Figure 2 depicts the observation of crystalline phases using the X-ray diffraction (XRD) technique. XRD is utilized to discern the crystal structure phases of the ZnAl2O4-MgTiO3 nanoceramic composite material prepared via the sol-gel method for microwave applications. The obtained sample confirms the presence of crystalline phases, notably the rutile phase of TiO2 and the wurtzite hexagonal phase of ZnAl2O4, alongside MgTi2O5 phases. The sintered ZTMg nanoceramics comprise wurtzite hexagonal zinc aluminate, magnesium dititanate, and rutile and anatase phases of TiO2, which are indexed in JCPDS file numbers 05-0669, 98-003-7232, 21-1276, and 22-1272, respectively [9, 10]. The prepared ZTMg nanoceramic compound exhibits excellent microwave dielectric properties through the sol-gel process (Fig. 1), specifically tailored for microwave applications.
Figure 3 presents the Raman spectra of the ZTMg nanoceramic composite material. The main vibration modes were identified at 346 and 717 cm⁻¹, with additional bands observed at 284, 390, and 484 cm⁻¹, corresponding to the A1g and B1g modes of TiO2. Similarly, distinctive vibrations of ZnO were detected at 346 cm⁻¹ (2nd order Raman scattering, E2-E1 vibration modes) and 390 cm⁻¹ (assigned to Eg mode). Additional bands, particularly at 346 and 717 cm⁻¹, suggested the interaction between ZnO and TiO2 in the case of hybrid ZnO/TiO2 particles, indicating the presence of ZnTiO3. The thermodynamically stable rutile phase exhibited a minor peak at 846 cm⁻¹ in the prepared ZTMg ceramic nanoparticle [11, 12, 13, 14].
The morphological structure was examined using field emission scanning electron microscopy (FESEM). Figure 4 displays FESEM images of the composite nanoceramic material, along with energy dispersive X-ray spectroscopy (EDS) measurements. In Fig. 4 (a, b), the spherical nanoparticles of the prepared nanoceramic composite sample (ZTMg) are visible, exhibiting agglomeration. The calculated mean diameter of the grain size is 19.26 nm. EDS measurements were employed to analyze the compositional elements, revealing distinct peaks corresponding to Zn, Al, Ti, and Mg at energy levels of 1, 1.5, 0.5, and 1.25 eV, respectively, as depicted in Fig. 4 (c). The corresponding histogram image of the ZTMg nanoparticles is presented in Fig. 4 (d). Following these characterizations, the dielectric characteristics of the produced nanoceramic composite material were evaluated using an LCR meter.
Before testing, the nanoceramic composite material underwent a transformation into pellets using the hot press method, as depicted in Fig. 5. In the pellet-making process, the initially prepared ceramic nanoparticles (ZTMg) were mixed with ethyl cellulose to improve stickiness. The resulting paste was then poured into the pellet machine, and the top handle was rotated to compress the powder into pellet form. Ultimately, pellets with precise dimensions, measuring 1.12 mm in thickness and 10 mm in diameter, were produced.
Moreover, we examined the dielectric properties of ZTMg nanoparticles using an LCR meter. The dielectric permittivity was calculated using the formula εr = Cd/(ε0A), where C represents capacitance, and d and A denote the distance and area of the cross-sections of the prepared pellet. Figure 6 (a) illustrates the relationship between dielectric permittivity and frequency across various temperatures, ranging from 30 to 150 degrees Celsius. The dielectric permittivity showed variations in the range of 13.46 to 19.39. The graph depicts an initial increase in dielectric permittivity from 100 Hz to 60 KHz, followed by a consistent trend as the frequency increases from 100 KHz to 10 MHz, across the different temperatures. These fluctuations in dielectric permittivity are attributed to the dipole movements of charge carriers. At higher frequencies, the permittivity of the material stabilizes and eventually decreases to lower levels due to dipole oscillations that impede dipole orientation in response to the applied field [15].
At elevated frequencies, ferroelectrics typically exhibit a reduction in dielectric permittivity. At high temperatures (150°C), the measured dielectric permittivity was 18.9. However, for each composition at lower temperatures, the dielectric permittivity value remained steady as the frequency increased, suggesting the absence of charge accumulation at the interface.
As depicted in the graph, the dielectric loss decreases with increasing frequency from 30°C to 150°C. This loss is attributed to a delay in polarization response to an alternating field, as well as impurities, imperfections, and electron hopping from Ti+ 2 to Zn+ 2. Maintaining a low loss tangent is critical for effective antennas as it enhances overall radiation efficiency [16]. The observed dielectric loss ranged between 0.25–0.64. According to the graph, the dielectric loss decreased from 1 KHz onwards, with a more significant reduction observed at each subsequent frequency step. This pattern repeated at higher frequencies and temperatures, as depicted in Fig. 6 (b).
Similarly, in Fig. 7, the conductivity graph depicts the relationship between dielectric conductivity (σAC) and frequency across various temperatures (30–150°C). The conductivity demonstrates a gradual rise from 100 Hz to 10 MHz, and beyond 10 KHz, all curves exhibit a slow but steady increase, reaching their peak levels. The conductivity can be explained using the expression σ = ωεoεtanδ. This pattern is attributed to thermally induced charge carrier hopping between different localized states. Furthermore, the maximum value of AC conductivity increases with frequency due to enhanced electron migration [16].
Figure 8 (a) presents the real section of the impedance for the developed nanoceramic composite material. The Z value decreases as frequency and temperature increase, indicating a direct correlation with the material's electrical conductivity. At different temperatures, the multiple curves display varying impedance values that eventually converge to sustain a constant value, possibly diminishing towards zero.
Concurrently, the imaginary component of the impedance (Fig. 8 (b)) holds significance in the fabrication of microstrip patch antennas employing microwave dielectric materials. It also fluctuates with frequency across various temperatures. The impedance values steadily diminish in a positive trajectory and stabilize as the curves converge, ultimately approaching zero [17].
Upon transforming the dielectric properties of the prepared composite material from powder to paste form for antenna fabrication, commence by blending 0.9 grams of ZTMg powder with 1 milliliter of CH3COOH for 20 minutes. Subsequently, introduce 0.20 milliliters of PEG and grind the mixture for 30 minutes. Transfer the resulting solution and excess acetic acid into a glass bottle and subject it to heating in a hot air oven at temperatures exceeding 120°C until it reaches a viscous paste consistency. Utilize a doctor blade to apply this thick paste onto the fluorine-doped tin oxide (FTO) plate. Once the FTO substrate is coated with ZTMg, employ magnetron sputtering to deposit a silver layer onto the patch antenna for electrical connectivity. Additionally, apply silver to the bottom of the FTO substrate. Finally, to assemble the 3.16 mm × 2.03 mm microstrip patch antenna, affix it to an SMA coaxial connector and conduct testing using a vector network analyzer. The entire process is depicted in Fig. 9.
Following the fabrication of the antenna, its performance underwent assessment utilizing a Vector Network Analyzer (VNA). The VNA furnished comprehensive measurements of the antenna's parameters, encompassing return loss, bandwidth, and impedance matching. These outcomes were meticulously documented and scrutinized. To ensure the precision of the physical measurements, the antenna's performance underwent verification through simulations employing the High-Frequency Structure Simulator (HFSS). The HFSS simulations established a comparative reference point to evaluate real-world performance, facilitating a comprehensive validation of the fabricated antenna's attributes against the designated specifications. The corresponding flow diagram is depicted in Fig. 10.
The fabricated prototype microstrip patch antenna showcases outstanding performance concerning return loss and voltage standing wave ratio (VSWR < 2) in both simulated and measured evaluations. Illustrated in Fig. 11 are the return loss (RL) characteristics of the synthesized 89Wt%(ZnAl2O4TiO2)11Wt%MgTiO3 (ZTMg) ceramic nanoparticles within the fabricated antenna, spanning operating frequencies from 1 to 10 GHz. Specifically, at the designated operating frequency of 3.25 GHz, accompanied by a bandwidth spanning 1.36 GHz, the return loss measured at -20.51 dB [18–20]. In comparison, Mazumdar et al. reported a fabricated antenna resonating at 3.68 GHz with a return loss of -16.15 dB and a bandwidth of 15.58 MHz [21]. Our investigation underscores that the microstrip patch antenna, leveraging ZTMg composite nanoparticles, demonstrates the lowest return loss, rendering it highly compatible with S-band communication applications. Furthermore, the inset of Fig. 10 showcases a digital image of the microstrip patch antenna, featuring dimensions of 3.16×2.03 mm², thereby providing additional insights into the physical dimensions and structure of the fabricated antenna.
Ultimately, the crafted ceramic dielectric nanoparticles emerged as exceptionally adept for microwave applications. Leveraging these nanoparticles, the fabricated antenna exhibited outstanding performance attributes, positioning it as an optimal choice for S-band applications. The harmonious integration of the nanoparticles' dielectric properties and the antenna's design culminated in an efficient and dependable device, proficient in meeting the rigorous demands of S-band communication systems. This symbiotic relationship not only ensures enhanced performance but also offers several advantages. For instance, the utilization of ceramic dielectric nanoparticles contributes to the antenna's compactness, lightweight nature, and durability, making it ideal for deployment in various S-band communication scenarios. Additionally, the inherent stability and reliability of ceramic materials enhance the antenna's longevity and operational consistency, ensuring uninterrupted communication even in challenging environments. Thus, the adoption of these advanced materials and design strategies underscores the antenna's prowess in delivering optimal performance and fulfilling the diverse requirements of modern S-band communication networks.