α-Fe2O3) nanoparticles prepared by hy-drothermal synthesis method. Appl. Surf. Sci. 2014, 320, 183–187. [Google Scholar] [CrossRef]
  26. Wang, W.; Ding, Z.; Zhao, X.; Wu, S.; Li, F.; Yue, M.; Liu, J.P. Microstructure and magnetic properties of MFe2O4 (M = Co, Ni, and Mn) ferrite nanocrystals prepared using colloid mill and hydrothermal method. J. Appl. Phys. 2015, 117, 17A328. [Google Scholar] [CrossRef] [Green Version]
  27. Mallesh, S.; Sunny, A.; Vasundhara, M.; Srinivas, V. Structure and magnetic properties of ZnO coated MnZn ferrite nanoparti-cles. J. Magn. Magn. Mater. 2016, 418, 112–117. [Google Scholar] [CrossRef]
  28. Wu, K.; Chang, Y.; Chen, H.; Yang, C.; Horng, D. Variable temperature electron paramagnetic resonance studies of the NiZn ferrite/SiO2 nanocomposite. J. Magn. Magn. Mater. 2004, 278, 156–163. [Google Scholar] [CrossRef]
  29. Mansour, H.; Letifi, H.; Bargougui, R.; De Almeida-Didry, S.; Negulescu, B.; Autret-Lambert, C.; Gadri, A.; Ammar, S. Structural, optical, magnetic and electrical properties of hematite (α-Fe2O3) nanoparticles synthesized by two methods: Polyol and precipitation. Appl. Phys. A 2017, 123, 787. [Google Scholar] [CrossRef]
  30. Mrabet, C.; Ben Amor, M.; Boukhachem, A.; Amlouk, M.; Manoubi, T. Physical properties of La-doped NiO sprayed thin films for optoelectronic and sensor applications. Ceram. Int. 2016, 42, 5963–5978. [Google Scholar] [CrossRef]
  31. Ortiz-Quiñonez, J.L.; Pal, U.; Villanueva, M.S. Structural, Magnetic, and Catalytic Evaluation of Spinel Co, Ni, and Co–Ni Ferrite Nanoparticles Fabricated by Low-Temperature Solution Combustion Process. ACS Omega 2018, 3, 14986–15001. [Google Scholar] [CrossRef] [PubMed]
  32. Bruvera1, I.J.; Zélis, P.M.; Calatayud, M.P.; Goya, G.F.; Sánchez, F.H. Determination of the blocking tempera-ture of magnetic nanoparticles: The good, the bad, and the ugly. J. Appl. Phys. 2015, 118, 184304. [Google Scholar] [CrossRef]
  33. Bhowmik, R.N.; Saravanan, A. Surface magnetism, Morin transition, and magnetic dynamics in antiferromagnetic α-Fe2O3 (hematite) nanograins. J. Appl. Phys. 2010, 107, 053916. [Google Scholar] [CrossRef]
  34. Mitra, S.; Das, S.; Basu, S.; Sahu, P.; Mandal, K. Shape- and field-dependent Morin transitions in structured α-Fe2O3. J. Magn. Magn. Mater. 2009, 321, 2925–2931. [Google Scholar] [CrossRef]
  35. Szafraniak-Wiza, I.; Andrzejewski, B.; Hilczer, B. Magnetic Properties of Bismuth Ferrite Nanopowder Obtained by Mechanochemical Synthesis. Acta Phys. Pol. A 2014, 126, 1029–1031. [Google Scholar] [CrossRef]
  36. Rana, R.; Pandey, P.; Singh, R.P.; Rana, D.S. Positive exchange-bias and giant vertical hysteretic shift in La0.3Sr0.7FeO3/SrRuO3 bilayers. Sci. Rep. 2014, 4, 4138. [Google Scholar] [CrossRef] [PubMed]
  37. Testa-Anta, M.; Murias, B.R.; Salgueiriño, V. Spin Frustration Drives Exchange Bias Sign Crossover in CoFe2O4-Cr2O3 Nanocom-posites. Adv. Funct. Mater. 2019, 29, 1900030. [Google Scholar] [CrossRef]
  38. Artus, M.; Ammar, S.; Sicard, L.; Piquemal, J.-Y.; Herbst, F.; Vaulay, M.-J.; Fiévet, F.; Richard, V. Synthesis and Magnetic Properties of Ferrimagnetic CoFe2O4 Nanoparticles Embedded in an Antiferromagnetic NiO Matrix. Chem. Mater. 2008, 20, 4861–4872. [Google Scholar] [CrossRef]
  39. Liu, X.H.; Cui, W.B.; Lv, X.K.; Liu, W.; Zhao, X.G.; Li, D.; Zhang, Z.D. Exchange bias and phase transformation in α-Fe2O3+NiO nanocomposites. J. Appl. Phys. 2008, 103, 103906. [Google Scholar] [CrossRef]
  40. He, X.-M.; Zhang, C.-W.; Guo, F.-F.; Yan, S.-M.; Li, Y.-T.; Liu, L.-Q.; Zhang, H.-G.; Du, Y.-W.; Zhong, W. Exchange-biased hybrid γ-Fe2O3/NiO core-shell nanostructures: Three-step synthesis, microstructure, and magnetic properties. Phys. Chem. Chem. Phys. 2019, 21, 11967–11976. [Google Scholar] [CrossRef]
Figure 1. (a,b) XRD profiles of FNN-221 and FNN-211 sample, respectively, phase fraction chart has been obtained through the Rietveld method in HighScore plus.
Figure 1. (a,b) XRD profiles of FNN-221 and FNN-211 sample, respectively, phase fraction chart has been obtained through the Rietveld method in HighScore plus.
Crystals 11 00613 g001
Figure 2. FE-SEM micrographs of (a) FNN-221 sample and the left side image (b) SEM image of FNN-211 compound, (cn) corresponding EDX mapping of the site indicating the distribution of Au, C, Fe, O, Ni elements, respectively, for both the samples.
Figure 2. FE-SEM micrographs of (a) FNN-221 sample and the left side image (b) SEM image of FNN-211 compound, (cn) corresponding EDX mapping of the site indicating the distribution of Au, C, Fe, O, Ni elements, respectively, for both the samples.
Crystals 11 00613 g002
Figure 3. The deconvoluted Raman peaks corresponding to the (a) FNN-221 and (b) FNN-211 samples, respectively.
Figure 3. The deconvoluted Raman peaks corresponding to the (a) FNN-221 and (b) FNN-211 samples, respectively.
Crystals 11 00613 g003
Figure 4. Temperature dependent ZFC-FC magnetization of (a) FNN-221 and (b) FNN-211 (c) α-Fe2O3 compounds.
Figure 4. Temperature dependent ZFC-FC magnetization of (a) FNN-221 and (b) FNN-211 (c) α-Fe2O3 compounds.
Crystals 11 00613 g004
Figure 5. Comparison of different figure of merits at 50 K, 100 K, 200 K and 300 K isotherms for (a) FNN-221 and (b) FNN-211 (c) α-Fe2O3. (a1,a2,b1,b2,c1,c2) are the corresponding local magnified image of magnetization curves representing the behavior of total magnetization and coercivity with respect to the temperature.
Figure 5. Comparison of different figure of merits at 50 K, 100 K, 200 K and 300 K isotherms for (a) FNN-221 and (b) FNN-211 (c) α-Fe2O3. (a1,a2,b1,b2,c1,c2) are the corresponding local magnified image of magnetization curves representing the behavior of total magnetization and coercivity with respect to the temperature.
Crystals 11 00613 g005aCrystals 11 00613 g005b
Figure 6. Schematic illustration of spin interactions that may have taken place inside the magnetic domains of the different components consisting variable spin functionalities.
Figure 6. Schematic illustration of spin interactions that may have taken place inside the magnetic domains of the different components consisting variable spin functionalities.
Crystals 11 00613 g006
Figure 7. (a) Low temperature (50 K) FC-ZFC magnetization behavior of FNN-221 and FNN-211 samples under a variable applied field. (b) Local magnification of the MH curve indicating the difference between the ZFC and FC magnetization of both samples. (c) Shift in the hysteresis loop of FNN-211 sample in comparison to FNN-221 sample, signifying the exchange bias field (Hex). (d) Comparison of Low temperature (50 K) magnetization for α-Fe2O3 compound.
Figure 7. (a) Low temperature (50 K) FC-ZFC magnetization behavior of FNN-221 and FNN-211 samples under a variable applied field. (b) Local magnification of the MH curve indicating the difference between the ZFC and FC magnetization of both samples. (c) Shift in the hysteresis loop of FNN-211 sample in comparison to FNN-221 sample, signifying the exchange bias field (Hex). (d) Comparison of Low temperature (50 K) magnetization for α-Fe2O3 compound.
Crystals 11 00613 g007
Table 1. Amount reagents and solutions taken during the synthesis process; ratio of individual component taken in each sample.
Table 1. Amount reagents and solutions taken during the synthesis process; ratio of individual component taken in each sample.
ReagentsFe2O3NiONi2FeO4
FeCl3.6H2O0.25 M0.2 M
NiCl2.6H2O0.4 M
Ni(NO3)2.6H2O0.5 M3 M
NaOH0.5 M
NH4OH5 mL
D-I water (mL)150150150
Nanocomposites
Sample labelFe2O3 (mol)NiO (mol)Ni2FeO4 (mol)
FNN-221221
FNN-211211
Table 2. The phase fraction (in molar) of components and concentration of each element present in the sample (obtained from XRD and EDX analysis, respectively); comparison of different figure of merits acquired from the magnetization measurements. For each component the thermal parameters were kept constant while performing the calculations.
Table 2. The phase fraction (in molar) of components and concentration of each element present in the sample (obtained from XRD and EDX analysis, respectively); comparison of different figure of merits acquired from the magnetization measurements. For each component the thermal parameters were kept constant while performing the calculations.
ParametersFNN-221FNN-211
Unit cell parameters (Å) Fe2O3 (R-3c): a = 5.034 ± 0.00045; c = 13.7368 ± 0.00178; Biso O1 = 0.50000; Biso Fe1 = 0.50000Fe2O3 (R-3c): a = 5.029 ± 0.00030; c = 13.7300 ± 0.00119; Biso O1 = 0.150000; Biso Fe1 = 0.10000
NiO (Fm-3m): a = b = c = 4.1760 ± 0.00012; Biso O1 = 0.50000; Biso Ni1 = 0.50000NiO (Fm-3m): a = b = c = 4.1742 ± 0.00012; Biso O1 = 0.500000; Biso Ni1 = 0.50000
Ni2FeO4 (Fd-3m): a = b = c = 8.3275 ± 0.00199; Biso O1 = 0.50000; Biso Fe1 = 0.50000; Biso Ni1 = 0.50000Ni2FeO4 (Fd-3m): a = b = c = 8.3428 ± 0.00210; Biso O1 = 0.500000; Biso Fe1 = 0.50000; Biso Ni1 = 0.50000
Rexp 12.4914.28
Rp 18.8217.85
Rwp 17.9317.39
χ22.061.84
Nanocomposites
FNN-221FNN-211
Phase fraction (Rietveld refinement) Fe2O3: 41.2% ± 0.39; NiO: 39.7% ± 0.14; Ni2FeO4: 19.1% ± 0.51Fe2O3: 49.5% ± 0.62; NiO: 26.4% ± 0.17; Ni2FeO4: 24.1% ± 0.33
crystallite size (nm) Scherrer’s formula Fe2O3: 58.4 NiO: 292.1 Ni2FeO4: 51.6
crystallite size (nm) Rietveld method Fe2O3: 49 NiO: 286.3 Ni2FeO4: 48.4
Element analysis (EDX) Crystals 11 00613 i001 Crystals 11 00613 i002
Magnetization figures of merit (at 50K-FC), Mt/Mr (emu/g), HC (Oe) Mt = 14.8 (at 3T), Mr: 4.83, HC = 428.1 Mt = 12.1 (at 3T), Mr: 4.01, HC = 440.72
Exchange Bias field (Hex) (Oe) 35.1
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Alhashem, Z.; Awada, C.; Ahmed, F.; Farha, A.H. Structural and Magnetic Properties Study of Fe2O3/NiO/Ni2FeO4 Nanocomposites. Crystals 2021, 11, 613. https://doi.org/10.3390/cryst11060613

AMA Style

Alhashem Z, Awada C, Ahmed F, Farha AH. Structural and Magnetic Properties Study of Fe2O3/NiO/Ni2FeO4 Nanocomposites. Crystals. 2021; 11(6):613. https://doi.org/10.3390/cryst11060613

Chicago/Turabian Style

Alhashem, Zakia, Chawki Awada, Faheem Ahmed, and Ashraf H. Farha. 2021. "Structural and Magnetic Properties Study of Fe2O3/NiO/Ni2FeO4 Nanocomposites" Crystals 11, no. 6: 613. https://doi.org/10.3390/cryst11060613

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