3.1. Structural analysis
The X-ray diffraction (XRD) analysis was conducted to assess the crystallinity, structural characteristics, and phase purity of the synthesized nanostructures. Figure 1 illustrates the powder XRD patterns obtained from samples prepared using various concentrations of NaHCO3 and different precursor ratios of Ni and Co. The synthesis involved chemical co-precipitation followed by calcination at 300°C for 3 hours, employing NaHCO3 concentrations of 0.5 M, 1 M, and 2 M, and precursor molar ratios of Ni to Co at 1:0.5 and 1:1, respectively. Distinct samples were synthesized with different molarities, including 1.5 M (NCO1), 0.5 M (NCO2), 1 M (NCO3), and 2 M (NCO4), as well as precursor molar ratios of Ni to Co at 1:0.5 (NCO5) and 1:1 (NCO6), respectively. The diffraction peaks observed at 18.6°, 30.9°, 35.5°, 38.3°, 44.4°, 54.9°, 58.8°, and 75.0° correspond to the crystallographic planes of (111), (220), (311), (222), (400), (422), (511), and (440), respectively. These peaks are consistent with the cubic spinel NiCo2O4 phase, as confirmed by comparison with the JCPDS file (JCPDS card No.00-20-0781) [42]. The XRD patterns (Fig. 1A) exhibit prominent peaks indicative of well-crystalline samples, while low-intensity peaks remain unindexed. Notably, no characteristic peaks associated with impurities, such as NiO or CoO, were detected, underscoring the high purity of the synthesized NiCo2O4 samples [43]. Furthermore, variation in the stoichiometric ratio of precursor concentrations during the synthesis of NCO1, from 1:0.5 (NCO5) to 1:1 (NCO6), revealed consistent formation of the cubic spinel NiCo2O4 phase across all XRD peaks. Importantly, no significant alterations in peak position or additional peaks were observed in the XRD patterns of NiCo2O4 samples synthesized under different parameters, including molarity and precursor molar ratios.
The FTIR analysis of NiCo2O4 revealed significant peaks near 2104 cm− 1, suggesting potential bending of water molecules. Additionally, the presence of the H–O–H bending vibration mode was indicated by peaks observed at 1602 cm− 1and 1078 cm− 1. Weak signals at 640 cm− 1 and 487 cm− 1, consistent with all peaks detected in the NiCo2O4 nanostructured materials illustrated in Fig. 2, were attributed to metal oxide vibrations [44].
3.2. Morphological analysis
By varying the concentration of sodium bicarbonate and adjusting the precursor molar ratio, the surface characteristics of the produced NiCo2O4 nanostructures (as depicted in Fig. 3) were investigated. Figure 3 (NCO1) presents a scanning electron microscope (FESEM) image of nickel cobaltite nanoparticles synthesized using water as the solvent and a concentration of 1.5 M NaHCO3.Additionally, nanoflowers were produced when the molarity concentration of NaHCO3 was switched from 1.5 M to 0.5 M (NCO2) while keeping the remaining synthesis parameters constant (Fig. 3(NCO2)). In addition, compared to using little high molarity concentration, the morphology of the sample is significantly changed when using 1 M concentration of sodium bicarbonate (NCO3). Figure 3 (NCO3) shows how nanorods are shaped which have also good electrochemical performance for making efficient electrocatalyst. Using the highest molar concentration of 2 M (NCO4) and keeping the other synthesis variables constant, the shape of nanoparticles similar to those shown in Fig. 3 (NCO4) was found. It is worth emphasised that the solvent is essential for the creation of different NiCo2O4 nanostructures with morphologies resembling nanoparticles, nanoflowers, and nanorods [45]. Further, we achieved 1:0.5 and 1:1 NiCl2.6H2O and CoCl2.6H2O precursor ratios throughout the synthesis of NCO1, while maintaining the other reaction conditions constant. This allowed us to better understand the impact of precursor molar ratio. NiCl2.6H2O and CoCl2.6H2O were mixed in a 1:0.5 molar ratio to produce nanosheets (Fig. 3, NCO5). CoCl2.6H2O and NiCl2.6H2O were combined in a 1:1 molar ratio to create encapsulated nanoparticles with the same shape (Fig. 3, NCO6).
The comprehensive microstructural analysis of NCO3 was completed utilising TEM in addition to support FESEM. Using a 0.05 M of NiCl2.6H2O and 0.1 M of CoCl2.6H2O and 1 M of NaHCO3 in 80 mL of aqueous solution and calcinated at 300°C for 3 hours, NCO3 was synthesised. Figure 4 illustrates the TEM image, selected area electron diffraction (SAED) pattern, and high-resolution transmission electron microscopy (HRTEM) images of NCO3.The TEM picture (Fig. 4a) displays a morphology resembling nanorods made of nanoplates. Moreover, the location of the selected area electron diffraction (SAED) pattern of NCO3 (Fig. 4b) indicates the poly-crystalline structure of the samples. Furthermore, the edge of a nanosheet was used to catch the high-resolution TEM (HRTEM) (Fig. 4c). The (311) crystal plane of the synthesized nanocomposite, identified as the most prominent peak in XRD, corresponds to a fringe spacing of 0.22 nm, as observed in Fig. 1a (NCO3) [46].
3.3. Compositional analysis
X-ray photoelectron spectroscopy (XPS) is employed to gain insights into the elemental composition and chemical states of the prepared NiCo2O4 sample.. Figure 5a shows the Ni, Co, and O components present in the survey region from the XPS survey pattern of NiCo2O4 nanoflowers (1200-0 eV). Using high resolution Ni, Co, and O spectra, the sample's spinel structure's oxidation condition was determined. In Fig. 5b, the high-resolution Ni 2p spectrum is depicted, delineated into the Ni 2p3/2 peak at 850.2 eV and the Ni 2p1/2 peak at 856.6 eV, forming a spin-orbit doublet. The Ni3+2p3/2 and Ni3+2p1/2 states exhibit dual peaks with spin-orbit binding energies of 852.5 and 860.7 eV, correspondingly, while Ni2+2p3/2 and Ni2+2p1/2 display two peaks with binding energies of 848.5 and 856.8 eV, respectively. Additionally, the Co 2p spectrum demonstrates a well-fitted profile with two prominent peaks corresponding to Co 2p3/2 and Co 2p1/2, positioned at 772.5 eV and 792.4 eV, as illustrated in Fig. 5c. In the O 1s spectra (depicted in Fig. 5d), the binding energies of 526.1 eV, 528.2 eV, and 529.5 eV are attributed to the metal-oxygen bond in NiCo2O4, as well as to defect sites with reduced oxygen coordination and oxygen absorption from the surrounding atmosphere[47].
3.4. BET analysis
The BET technique, founded on the adsorption of particular molar species in their gaseous state on the surface, stands as one of the pivotal methods for accurately determining the total specific area of porous samples [48]. The crucial factor in any electrochemical reaction, the active surface area of the synthesized NiCo2O4-based samples, was assessed utilizing the Quantachrome Nova station 100 apparatus through gas adsorption analysis conducted at 77 K. The synthesised NCO3 (Fig. 6(a)) has high surface area of 91.2 m2/g, compared to 30.76 m2/g for the NCO5 (Fig. 6(b)). When compared to NCO3, the nanorod morphology of the synthesized NCO3 substantially enhances the specific surface area. The total pore volume was calculated by assessing the quantity of gas absorbed relative to the applied pressure. The investigation of the pore size distribution of the synthesized NiCo2O4 active materials was conducted using non-local density functional theory. The synthesised NCO3 has a pore volume of 2.184 cm3/g, while the synthesised NCO5 has a pore volume of 0.135 cm3/g. Due to high pore volume of as synthesized NCO3, it indicates the superior electrochemical performance.
3.5. Electrocatalytic study on OER/ORR
The electrocatalytic characteristics of NiCo2O4 nanostructures were investigated using the polarisation plot of the oxygen evolution reaction (OER) (Fig. 7a). The onset potentials of the NiCo2O4 nanostructures (NCO1, NCO2, NCO3, NCO4, NCO5, NCO6, and IrO2) synthesised using the various parameters stated were 1.52, 1.5, 1.49, 1.52, 1.54, 1.56, 1.57, and 1.6, respectively, vs. RHE. In compared to other nickel cobaltite samples and IrO2, the NiCo2O4 material (NCO3) used 260 mV less energy to conduct the oxygen evolution when a reference current density of 10 mA.cm− 2 was used. The highest electrocatalytic characteristics for oxygen evolution are defined by the high double layer region in NiCo2O4 samples with IrO2, while NCO3 revealed the low double layer region in Cdl plots (Fig. 7b). The decreased Tafel slope, which denotes a successful charge-transfer process, can be linked to the increased OER activity NCO3 exhibits IrO2's Tafel slope indicated that the dissociative water adsorption mechanism was its limiting factor [49]. All of the NiCO2O4 samples' tafel slopes were further displayed (SI). In addition to NCO5's outstanding ORR performance in compared to Pt/C, NCO3 also demonstrated strong OER activity. Figure 7(c) shows the Nyquist plot of the synthesized materials using a frequency range of 0–10 kHz and amplitude of 5 kV. In comparison to other NiCo2O4 samples, the synthesised materials NCO3 and NCO5 have a lower Rct value, which indicates that they have good electrocatalytic capabilities for oxygen reduction and evolution reactions.
NCO5 was found to achieve a greater current density than other NiCo2O4 nanomaterials, it was also discovered. Thus, linear sweep voltammetry curve for the oxygen reduction reaction of NiCo2O4 nanoparticles with Pt/C at different rotatory speeds are given in Fig. 7e and the LSV curve of other synthesized NiCo2O4 nanomaterials were shown in supporting information (SI, Fig. S1). As can be seen, at all rotational speeds, all materials reached clearly defined diffusion limiting currents. An activity comparison of Pt/C and NiCo2O4 nanomaterials at 1600 rpm is shown in Fig. 7(d). While Pt/C displayed 15.8 mA.cm− 2 at the same potential, NCO5 displayed a highest current density of -12.3 mA.cm− 2 among all NiCo2O4 catalysts. So Changes in surface area cause a shift in the onset potential. With a nearly 2-fold greater surface area than Co3O4, NiCo2O4 exhibits better activity in terms of current density. As a result, there are more active sites available for oxygen molecule reactions in NiCo2O4. The introduction of Ni species into the NiCo2O4 catalyst enhanced its activity by boosting surface area and shifting the onset potential. The performance exhibited by the NiCo2O4 material surpassed that of comparable reported materials, with the added advantage of being synthesized through a simpler process [50–51]. The presence of nickel (Ni) molecules within the NiCo2O4 catalyst facilitated an improvement in the onset potential. Ni played a crucial role in augmenting electrical conductivity, thereby enhancing electronic properties and facilitating electron transfer [52]. The enhancement in onset potential resulting from nickel substitution is attributed to the reduction in activation energy [53].
Because of the enhancement of the electronic characteristics, the NiCo2O4 material demonstrated an enhancement in the electrocatalytic performance for ORR. The following equations were used to conduct a Koutecky-Levich correlation and further analyse the ORR mechanism.
$$\frac{1}{J}=\frac{1}{{J}_{L}}+\frac{1}{{J}_{K}}=\frac{1}{B{\omega }^{\frac{1}{2}}}+\frac{1}{{J}_{K}}$$
1
B = 0.62nFC0 (D0)2/3\({\vartheta }^{-1/6}\) (2)
JK =nFkC0 (3)
Where Jd, Jk, and Jl are the disc current density, kinetic current density, and diffusion current density, respectively. Furthermore, n represents the electron transfer number, F represents the Faraday constant (96485 C/mol), CO2 represents the oxygen concentration in KOH solution (1.4x10− 6 mol/cm3), DO2 represents the oxygen diffusion coefficient in KOH (1.73x10− 5 cm2/s), is the kinematic viscosity of the KOH solution (0.01 cm2/s), and is the electrode rotation rate (rad/s). Figure 7(f) shows the KL plots obtained at two different potentials of NCO5 samples and other plots were shown in Fig S2. This phenomenon can be visually illustrated by contrasting the outcomes with theoretical K–L plots for 2 and 4 electrons, demonstrating a consistent adherence to a 4-electron trend across all materials. The number of transferred electrons was determined using Eq. 2, yielding values of 3.12, 1.2, 1.9, 1.4, 3.3, 1.9, and 4e- for NCO1, NCO2, NCO3, NCO4, NCO5, NCO6, and Pt/C, respectively. Otherwise, KL plots showed linear behaviour, indicating that these materials had a first-order influence on O2 kinetics [54].