Next Article in Journal
Influence of Chloride Concentration on Fretting Wear Behavior of Inconel 600 Alloy
Previous Article in Journal
The Impact of Abiotic and Biotic Conditions for Degradation Behaviors of Common Biodegradable Products in Stabilized Composts
Previous Article in Special Issue
Efficient Chlorostannate Modification of Magnetite Nanoparticles for Their Biofunctionalization
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Preparation and Characterization of Zinc Ferrite and Gadolinium Iron Garnet Composite for Biomagnetic Applications

1
i3N and Department of Physics, University of Aveiro, Campus Universitário de Santiago, 3810-193 Aveiro, Portugal
2
i3N/CENIMAT, Physics Department, NOVA School of Science and Technology, Campus de Caparica, NOVA University Lisbon, 2829-516 Caparica, Portugal
3
i3N/CENIMAT, Science Materials Department, Faculty of Sciences and Technology, Nova University of Lisbon, 2829-516 Caparica, Portugal
*
Authors to whom correspondence should be addressed.
Materials 2024, 17(12), 2949; https://doi.org/10.3390/ma17122949
Submission received: 30 April 2024 / Revised: 4 June 2024 / Accepted: 11 June 2024 / Published: 17 June 2024

Abstract

:
Cancer is a major worldwide public health problem. Although there have already been astonishing advances in cancer diagnosis and treatment, the scientific community continues to make huge efforts to develop new methods to treat cancer. The main objective of this work is to prepare, using a green sol–gel method with coconut water powder (CWP), a new nanocomposite with a mixture of Gd3Fe5O12 and ZnFe2O4, which has never been synthesized previously. Therefore, we carried out a structural (DTA-TG and X-ray diffraction), morphological (SEM), and magnetic (VSM and hyperthermia) characterization of the prepared samples. The prepared nanocomposite denoted a saturation magnetization of 11.56 emu/g at room temperature with a ferromagnetic behavior and with a specific absorption rate (SAR) value of 0.5 ± 0.2 (W/g). Regarding cytotoxicity, for concentrations < 10 mg/mL, it does not appear to be toxic. Although the obtained results were interesting, the high particle size was identified as a problem for the use of this nanocomposite.

Graphical Abstract

1. Introduction

According to the World Health Organization, there were approximately 20 million new cases and 10 million deaths in 2020 worldwide due to cancer [1]. In 2040, it is estimated that there will be nearly 30 million diagnoses and 16 million deaths from cancer worldwide [2]. The evolution of science has led to a better understanding of the mechanisms behind cancer pathophysiology. Despite the visible decline in mortality, the increasing incidence of cancer justifies the continuous development of new treatment methods for the different types of cancer [3].
Since Richard Feynman’s lecture “There’s Plenty of Room at the Bottom” (1959), the advances of nanotechnology towards biomedical applications have undergone a major revolution in the diagnosis and treatment of cancer [4]. The combination of therapeutic and diagnostic (theranostics) in the same nanoparticle is an emerging tool that intends to act as a precise and personalized approach to cancer treatment [5]. However, the lack of knowledge on certain issues, such as the biological response to nanoparticles and their elimination in an organism, limits their use in clinical applications [5].
Magnetic nanoparticles (MNPs) are defined as nanostructures that possess at least one dimension on the nanoscale, with remarkable magnetic properties [4]. Their high surface-to-volume ratio [6] is responsible for their outstanding magnetic properties. In accordance with B. Issa et al. [7,8], the primary characteristics that they must have to be used for biomedical applications are biocompatibility and non-toxicity; a particle size distribution between 10 and 200 nm [9]; a high saturation magnetization (Ms) to provide easy control of the particles in the blood through a moderate external magnetic field; and the possibility of better targeting of the pathologic tissue. By controlling the size, material, and coating of the MNPs, it is possible to improve and modify their properties considering the biomedical application in question.
In magnetic hyperthermia (MH) therapy, the magnetic heating efficiency of MNPs is a parameter that is extremely relevant. To evaluate their performance, specific parameters are usually used, such as specific heat absorption rate (SAR). SAR is defined as the capacity that a certain magnetic material has to generate heat. It is used to characterize the efficiency of heating a magnetic material through its absorption of energy during the exposition to an alternating magnetic field (AMF). In recent years, most of the studies have been focused on improving the SAR value of the MNPs for MH therapy use. Different methods have been used, such as controlling the particle core size, shape, composition, and surface shell, and selecting specific magnetic materials [10]. According to Liu et al. [11], due to the correlation between the decreasing of Ms as the size decreases, and knowing that SAR is proportional to Ms, it is possible to conclude that SAR value increases with the increasing size of MNPs. Thus, MNPs with a higher Ms are desirable for effective magnetic loss [11].
Gadolinium nanoparticles (GdNPs) were first presented as a Magnetic Resonance Imaging (MRI) contrast agent by Carr et al. in 1984 [12]. Since then, MRI as well as other methods of medical imaging commonly employ this metal lanthanide as a contrast agent [12,13]. Due to its seven unpaired electrons and delayed electronic relaxation, the trivalent cation is regarded as a hard acid and is therefore employed in MRI [12]. In addition to its use in MRI, this element has drawn interest for various biomedical uses, most notably magnetic hyperthermia [14,15]. Jiang P. et al. [16] synthesized gadolinium-doped iron oxide nanoparticles with an elevated SAR value.
Spinel ferrite nanoparticles, such as MFe2O4 (where M = divalent metal ions, such as Co2+, Ni2+, Zn2+, Mn2+), are commonly used as MNPs for MH due to their astonishing chemical and physical properties [16,17]. Very good chemical stability, enhanced saturation magnetization, and high electrical resistivity are some of the outstanding properties that characterize them [17]. Zinc ferrite (ZnFe2O4) is a material that shows a lot of potential to be used in the biomedical field [18]. Bulk ZnFe2O4 has a normal spinel structure in which Zn2+ cations occupy the tetrahedral positions, while the octahedral positions are occupied by Fe3+ cations. However, when the dimensions are reduced to the nanoscale, it is described that the Zn2+ cations are distributed both on tetrahedral and octahedral sites, leading to their partially inverse spinel structure [19,20,21]. Among its different applications, zinc ferrite is seen as “the potential alternative material” [21], regarding its properties. In terms of biomedical applications, due to the lack of knowledge on the biological response of the organism and toxicity of zinc ferrite nanoparticles, there are only a few available clinical studies where only zinc ferrite was explored for MH. Somvanshi S. et al. [18] synthesized zinc ferrite nanoparticles using chemical coprecipitation and further functionalization with oleic acid. The obtained results showed superparamagnetic behavior, with a Ms of 25.5 emu/g. According to these outcomes, it is possible to conclude that zinc ferrite has the potential to be employed as a material for cancer treatment through MH.
Being aware of the advantages of green synthesis methods, new methods have been described by the scientific community as promising, cheaper, and sustainable ways to produce nanoparticles. Thus, based on its basic features and rich composition, the coconut water powder (CWP)-assisted sol–gel method has been described as a successful method to produce different MNPs. Some of the examples described in the literature are the synthesis of LiFe5O8 [22], Y2O3:Eu3+ [23], Y2O3:Nd3+ [24], niobium oxides [25], SrFe12O19 [26], and BaFe12O19 [27]. The main advantages of using this proteic sol–gel method based on coconut water are (1) low-cost method; (2) higher concentration of sugars, promoting the gelation process; (3) material widely and easily available around the world, being available at an industrial scale; (4) the promotion of a homogeneous distribution of the precursor ions due to the presence of proteins and amino acids that have the ability to bind with metal ions; and (5) importance as a renewable source [25].
The main purpose of this work is to develop and characterize structurally, morphologically, and magnetically a new nanocomposite of zinc ferrite (ZnFe2O4) and gadolinium iron garnet (Gd3Fe5O4). The aim is to synthesize this nanocomposite through a green synthesis sol–gel method using coconut water powder. This nanocomposite is intended to be used for MH treatment. Therefore, this nanocomposite’s SAR and biocompatibility are evaluated.

2. Materials and Methods

2.1. Nanocomposite Synthesis

The coconut water powder (CWP)-assisted sol–gel method was used to prepare an efficient and novel nanocomposite of Gd3Fe5O12 and ZnFe2O4 able to be used for magnetic hyperthermia, as described by Teixeira S. et al. [22]. For Gd3Fe5O4, we used iron (III) nitrate (Fe(NO3)3·9H2O) (Merck KGaA, Darmstadt, Germany, ≥99.0%) and gadolinium nitrate hydrate (Gd(NO3)3·H2O) (Merck KGaA, Darmstadt, Germany, ≥99.0%), and for ZnFe2O4, we used iron (III) nitrate (Fe(NO3)3·9H2O) and zinc nitrate hexahydrate (Zn(NO3)2·6H2O) (Merck KGaA, Darmstadt, Germany, ≥99.9%) as precursors, respectively.
First, each compound was synthesized individually. Then, after ensuring the best phase of each ferrite, Gd3Fe5O12 and ZnFe2O4 powders were mixed with a planetary ball mill to obtain the nanocomposite. For both Gd3Fe5O12 and ZnFe2O4, we applied the following steps (Figure 1):
(1)
The metal precursors were dissolved into a CWP solution with a concentration of 0.58 mol/dm3, the critical micelle concentration (whose determination is explained in Section 3), and mixed with a magnetic stirrer in two steps:
(a)
At T = 80 °C, for 2 h;
(b)
At T = 100 °C, for 2 h.
(2)
The obtained viscous brown gel was heat-treated at 350 °C for 1 h to remove the solvent;
(3)
Preparation of pellets of Gd3Fe5O12 and ZnFe2O4 powders 10 mm in diameter and with a thickness of 2 mm, approximately, with three tons applied;
(4)
The pellets were then submitted to different heat treatments using different values of dwell time and heating rate. The cooling process was performed according to the furnace’s thermal inertia when the power was switched off;
(5)
After a structural analysis with X-ray diffraction, the purest phase of both Gd3Fe5O12 and ZnFe2O4 was chosen. To perform this analysis, the pellets were ground with the help of a mortar and pestle;
(6)
The best phase of each compound was mixed with a planetary ball mill, the Pulverisette 7. The planetary ball mill was used due to its high efficiency in the process of mixing different materials [28].
Finally, the composite was obtained and characterized structurally, morphologically, magnetically, and biologically.

2.2. Structural and Morphological Characterization

A structural and morphological analysis was performed to characterize structurally and morphologically the samples. The differential thermal analysis (DTA) and thermogravimetric analysis (TG) were conducted with Hitachi STA7300 equipment (Tokyo, Japan) in a nitrogen atmosphere with a flux of 200 mL. The analysis was performed with a heating rate of 5 °C/min in a range of temperatures from room temperature up to 1400 °C.
To characterize the crystalline structure of the samples, X-ray diffraction (XRD) was used. Therefore, Panalytical AERIS equipment from Malvern Panalytical (Westborough, MA, USA) with CuKα radiation with a 2θ angle (10–60°), a wavelength of 1.54060 Å, and at 40 kV and 15 mA was used. The XRD was performed using the powders obtained from the crushed pellets for each sample. The crystalline phases were identified with X’Pert HighScore Panalytical software version 5.2, which contains the database of the Joint Committee for Powder Diffraction Standards–International Center for Diffraction Data (JCPDS).
Scanning electronic microscopy (SEM) is a widely used technique that enables us to analyze the microstructure morphology. Thus, the TESCAN Vega3 SB (Warrendale, PA, USA) was used with an accelerating beam voltage of 30 kV. Carbon deposition was performed to ensure the samples’ conductivity. Using ImageJ 1.52v, the average grain size was determined.

2.3. Magnetic Characterization

A vibrating sample magnetometer (VSM), the Cryofree model from Cryogenic (London, UK), was used to measure the magnetization (M) of the samples vs. the magnetic field (H) up to 50 kOe. The values of M and H were measured in a range of temperature from −258.15 to 26.85 °C.
The magnetic hyperthermia assays were performed using a D5 series from the nB nanoscale Biomagnetics (Zaragoza, Spain). The samples were exposed to an alternating current magnetic field of 24 kAm−1, with a frequency of 388 Hz for 10 min. Each sample was immersed in 1 mL of ultra-pure water and was ultra-sonicated before each measurement.
In MH therapy, the specific heat absorption rate parameter (SAR) is used to evaluate the magnetic heating efficiency of MNPs. SAR is calculated using the following equation:
S A R = m l × c l + m F e × c N P m l + m F e × d T d t
where d T d t is given by the variation of temperature within a certain period of time, c l is the specific heat of the liquid, c N P is the specific heat of the magnetic material, m l is the fluid mass, and m F e is the mass concentration of the magnetic element in the solution [8].

2.4. Cytotoxicity Analysis

The cytotoxicity assays were performed according to the ISO 10993-5:2009 standard “Biological evaluation of medical devices—Part 5: Tests for in vitro cytotoxicity” [29]. Due to the presence of gadolinium in human bones and its importance in bone regeneration [30], the SaOs-2 cell line (human osteosarcoma, ATCC HTB-85) was chosen as a cellular model of cells from bone.
The extracts were produced by placing 20 mg of each of the samples in 1 mL of complete culture medium (McCoy’s 5 A from Sigma-Aldrich (St. Louis, MO, USA), catalog number M4892; supplemented with 2.2 g/L sodium bicarbonate, Sigma-Aldrich, catalog number S5761; 1% penicillin/streptomycin from Gibco (Miami, FL, USA)/ThermoFisher (Waltham, MA, USA), catalog number 15140122; and 10% FBS, Fetal Bovine Serum, from Biowest, Nuaillé, France, catalog number S1810) at a temperature of 37 °C for 48 h.
Cells were seeded at a density of 30,000 cells/cm2 in 96-well plates and were incubated at 37 °C in a 5% CO2 Sanyo MCO19AICUV incubator for 24 h. The extracts were used at the initial concentration of 20 mg/mL and were also diluted to obtain equivalent extract concentrations of 10, 5, 2.5, and 1.125 mg/mL. Each of the concentrations was tested four times. Two controls were set up: a negative control, where cells were cultured with a complete culture medium, and a positive control, where the culture medium was supplemented with 10% dimethyl sulfoxide, a cytotoxic compound. After 48 h of incubation, the cell culture media were aspirated and replaced by a medium containing 50% of the complete culture medium and 50% of a 0.04 mg/mL resazurin solution prepared using a phosphate-buffered saline (PBS) solution. After 3 h of incubation at 37 °C and 5% CO2, the absorbance of each well was measured at 570 and 600 nm, which correspond to the absorbance maxima of resorufin and resazurin, respectively. Metabolically active cells reduce resazurin to resorufin and the conversion rate is assumed to be proportional to the cell population. The OriginPro 2018 software was used to perform analysis of variance (ANOVA) to determine the significance of differences between samples for each concentration. Tukey’s test was used for multiple comparisons, and the differences were assumed to be statistically significant if p < 0.05. Cell viability is given as a percentage of viable cells in the samples to test relative to the negative control:
% c e l l   v i a b i l i t y = t r e a t e d   c e l l s c o n t r o l   c e l l s × 100

3. Results and Discussion

3.1. Critical Micelle Concentration (CMC)

The critical micelle concentration (CMC) is defined as a phenomenon that separates two distinct behaviors of the size distribution of micelles [31]. The surfactants are characterized as amphiphilic molecules that comprise two different parts, polar and hydrophilic [31]. In water, these molecules present a specific organized molecule configuration, defined as micelles. Nevertheless, this phenomenon is only verified for concentrations above the CMC. It is represented by an inflection point that can be calculated through the increase in the concentration of amphiphilic molecules, which results in changes in the physicochemical properties of the surfactant solution [32]. Therefore, with the selected technique, the synthesis of the nanocomposite can only be verified if the concentration of the CWP in the solution has a CMC value. As seen in Figure 2, it is possible to analyze the variation of conductivity as a function of frequency for the different concentrations of CWP under study. Figure 2a shows the variation of the conductivity for the different concentrations of CWP, which goes from 0.1 mol/dm3 to 0.9 mol/dm3. For a comparative study, 10 5 Hz was chosen as the reference frequency. Figure 2b represents the variation of conductivity as a function of the concentration of the CWP solutions, for a frequency of 10 5 Hz. The inflection point, in this case, is the intersection point of the two lines resulting from the linear fittings performed, representing the CMC value. Thus, a value of 0.58 mol/dm3 was obtained.

3.2. Thermal Analysis

Figure 3 shows the results of the differential thermal analysis (DTA) and thermogravimetric analysis (TG) performed on the Gd3Fe5O12 and ZnFe2O4 powders after the first heat treatment at 350 °C. In these thermograms, the exothermic peaks without associated weight loss are commonly related to phase crystallization.
In Figure 3a, thermal analysis of Gd3Fe5O12 sample, it is possible to highlight three exothermic peaks at 705, 988, and 1223 °C without associated weight loss. Thus, suggesting that for the indicated temperatures, it is possible for meaningful structural variation to occur. Regarding the ZnFe2O4 sample, (Figure 3b), five exothermal peaks without any loss of mass can be verified, making them relevant to analyze. The peaks that suggest the existence of structural changes are centered at 421, 668, 845, 1060, and 1220 °C. Thus, the heat treatments that will be applied to the Gd3Fe5O12 and ZnFe2O4 samples will be for temperatures of 500, 700, 850, 1000, 1200, and 1400 °C.

3.3. Morphological and Structural Characterization

Figure 4 illustrates the diffractograms taken for the different Gd3Fe5O12 samples. In Figure 4, we only presented the phases of HT at 1200 and 1400 °C since they were the only ones with a composition of Gd3Fe5O12. Other than that, different times of HT (4 h and 24 h) were chosen to analyze the impact of the heating rate on the samples’ composition. It is possible to conclude that none of the samples had a pure Gd3Fe5O12 composition. Fe2O3, Fe3O4, and Gd3Fe5O12 phases were identified. The sample with the highest Gd3Fe5O12 composition was heat-treated at 1200 °C for 24 h. The sample heat-treated at 1200 °C for 4 h is not presented because it has secondary crystal phases (GdFeO3, Fe2O3, and Fe3O4) and a high crystallite size.
A Rietveld refinement was applied to the sample with the highest Gd3Fe5O12 composition (HT at 1200 °C for 24 h). The Goodness of Fit (GoF) and the ratio of Rwp to Rexp are two parameters that, according to the literature, are used to examine the quality of Rietveld refinement [33,34]. By looking at the parameters presented in Figure 5, it is possible to confirm that Rwp > Rexp and that the GoF is ≈2, which demonstrates a good refinement quality.
Figure 6 illustrates the XRD of the ZnFe2O4. It should be noted that none of the synthesized samples show a pure composition of ZnFe2O4 since different phases of ZnO and even Fe2O3 have been identified. However, the sample that exhibits a higher ZnFe2O4 composition (97%) is the sample heat-treated at 1200 °C for 4 h. Figure 7 shows the Rietveld refinement fit, where the Rwp parameter value is higher than the Rexp and has a GoF of 2.0051.
In comparative terms, after performing a ball-milling procedure to decrease the particle size, an X-ray diffractogram was conducted on both samples with the highest composition of Gd3Fe5O12 (HT 1200 °C for 24 h) and ZnFe2O4 (1200 °C for 4 h). Observing Figure 8, it can be noticed that in the ZnFe2O4 sample, a SiO2 contamination is verified, probably caused by the agate grinding balls used for ball milling [35]. One explanation for this contamination not occurring in the Gd3Fe5O12 sample may be due to the intensity of the peaks, which are much more intense in the case of the Gd3Fe5O12 sample. Also, both diffractograms reveal some amorphous phase due to the presence of SiO2.
The phases with the highest composition of Gd3Fe5O12 (HT 1200 °C for 24 h) and ZnFe2O4 (1200 °C for 4 h) were chosen to synthesize the composite. The structural analysis performed on the composite, illustrated in Figure 9, shows that four different phases are present, in agreement with the results obtained previously: Gd3Fe5O12, ZnFe2O4, Fe2O3, and SiO2 (from the agate ball contamination). The composite presents a composition of 65% Gd3Fe5O12 and 10% ZnFe2O4. The Rietveld refinement obtained for the nanocomposite showed a GoF of 1.78 and a higher Rwp than Rexp.
Regarding the morphological analysis performed on the samples, Figure 10 illustrates the SEM analysis performed on the Gd3Fe5O12 (Figure 10a) and ZnFe2O4 (Figure 10b) samples after the heat treatment; Figure 10c,d represents the samples after the first ball-milling cycle performed on the Gd3Fe5O12 (Figure 10c) and ZnFe2O4 (Figure 10d) samples; and Figure 10e,f shows the results obtained at the end of the ball-milling process of the Gd3Fe5O12 (Figure 10e) and ZnFe2O4 (Figure 10f) samples.
Figure 10 shows the Gd3Fe5O12 particles after heat treatment, with large aggregates of circular structures, as described by Jiang L. et al. [34]. The average particle size determined by ImageJ 1.52v software is 1.09 μm. The SEM images obtained for the ZnFe2O4 sample are significantly irregular, with large agglomerates. The average size of the synthesized particles is approximately 1.59 μm. Firstly, these micrographs were obtained with pellets of the samples. Only after the morphological analysis were the pellets crushed.
In comparative terms, after conducting the first ball-milling cycle, each sample was analyzed to verify the size and uniformity. In both Figure 10c,d, it is possible to conclude that both samples are significantly irregular, especially in terms of grain size uniformity. The average grain size is 0.92 μm for the Gd3Fe5O12 sample (Figure 10c) and 0.51 μm for the ZnFe2O4 sample (Figure 10d). Due to their irregularity in terms of grain size distribution, it was decided that it was necessary to execute one more cycle of ball milling.
Therefore, micrographs in Figure 10e,f were acquired after the total ball-milling process for Gd3Fe5O12 and ZnFe2O4, respectively. The results allow us to emphasize the presence of agglomerates in both samples. The Gd3Fe5O12 particles, as seen in Figure 10e, present an average particle size of 273 nm, and the ZnFe2O4 samples have an average size of 196 nm. These two samples were used for the synthesis of the nanocomposite, with the help of a planetary ball milling, to make their mixing possible.
The morphology of the synthesized composite is shown in Figure 11, where the presence of large agglomerates is clearly visible. The average particle size obtained for the composite was 155 nm.

3.4. Magnetic Characterization

Figure 12a represents the hysteresis M–H curve for Gd3Fe5O12 sample at temperatures of −258.15 and 26.85 °C, i.e., room temperature. The sample tested at −258.15 °C has a saturation magnetization of 57.03 emu/g, denoting a dominant ferromagnetic behavior, as represented in Figure 12a. This behavior is in accordance with what is reported in the literature [36]. The sample tested at 26.85 °C, on the other hand, does not enable us to calculate the value of saturation magnetization since its magnetization rises with increasing values of the magnetic field, demonstrating no magnetic hysteresis. The magnetization’s tendency to increase linearly with the magnetic field can be attributed to a typical behavior of paramagnetic nanoparticles (Figure 12a) [34]. Additionally, analyzing the inverse of susceptibility with the variation of temperature (top inset of Figure 12a), above −272.15 °C, the sample demonstrates a variation from a ferromagnetic to paramagnetic behavior, corroborating the previous analysis. This temperature-dependent change in the magnetic behavior is a direct consequence of the paramagnetic contribution associated with Gd3+ and the antiferromagnetic contribution from the octahedral structure of the Fe sublattice [37]. Due to the magnetic unit cell’s ordered spins in the spin-canted FeO6 octahedra structure, the antiferromagnetic contribution was prominent at low temperatures. The ferromagnetism gradually diminished as the surrounding temperature increased and the paramagnetic contribution of Gd3+ ions took control [34].
The hysteresis curve for the ZnFe2O4 sample is shown in Figure 12b. First, it is clear that the nanoparticles have a larger saturation magnetization at −258.15 °C, with a value of roughly 37.16 emu/g, compared to the 23.67 emu/g obtained for a temperature of 26.85 °C. However, both samples exhibit ferromagnetic behavior in both circumstances. According to the literature [37], it is described that particles with a diameter between 29 and 35 nm, at room temperature present ferromagnetic behavior with clear evidence of a hysteresis curve. In the first instance, in terms of the comparison of the magnetic behavior of this present work, despite the difference in the particle size, its behavior is similar to what is described in the literature [37]. Regarding the values of saturation magnetization obtained for the synthesized particles, compared to that described in the literature [37], approximately 10 emu/g, it assumes a significantly higher value, which is noted to be quite favorable.
The examination of the M–H curve of the composite, as seen in Figure 13, obtained by the mixture of Gd3Fe5O12 and ZnFe2O4, allows us to infer that, as previously proven, the saturation magnetization has greater values at −258.15 °C, according to what was expected. Secondly, in terms of magnetic behavior, it is noted that due to the presence of ZnFe2O4 in the composite, the particles behave ferromagnetically at both analysis temperatures. While the saturation magnetization for −258.15 °C was 57.30 emu/g, the result for 26.85 °C was 11.56 emu/g. The low saturation magnetization and ferromagnetic behavior of the synthesized composite at room temperature can be explained by the constitution of the sample (XRD analysis (Figure 9)). At room temperature, the paramagnetic behavior from Gd3Fe5O12 with the dominant ferromagnetism behavior from ZnFe2O4 and the antiferromagnetism from Fe2O3 [38] result in a magnetization with values lower than the obtained for ZnFe2O4 but slightly higher than the ones obtained for Gd3Fe5O12.
Individually, the SAR obtained for the synthesized Gd3Fe5O12 sample was 0.3 ± 0.2   ( W · g 1 ) . By itself, the SAR value obtained is lower than the ambitioned. The size of the nanoparticles, their state of aggregation, and their composition are some of the variables that might be at the root of this considerable variation. The SAR value obtained for the produced ZnFe2O4 nanoparticles was approximately 0 (W/g). The low SAR value obtained, compared to the literature [39], may be due to the agglomeration of the particles, resulting in the increase in the average particle size (Figure 11) compared with the ones referred to in the literature [39]. Due to their limited thermal efficiency, these particles do not appear to be feasible for use in magnetic hyperthermia on their own according to the data obtained. However, their low SAR value may be due to their high particle size, and they may have greater values if a size decrease could be achieved.
Finally, the SAR analysis of the synthesized nanocomposite obtained a value of 0.5 ± 0.2   ( W · g 1 ) . Firstly, the obtained SAR value is higher than the values obtained for Gd3Fe5O12 and ZnFe2O4. The composite’s SAR value would be predicted to be lower than the one obtained for the Gd3Fe5O12 sample due to its composition of 65% Gd3Fe5O12 and 10% ZnFe2O4. However, the obtained SAR value allows us to conclude that the mixing of these two crystal phases resulted in a composite with superior thermal efficiency. In this approach, and despite the low value attained, the nanocomposite appears to be more feasible for magnetic hyperthermia treatment than Gd3Fe5O12 and ZnFe2O4 particles alone.

3.5. Biological Analysis

To evaluate the cytotoxicity of the samples, the extract method was used based on different extract concentrations. Analyzing the three different compounds, ZnFe2O4 is the one that presents the highest cytotoxicity. In the case of the synthesized Gd3Fe5O12 particles, it is worth noting that they appear to have a non-cytotoxic behavior for all the studied concentrations. ZnFe2O4 particles appear to be non-cytotoxic only for concentrations lower than 5 mg/mL. The composite exhibits cytotoxic behavior at concentrations of 20 mg/mL. This behavior is a consequence of the strong cytotoxic effect of ZnFe2O4 (Figure 14).
The particles, when in contact with culture media, release ions and present a low dissolution rate due to their high crystallinity. This type of particle can be used not only for MH but also for bone regeneration since the release of gadolinium ions can improve biocompatibility, osteoconductivity, and osteoinductivity [40].

4. Conclusions

This study focuses on a novel composite made of gadolinium iron garnet (Gd3Fe5O12) and zinc ferrite (ZnFe2O4). The nanocomposite was created using an eco-friendly coconut water-assisted sol–gel method. The purest phase of each ferrite (Gd3Fe5O12 and ZnFe2O4) was used to guarantee the purest composition and the best magnetic properties of the nanocomposite. With an average particle size of 155 nm, the composite exhibits ferromagnetic behavior with a saturation magnetization of 11.56 emu/g and a specific absorption rate (SAR) of 0.5 ± 0.2 (W/g). Cytotoxicity tests showed no harmful effects at doses below 10 mg/mL. While its efficiency for magnetic hyperthermia application is poor, the saturation magnetization values obtained for the nanocomposite denote its importance in being used for biomagnetic applications.

Author Contributions

Conceptualization, M.G. and S.S.; Formal analysis, B.C., J.C., T.V., J.C.S., P.I.P.S., M.A.V., S.S. and M.G.; Investigation, B.C., J.C., S.G., T.V., J.C.S., P.I.P.S., M.A.V., S.S. and M.G.; Methodology, B.C., T.V., S.S. and M.G.; Supervision, S.S. and M.G.; Validation, B.C., S.G., J.C.S., P.I.P.S. and M.G.; Writing—original draft, B.C.; Writing—review and editing, B.C., J.C., S.G., T.V., J.C.S., P.I.P.S., M.A.V., S.S. and M.G. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by national funds (OE) through FCT—Fundação para a Ciência e a Tecnologia (PhD grant SFRH/BD/148233/2019), I.P., in the scope of the framework contract foreseen in the numbers 4, 5, and 6 of article 23 of the Decree-Law 57/2016, 29 August, changed by Law 57/2017, 19 July.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data are contained within the article.

Acknowledgments

The authors acknowledge the support of FCT—Fundação para a Ciência e a Tecnologia, I.P., in the scope of the projects LA/P/0037/2020, UIDP/50025/2020 and UIDB/50025/2020 of the Associate Laboratory Institute of Nanostructures, Nanomodelling, and Nanofabrication-i3N. S.R. Gavinho acknowledges FCT—the Portuguese Foundation for Science and Technology for the PhD grant (SFRH/BD/148233/2019).

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. GLOBOCAN. The Global Cancer Observatory—Protaste Cancer in Portugal; International Agent for Research on Cancer, WHO: Geneva, Switzerland, 2020; Volume 501, pp. 1–2.
  2. International Agency for Research on Cancer, W.H.O. Global Cancer Observatory. Available online: https://gco.iarc.fr/tomorrow/en (accessed on 2 May 2023).
  3. Siegel, R.L.; Miller, K.D.; Wagle, N.S.; Jemal, A. Cancer Statistics, 2023. CA Cancer J. Clin. 2023, 73, 17–48. [Google Scholar] [CrossRef] [PubMed]
  4. Bayda, S.; Adeel, M.; Tuccinardi, T.; Cordani, M.; Rizzolio, F. The History of Nanoscience and Nanotechnology: From Chemical–Physical Applications to Nanomedicine. Molecules 2020, 25, 112. [Google Scholar] [CrossRef] [PubMed]
  5. Yu, N.; Zhao, L.; Cheng, D.; Ding, M.; Lyu, Y.; Zhao, J.; Li, J. Radioactive Organic Semiconducting Polymer Nanoparticles for Multimodal Cancer Theranostics. J. Colloid Interface Sci. 2022, 619, 219–228. [Google Scholar] [CrossRef] [PubMed]
  6. Jose, J.; Kumar, R.; Harilal, S.; Mathew, G.E.; Parambi, D.G.T.; Prabhu, A.; Uddin, M.S.; Aleya, L.; Kim, H.; Mathew, B. Magnetic Nanoparticles for Hyperthermia in Cancer Treatment: An Emerging Tool. Environ. Sci. Pollut. Res. 2020, 27, 19214–19225. [Google Scholar] [CrossRef] [PubMed]
  7. da Silva, F.A.S.; de Campos, M.F. Study of Heating Curves Generated by Magnetite Nanoparticles Aiming Application in Magnetic Hyperthermia. Braz. J. Chem. Eng. 2020, 37, 543–553. [Google Scholar] [CrossRef]
  8. Issa, B.; Obaidat, I.M.; Albiss, B.A.; Haik, Y. Magnetic Nanoparticles: Surface Effects and Properties Related to Biomedicine Applications. Int. J. Mol. Sci. 2013, 14, 21266–21305. [Google Scholar] [CrossRef] [PubMed]
  9. Liu, X.; Wang, N.; Liu, X.; Deng, R.; Kang, R.; Xie, L. Vascular Repair by Grafting Based on Magnetic Nanoparticles. Pharmaceutics 2022, 14, 1433. [Google Scholar] [CrossRef]
  10. Gavilán, H.; Simeonidis, K.; Myrovali, E.; Mazarío, E.; Chubykalo-Fesenko, O.; Chantrell, R.; Balcells, L.; Angelakeris, M.; Morales, M.P.; Serantes, D. How Size, Shape and Assembly of Magnetic Nanoparticles Give Rise to Different Hyperthermia Scenarios. Nanoscale 2021, 13, 15631–15646. [Google Scholar] [CrossRef] [PubMed]
  11. Liu, X.; Zhang, Y.; Wang, Y.; Zhu, W.; Li, G.; Ma, X.; Zhang, Y.; Chen, S.; Tiwari, S.; Shi, K.; et al. Comprehensive Understanding of Magnetic Hyperthermia for Improving Antitumor Therapeutic Efficacy. Theranostics 2020, 10, 3793–3815. [Google Scholar] [CrossRef]
  12. Carr, D.H.; Brown, J.; Bydder, G.M.; Weinmann, H.-J.; Speck, U.; Thomas, D.J.; Young, I.R. Intravenous Chelated Gadolinium as a Contrast Agent in NMR Imaging of Cerebral Tumours. Lancet 1984, 323, 484–486. [Google Scholar] [CrossRef]
  13. Spirou, S.V.; Basini, M.; Lascialfari, A.; Sangregorio, C.; Innocenti, C. Magnetic Hyperthermia and Radiation Therapy: Radiobiological Principles and Current Practice. Nanomaterials 2018, 8, 401. [Google Scholar] [CrossRef] [PubMed]
  14. Liu, Y.; Zhang, N. Biomaterials Gadolinium Loaded Nanoparticles in Theranostic Magnetic Resonance Imaging. Biomaterials 2012, 33, 5363–5375. [Google Scholar] [CrossRef]
  15. Lux, F.; Sancey, L.; Bianchi, A.; Crémillieux, Y.; Roux, S.; Tillement, O. Gadolinium-Based Nanoparticles for Theranostic MRI-Radiosensitization. Nanomedicine 2015, 10, 1801–1815. [Google Scholar] [CrossRef] [PubMed]
  16. Jiang, P.S.; Tsai, H.Y.; Drake, P.; Wang, F.N.; Chiang, C.S. Gadolinium-Doped Iron Oxide Nanoparticles Induced Magnetic Field Hyperthermia Combined with Radiotherapy Increases Tumour Response by Vascular Disruption and Improved Oxygenation. Int. J. Hyperth. 2017, 33, 770–778. [Google Scholar] [CrossRef] [PubMed]
  17. Jadhav, S.V.; Shewale, P.S.; Shin, B.C.; Patil, M.P.; Kim, G.D.; Rokade, A.A.; Park, S.S.; Bohara, R.A.; Yu, Y.S. Study of Structural and Magnetic Properties and Heat Induction of Gadolinium-Substituted Manganese Zinc Ferrite Nanoparticles for in Vitro Magnetic Fluid Hyperthermia. J. Colloid Interface Sci. 2019, 541, 192–203. [Google Scholar] [CrossRef] [PubMed]
  18. Somvanshi, S.B.; Kumar, R.V.; Kounsalye, J.S.; Saraf, T.S.; Jadhav, K.M. Investigations of Structural, Magnetic and Induction Heating Properties of Surface Functionalized Zinc Ferrite Nanoparticles for Hyperthermia Applications. AIP Conf. Proc. 2019, 2115, 4–7. [Google Scholar] [CrossRef]
  19. Somvanshi, S.B.; Kharat, P.B.; Khedkar, M.V.; Jadhav, K.M. Hydrophobic to Hydrophilic Surface Transformation of Nano-Scale Zinc Ferrite via Oleic Acid Coating: Magnetic Hyperthermia Study towards Biomedical Applications. Ceram. Int. 2020, 46, 7642–7653. [Google Scholar] [CrossRef]
  20. Alhadlaq, H.A.; Akhtar, M.J.; Ahamed, M. Zinc Ferrite Nanoparticle-Induced Cytotoxicity and Oxidative Stress in Different Human Cells. Cell Biosci. 2015, 5, 55. [Google Scholar] [CrossRef]
  21. Kerroum, M.A.A.; Essyed, A.; Iacovita, C.; Baaziz, W.; Ihiawakrim, D.; Mounkachi, O. The Effect of Basic PH on the Elaboration of ZnFe2O4 Nanoparticles by Co-Precipitation Method: Structural, Magnetic and Hyperthermia Characterization. J. Magn. Magn. Mater. 2019, 478, 239–246. [Google Scholar] [CrossRef]
  22. Teixeira, S.S.; Graça, M.P.F.; Lucas, J.; Valente, M.A.; Soares, P.I.P.; Lança, M.C.; Vieira, T.; Silva, J.C.; Borges, J.P.; Jinga, L.I.; et al. Nanostructured LiFe5O8 by a Biogenic Method for Applications from Electronics to Medicine. Nanomaterials 2021, 11, 193. [Google Scholar] [CrossRef]
  23. Gomes, M.D.A.; Valerio, G.; Macedo, S. Particle Size Control of Y2O3:Eu3+ Prepared via a Coconut Water-Assisted Sol-Gel Method. J. Nanomater. 2011, 2011, 469685. [Google Scholar] [CrossRef]
  24. Gomes, M.A.; Brandão-silva, A.C.; Avila, J.F.M.; Alencar, M.A.; Rodrigues, J.J., Jr.; Macedo, Z.S. Particle Size Effect on Structural and Optical Properties of Y2O3:Nd3+ Nanoparticles Prepared by Coconut Water-Assisted Sol-Gel Route. J. Lumin. 2018, 200, 43–49. [Google Scholar] [CrossRef]
  25. Soreto, J.M.F.L.S.; Gavinho, T.S.R.; Silva, P.R.P.C.C.; Valente, A.J.M.S.M.A. Niobium Oxide Prepared by Sol–Gel Using Powder Coconut Water. J. Mater. Sci. Mater. Electron. 2019, 30, 11346–11353. [Google Scholar] [CrossRef]
  26. Brito, P.C.A.; Gomes, R.F.; Duque, J.G.S.; Macêdo, M.A. SrFe12O19 Prepared by the Proteic Sol-Gel Process. Phys. B Condens. Matter. 2006, 384, 91–93. [Google Scholar] [CrossRef]
  27. Fortes, S.S.; Duque, J.G.S.; Macêdo, M.A. Nanocrystals of BaFe12O19 Obtained by the Proteic Sol-Gel Process. Phys. B Condens. Matter. 2006, 384, 88–90. [Google Scholar] [CrossRef]
  28. Burmeister, C.F.; Kwade, A. Process Engineering with Planetary Ball Mills. Chem. Soc. Rev. 2013, 42, 7660–7667. [Google Scholar] [CrossRef]
  29. ISO 10993-5:2009; Biological Evaluation of Medical Devices—Part 5: Tests for In Vitro Cytotoxicity. ISO: Geneva, Switzerland, 2009.
  30. Huang, Y.; Zhai, X.; Ma, T.; Zhang, M.; Pan, H.; Weijia Lu, W.; Zhao, X.; Sun, T.; Li, Y.; Shen, J.; et al. Rare Earth-Based Materials for Bone Regeneration: Breakthroughs and Advantages. Coord. Chem. Rev. 2022, 450, 214236. [Google Scholar] [CrossRef]
  31. Nagarajan, R.; Ruckenstein, E. Critical Micelle Concentration: A Transition Point for Micellar Size Distribution. A Statistical Thermodynamical Approach. J. Colloid Interface Sci. 1977, 60, 221–231. [Google Scholar] [CrossRef]
  32. Castro, M.J.L.; Ritacco, H.; Kovensky, J.; Fernández-Cirelli, A. A Simplified Method for the Determination of Critical Micelle Concentration. J. Chem. Educ. 2001, 78, 347–348. [Google Scholar] [CrossRef]
  33. Toby, B.H. R Factors in Rietveld Analysis: How Good Is Good Enough? Powder Diffr. 2006, 21, 67–70. [Google Scholar] [CrossRef]
  34. Jiang, L.; Yang, S.; Zheng, M.; Chen, H.; Wu, A. Synthesis and Magnetic Properties of Nanocrystalline Gd3Fe5O12 and GdFeO3 Powders Prepared by Sol–Gel Auto-Combustion Method. Mater. Res. Bull. 2018, 104, 92–96. [Google Scholar] [CrossRef]
  35. Nakai, A.; Shun’ichi, T. Contamination Introduced during Rock Sample Powdering: Effects from Different Mill Materials on Trace Element Contamination. Geochem. J. 2009, 43, 389–394. [Google Scholar] [CrossRef]
  36. El Dek, S.I.; Amin, R.M. Dose Dependent Modifications in Structural, Magnetic and Optical Properties of Gamma-Irradiated Nanocrystalline Gd3Fe5O12 Garnet. Radiat. Phys. Chem. 2023, 204, 110709. [Google Scholar] [CrossRef]
  37. Anjaneyulu, T.; Raghavender, A.; Vijaya Kumar, K.; Narayana Murthy, P.; Narendra, K. Effect of Particle Size on the Structural and Magnetic Properties of Nanocrystalline Zinc Ferrite. Sci. Technol. Arts Res. J. 2014, 3, 48. [Google Scholar] [CrossRef]
  38. Vasquez-Mansilla, M.; Zysler, R.D.; Arciprete, C.; Dimitrijewits, M.I.; Saragovi, C.; Greneche, J.M. Magnetic Interaction Evidence in α-Fe2O3 Nanoparticles by Magnetization and Mossbauer Measurements. J. Magn. Magn. Mater. 1999, 204, 29–35. [Google Scholar] [CrossRef]
  39. Raland, R.D.; Borah, J.P. Efficacy of Heat Generation in CTAB Coated Mn Doped ZnFe2O4 Nanoparticles for Magnetic Hyperthermia. J. Phys. D Appl. Phys. 2017, 50, 17. [Google Scholar] [CrossRef]
  40. Zhao, P.P.; Hu, H.R.; Liu, J.Y.; Ke, Q.F.; Peng, X.Y.; Ding, H.; Guo, Y.P. Gadolinium Phosphate/Chitosan Scaffolds Promote New Bone Regeneration via Smad/Runx2 Pathway. Chem. Eng. J. 2019, 359, 1120–1129. [Google Scholar] [CrossRef]
Figure 1. Schematic representation of the nanocomposite synthesis process.
Figure 1. Schematic representation of the nanocomposite synthesis process.
Materials 17 02949 g001
Figure 2. (a) Conductivity vs. frequency for different concentrations of the CWP solutions. The different concentrations are represented in the graph in the form of [CWP] = “concentration of the solution” mol·dm−3. (b) CMC vs. CWP concentration; f = 10 5 Hz.
Figure 2. (a) Conductivity vs. frequency for different concentrations of the CWP solutions. The different concentrations are represented in the graph in the form of [CWP] = “concentration of the solution” mol·dm−3. (b) CMC vs. CWP concentration; f = 10 5 Hz.
Materials 17 02949 g002
Figure 3. DTA and TG of (a) Gd3Fe5O12 and (b) ZnFe2O4 samples.
Figure 3. DTA and TG of (a) Gd3Fe5O12 and (b) ZnFe2O4 samples.
Materials 17 02949 g003
Figure 4. XRD diffractograms of the Gd3Fe5O12 samples heat-treated at 1200 °C for 24 h (red line) and at 1400 °C for 4 h (blue line).
Figure 4. XRD diffractograms of the Gd3Fe5O12 samples heat-treated at 1200 °C for 24 h (red line) and at 1400 °C for 4 h (blue line).
Materials 17 02949 g004
Figure 5. Rietveld refinement of the Gd3Fe5O12 sample heat-treated at 1200 °C for 24 h.
Figure 5. Rietveld refinement of the Gd3Fe5O12 sample heat-treated at 1200 °C for 24 h.
Materials 17 02949 g005
Figure 6. XRD of the ZnFe2O4 samples heat-treated at 500, 700, 850, 1000, and 1200 °C for 4 h.
Figure 6. XRD of the ZnFe2O4 samples heat-treated at 500, 700, 850, 1000, and 1200 °C for 4 h.
Materials 17 02949 g006
Figure 7. Rietveld refinement of the sample of ZnFe2O4 heat-treated at 1200 °C for 4 h.
Figure 7. Rietveld refinement of the sample of ZnFe2O4 heat-treated at 1200 °C for 4 h.
Materials 17 02949 g007
Figure 8. XRD diffractograms of the ZnFe2O4 (red line) and Gd3Fe5O12 (blue line) samples after one cycle of ball milling.
Figure 8. XRD diffractograms of the ZnFe2O4 (red line) and Gd3Fe5O12 (blue line) samples after one cycle of ball milling.
Materials 17 02949 g008
Figure 9. XRD diffractograms of the produced composite.
Figure 9. XRD diffractograms of the produced composite.
Materials 17 02949 g009
Figure 10. SEM analysis performed on the (a) Gd3Fe5O12 (HT 1200 °C 24 h) and (b) ZnFe2O4 (HT 1200 °C 4 h) samples after heat treatment. These samples were used for the synthesis of the nanocomposite. (c) Gd3Fe5O12, and (d) ZnFe2O4, represent the samples after the first ball-milling cycle performed. These images were used to evaluate if it was necessary to execute another cycle of ball milling. In the end, after the total ball-milling procedure, (e) Gd3Fe5O12 and (f) ZnFe2O4 were obtained.
Figure 10. SEM analysis performed on the (a) Gd3Fe5O12 (HT 1200 °C 24 h) and (b) ZnFe2O4 (HT 1200 °C 4 h) samples after heat treatment. These samples were used for the synthesis of the nanocomposite. (c) Gd3Fe5O12, and (d) ZnFe2O4, represent the samples after the first ball-milling cycle performed. These images were used to evaluate if it was necessary to execute another cycle of ball milling. In the end, after the total ball-milling procedure, (e) Gd3Fe5O12 and (f) ZnFe2O4 were obtained.
Materials 17 02949 g010
Figure 11. SEM analysis performed on the synthesized composite.
Figure 11. SEM analysis performed on the synthesized composite.
Materials 17 02949 g011
Figure 12. (a) Hysteresis loop and the dependence of the inverse with temperature (top inset) for Gd3Fe5O12 and (b) hysteresis loop of ZnFe2O4 samples tested at different temperatures.
Figure 12. (a) Hysteresis loop and the dependence of the inverse with temperature (top inset) for Gd3Fe5O12 and (b) hysteresis loop of ZnFe2O4 samples tested at different temperatures.
Materials 17 02949 g012
Figure 13. Hysteresis loop for composite samples tested at different temperatures.
Figure 13. Hysteresis loop for composite samples tested at different temperatures.
Materials 17 02949 g013
Figure 14. The cell viability of synthesized powders as a function of concentration. The composite is a mixture of Gd3Fe5O12 and ZnFe2O4. C− and C+ is the negative and positive control, respectively. The results were statistically compared for each concentration between each sample, using ANOVA with a significance level of p < 0.05 (represented by an asterisk).
Figure 14. The cell viability of synthesized powders as a function of concentration. The composite is a mixture of Gd3Fe5O12 and ZnFe2O4. C− and C+ is the negative and positive control, respectively. The results were statistically compared for each concentration between each sample, using ANOVA with a significance level of p < 0.05 (represented by an asterisk).
Materials 17 02949 g014
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

Costa, B.; Carvalho, J.; Gavinho, S.; Vieira, T.; Silva, J.C.; Soares, P.I.P.; Valente, M.A.; Soreto, S.; Graça, M. Preparation and Characterization of Zinc Ferrite and Gadolinium Iron Garnet Composite for Biomagnetic Applications. Materials 2024, 17, 2949. https://doi.org/10.3390/ma17122949

AMA Style

Costa B, Carvalho J, Gavinho S, Vieira T, Silva JC, Soares PIP, Valente MA, Soreto S, Graça M. Preparation and Characterization of Zinc Ferrite and Gadolinium Iron Garnet Composite for Biomagnetic Applications. Materials. 2024; 17(12):2949. https://doi.org/10.3390/ma17122949

Chicago/Turabian Style

Costa, Bárbara, João Carvalho, Sílvia Gavinho, Tânia Vieira, Jorge Carvalho Silva, Paula I. P. Soares, Manuel A. Valente, Sílvia Soreto, and Manuel Graça. 2024. "Preparation and Characterization of Zinc Ferrite and Gadolinium Iron Garnet Composite for Biomagnetic Applications" Materials 17, no. 12: 2949. https://doi.org/10.3390/ma17122949

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