3.1. Analysing structures using XRD
The X-Ray Diffraction (XRD) patterns of graphene oxide (GO) and reduced graphene oxide (rGO) reveal significant structural differences (Fig. 2). The GO pattern (red line) has a significant peak at 2θ = 10.43°, which corresponds to the (001) reflection [29]. This implies larger interlayer spacing due to the presence of oxygen-containing functional groups and intercalated water molecules. In comparison, the rGO pattern (blue line) suffers a significant change [30]. The large GO peak at 10.43° is much diminished or absent, indicating that the bulk of oxygen-containing groups were eliminated during the reduction. The peak at 2θ = 25.08° represents the (002) reflection of graphitic structures, showing partial structural restoration [31]. The broadness of this peak implies that rGO sheets are not as well-ordered as pure graphite, most likely due to remaining defects and residual functional groups. The Debye-Scherrer formula may be used to calculate the size of these materials' crystallites.
D = Kλ / (β cosθ),
Where D is the mean crystallite size, K is the form factor (usually 0.9), λ is the X-ray wavelength, β is the peak's FWHM in radians, and θ is the Bragg angle. The Debye Scherer formula was used to calculate the crystallite size 15nm and 27nm for GO and rGO respectively [32]. The greater backdrop of the rGO pattern indicates a more disordered structure than GO. These XRD patterns, together with the Debye-Scherrer analysis, clearly show the structural development from GO to rGO, emphasizing the removal of oxygen groups, partial recovery of graphitic structure, and variations in crystallite size during the reduction process [33].
3.2. Morphological Study
The SEM and EDAX images provide complementary information on the structural and chemical characteristics of Graphene Oxide (GO) and reduced graphene oxide (Fig. 3). SEM micrographs reveal distinct morphological differences: GO (picture a) exhibits a highly wrinkled and crumpled sheet-like shape that is typical of its oxidized condition, but rGO (image b) appears more compact with reduced folding, indicating that the graphene structure has been partially restored following reduction. A particles size of GO is 29nm and rGO is 32nm calculated by Image J software. [34, 35].
The EDAX spectra presented below support these compositional observations. Spectrum 1 (Fig. 4a), which most likely corresponds to GO, has distinct peaks for both carbon and oxygen, with a notably strong oxygen signal suggesting GO's high oxygen content. Spectrum 2 (Fig. 4b), which likely represents rGO, shows a substantially smaller oxygen peak than carbon, which is consistent with the removal of oxygen-containing functional groups during the reduction process. These combined studies successfully demonstrate the structural and chemical changes that occur when GO is converted to rGO, highlighting the need of integrating several analytical methods in material characterization [36]. There are many elements present in EDAX because the modified Hummers' method involves the use of reagents like potassium permanganate (KMnO₄), sulfuric acid (H₂SO₄), and others that may leave behind traces of elements like sulfur (S), manganese (Mn), potassium (K), and sodium (Na).
3.3. Vibrational Spectral Analysis
Figure 5 shows the FTIR spectra for GO-rGO. The vibrational spectra of the materials were analysed using FTIR to determine the presence of functional groups. This study is based on the vibrational excitation of molecular bonds generated by the absorption of infrared light energy at wavelengths ranging from 4000 to 500 cm− 1. Following oxidation, Graphite Oxide (GO) is expected to include more oxygen-containing functional groups [38]. However, after reduction, rGO is expected to have a minimal number of oxygen-containing functional groups. In contrast, for GO, the strong and broad signal at 3432cm− 1 demonstrated the presence of an O-H bond (hydroxyl group). Furthermore, -C = O stretching (-COOH group) was seen at 1642 cm-1, whilst C-O-C stretching (epoxy group) may be seen between 1388 and 1111 cm-1[38]. The presence of all of these carboxylic, hydroxyl, epoxide, and carbonyl groups resulted in oxygen molecules (O) being highly occupied at the edge and basal planes of GO, indicating that GO was properly synthesized [39]. The rGO signal at 3425 cm-1 grew narrower than GO's, indicating that the hydroxyl group had been substantially eliminated. Other peaks at 1642, 1384, and 1112 cm-1 became less prominent than those in GO FTIR spectra, indicating that oxygen was eliminated during the reduction with hydrazine hydrate. As a consequence, oxygen-containing functional groups were partially removed, leaving only trace amounts of functional group residue at the rGO's edge and basal plane [40].
3.4. Optical Properties of GO-rGO
The UV-Visible absorption spectra in this image show a comparative analysis of Graphene Oxide (GO) and reduced Graphene Oxide (rGO), revealing substantial structural and electrical differences. GO exhibits two distinct spectral features: a prominent peak at 232 nm owing to π→π* transitions of aromatic C = C bonds, and a pronounced shoulder at 300 nm due to n→π* transitions of C = O bonds [41]. In contrast, rGO's spectrum changes dramatically, suggesting a reduced state. The red shift of the C = C bond peak to 260–270 nm suggests that the sp2 carbon network has been re-established (Fig. 6). Simultaneously, the C = O shoulder has dropped dramatically, suggesting that oxygen-containing groups were effectively eliminated during the reduction process. Perhaps, rGO shows higher absorbance throughout the whole visible spectrum (300–800 nm), suggesting better electronic conjugation than its GO precursor. These spectrum characteristics, when combined, provide strong evidence of the structural change from GO to rGO, as well as showing the reduced material's superior electrical capabilities [42].
The Tauc plots (Fig. 6) in the right-hand graphs give important insights into the optical bandgaps of Graphene Oxide (GO) and reduced Graphene Oxide (rGO). The charts show the relationship between (αhv)² and photon energy (hν), allowing for linear extrapolation to the x-axis to determine each material's bandgap. The top figure, which most likely represents GO, shows a significant optical bandgap of 2.42 eV, which is consistent with GO's broken sp2 network and oxygen-rich structure [43]. In sharp contrast, the bottom figure for rGO indicates a significantly lower bandgap of 1.33 eV. This considerable decrease in bandgap energy demonstrates the restoration of the sp2 carbon network and the successful removal of oxygen functions during the reduction process, resulting in increased electrical conductivity. These Tauc plot investigations, when paired with the UV-Visible spectra, offer a comprehensive picture of the GO to rGO transition [44]. The observed red-shift in absorption peak, increased absorption across the visible spectrum, and significant reduction in bandgap all show the success of the reduction technique in removing oxygen groups and restoring a more graphene-like structure in rGO. The changes in electronic structure and optical properties are important for applications in electronics, sensors, and energy storage devices, where rGO's increased conductivity and altered electronic characteristics might boost performance [45].
3.5. Raman spectra
Raman spectroscopy gives information based on the inelastic (Raman) scattering of a molecule when subjected to monochromatic light, often a laser. Figure 7 depicts the Raman spectra for GO and RGO. Figure 6 shows that GO and RGO have two basic vibrations at 1100 and 1700 cm-1. The D vibration band, which is created by a breathing mode of j-point photons with A1g symmetry, may be discovered at 1348.31 and 1353.20 cm-1 for GO and RGO, respectively [46]. In contrast, the G vibration band generated by first-order scattering of E2g phonons by sp2 carbon was detected at 1594.19 cm-1 for GO and 1586.56 cm-1 for RGO. In addition, the presence of the stretching C-C bond, which is prevalent in all sp2 carbon systems, adds to the G vibration band. The Raman spectrum (Fig. 6) displays the disorder and tangential bands, denoted by the letters D and G [47].
Figure 7 shows a large and shifted to higher wavenumber 2D band for GO at 2716.77 cm-1. Because the 2D band is highly sensitive to graphene layer stacking, it may be used to identify graphene layers (monolayer, double layer, or multilayer). Thus, the location of the 2D band confirms that the produced GO was multilayer, as monolayer graphene is typically detected around 2679 cm-1 in the spectra [48]. The displaced location of the 2D band is due to the presence of oxygen-containing functional groups that prevent the graphene layer from stacking. In addition, rGO has a 2D band at 2706.20 cm-1, as seen in Fig. 7. It is because, after reducing GO to rGO, less residue of oxygen-containing functional groups remained, causing RGO to stack [49]. The location of the 2D band in this experiment is extremely similar to the work presented by Thakur and Karak, in which GO was reduced by employing phytochemicals taken from leaves, peels, or other parts of plants. Phytochemicals were employed instead of chemicals as a reducing agent [50].
The ID/IG ratio for GO was calculated to be 0.85. Following reduction, the ID/IG for rGO improved to 0.91 due to the restoration of sp2 carbon, despite the fact that the average size of sp2 domains fell. The higher intensity in the D band also indicated that RGO included more isolated graphene domains than GO, as well as the removal of oxygen constituent from GO upon reduction [51].
3.6. Electrochemical analysis
The electrochemical properties were assessed using CV and EIS. The measurements were performed using three electrodes. The working electrode is a Glassy Carbon Electrode (GCE) covered with an active substance, the counter electrode is platinum wire, and the reference electrode is a Saturated Calomel Electrode (SCE). Tetrabutylammonium perchlorate 1M (C16H36ClNO4) was used as the electrolyte. The potential window CV extends from − 0.6 to 0.8 V, with an EIS frequency range of 0.1 Hz to 100 kHz.
3.6.1. Cyclic Voltammetry
CV measurements for GO-rGO were carried out in a 1 M C16H36ClNO4 electrolyte at a potential range of -0.6 to 0.8 V. These experiments revealed a strong redox peak, showing that both electrode materials functioned in a pseudocapacitive manner. Figures 8 illustrate the CV curves for GO-rGO, respectively [52]. Curves for newly generated samples were obtained at nine different scan speeds ranging from 10 to 200 mV/s at room temperature. Even at the slowest scan rate of 10 mV/s, the area under the CV curve for rGO is clearly greater than that for GO. Because specific capacitance (Csp) is proportional to area under the CV curve, rGO has a greater specific capacitance. As shown in Fig. 8, the CV curves displayed discrete redox peaks indicating pseudocapacitive behavior, implying effective ion migration from the electrolyte to electrode [53].
The specific capacitance (Csp) values for the GO electrode were determined to be 219, 109, 73, 54, 43, 36, 27, 21, and 10 F/g− 1 at scan speeds of 10, 20, 30, 40, 50, 60, 80, 100, and 200 mV/s, respectively (Fig. 8). At the same scan speeds, the rGO electrode had Csp values of 359, 179, 119, 89, 71, 59, 44, 35, and 17 F/g− 1. The specific capacitance of rGO falls dramatically with increased scan speeds due to the limiting of ion migration by diffusion in the electrolyte [54]. Furthermore, peak currents for oxidation and reduction increased with scan rate, demonstrating a diffusion-controlled redox process for charge storage. Furthermore, when the scan rate increased, the peak potentials for oxidation and reduction changed slightly higher and lower, respectively, which can be ascribed to polarization effects as well as ion diffusion. The higher capacitive response of rGO in the 1 M aqueous C16H36ClNO4 electrolyte is ascribed to its increased surface area when compared to GO, showing strong reversibility and pseudocapacitance of the synthesised materials [55].
Figures 9 plot the specific capacitance curves for GO-rGO vs scan rate. This graphic depicts how raising the scan rate lowers specific capacitance, mostly through the ion exchange mechanism. The higher specific capacitance of rGO was attributable to its bigger surface area when compared to GO. The increased number of redox sites in the synthesized materials results in pseudocapacitive behaviour, which improves rGO's capacitive activity. Thus, CV analysis shows that the rGO electrode is one of the most promising candidates for supercapacitor applications [56].
3.6.2. Electrochemical Impedance Spectroscopy
The Electrochemical Impedance Spectroscopy (EIS) graph (Fig. 10) presented compares the electrochemical behavior of Graphene Oxide (GO) and reduced Graphene Oxide (rGO) in the context of supercapacitor applications. The Nyquist plot displays the complex impedance, with the imaginary part (-Z (im)) plotted against the real part (Z (re)), revealing crucial insights into the materials' electrochemical properties [57]. The large semicircles in the Nyquist plots indicate high charge transfer resistance, which explains why the resistances are so large. GO is known to be less conductive than RGO due to oxygen-containing functional groups disrupting the sp2 carbon network. The RGO sample shows slightly lower impedance (smaller semicircle) compared to GO, which is shown due to its more conductive nature [58].
Both GO-rGO exhibit characteristic super capacitor behavior, with a semicircle in the high-frequency region followed by a sloped line in the low-frequency domain. However, rGO demonstrates markedly superior performance across several key parameters. The smaller semicircle diameter of rGO indicates lower charge transfer resistance, suggesting enhanced electrical conductivity and faster electron transfer kinetics compared to GO [59]. This translates to improved power delivery capabilities, a critical factor in supercapacitor performance. In the low-frequency region, rGO's more vertical line approaches ideal capacitor behavior more closely than GO's more sloped line. This indicates that rGO offers better capacitive characteristics and lower diffusion resistance, which are essential for efficient energy storage and release in supercapacitor applications [60]. The overall lower impedance values observed for rGO further reinforce its superior electrochemical performance.
The equivalent circuit model included in the graph helps interpret these results, illustrating the interplay between solution resistance (Rs), charge transfer resistance (Rct), double layer capacitance (Cdl), and Warburg impedance (W). rGO's performance in this model suggests it would excel in high-frequency applications, crucial for rapid charge/discharge cycles in supercapacitors [61]. In conclusion, this EIS analysis strongly indicates that RGO would be the preferred material for supercapacitor applications. Its lower impedance, superior charge transfer properties, and more ideal capacitive behavior promise better rate capability, higher power performance, and overall more efficient energy storage and delivery compared to GO. These characteristics make RGO a promising candidate for advancing supercapacitor technology and meeting the demands of high-performance energy storage systems [62].