3.1 Structure and morphological investigations
The photocatalytic material’s crystalline phase structures were investigated by XRD technique. Figure 1a demonstrates the XRD patterns of the bare NCO, NCO-OV, BiOI and NCBI nanohybrid photocatalysts. The XRD of BiOI was well coordinated to the tetragonal phase of BiOI (JCPDS No. 10–0445) for the peaks at 29.7, 31.9, 45.4, 55.3 and 68.8o corresponds to (102), (110), (200), (212) and (221) planes, respectively (Sasikala et al. 2023). The XRD peaks of NCO and NCO-OV are assigned to the cubic phase of NCO (JCPDS No. 73-1702), while the characteristic peaks of NCO-OV are slightly shifted to lower direction owing to the OV compared from the NCO peaks at 2θ-18.8o, 32.7o, 38.1o, 51.4o, 58.6o, 62.4o and 69.6o corresponds to the (111), (200), (311), (400), (422), (440) and (531) planes (Naz et al. 2024). In the case of NCO-OV/BiOI nanohybrid, both the peaks of BiOI and NCO-OV are in good agreement. Further, the distinctive peaks of NCO-OV in the composites are stronger as the loading of NCO-OV increases. The significant peaks of NCO-OV are apparent on the NCBI nanocomposite, which signifies the efficient formation of NCO-OV with BiOI.
The functional groups of the prepared photocatalytic materials have been determined using FT-IR spectroscopy. From Fig. 1b, the pure BiOI displays peaks at 517 and 779 cm− 1 for Bi-O bond and peaks located at 872–1047 cm− 1 corresponds to the I-O bond. At the time, the stretching and bending modes of surface adsorbed water molecules are identified at 1593 and 3739 cm− 1 (Sasikala et al. 2023). Besides, as seen in Fig. 1b, the NCO possesses two bands at 1631 and 3543 cm− 1 for the -OH bending and stretching vibrations (Katubi et al. 2024). Furthermore, the vibrations of O–M–O, M–O, and M–O–M (M- Co and Ni) can be attributed at 676–823 cm− 1. The FT-IR spectra of NCBI nanohybrid affirms the presence of characteristics peaks of NCO-OV and BiOI.
The material’s structural morphology was investigated using SEM. The SEM images of bare BiOI (Fig. 2a) demonstrate rose-like hierarchical structures formed by an array of 2D-nanosheets. The BiOI flower is composed of several self-assembled irregular plates, resembling petals interconnected in a manner like to the formation of a Rieger Begonia. As illustrated in Fig. 2b, the oxygen-deficient NCO reveals its 1D-nanorods shaped structures. The SEM images of pure NCO possess aggregated like structure, results shown in Fig. S1 (a and b). Many of 1D-NCO-OV nanorods are observed to regularly grow on 2D-BiOI nanosheets. This intimate 1D/2D interface contact between NCO-OV/BiOI is prominently obvious within different regions of NCBI-2 rod-on-rose microflowers (see Fig. 2c), underscoring the formation of heterojunctions between NCO-OV and BiOI, facilitating effective contact. The formation of close 1D/2D contact in the composite is further confirmed from EDAX spectra, shown in Fig. 2d, which exhibits the existence of Bi, O, I, Co and Ni elements in the BiOI and NCO-OV nanohybrid. Fig. S2 (a-c), illustrates the EDAX spectra of NCO without and with defect and BiOI. The existence of Bi, O, I, Co and Ni elements in the composites can be observed by elemental mapping images (Fig. 2e).
XPS investigation was employed to identify the compound's elemental composition and atomic chemical environment. Figure 3 (a-f) depicts the XPS spectra of pure NCO, NCO-OV, BiOI and NCBI nanohybrids. The wide survey XPS spectra depicts the presence of Bi, O, Ni, Co, and I elements in the NCO, BiOI, and NCBI nanohybrids has been demonstrated correspondingly in the survey spectrum of XPS as shown in Fig. 3a. The pure NCO without OV was used to verify the existence of OV defects in the NCO-OV. High resolution XPS spectra of Ni 2p peaks around 856.7 and 874.1 eV for the respective Ni 2p3/2 and Ni 2p1/2 states, and the satellite peaks at 862.6 and 880.2 eV, which represents Ni2+ from the Ni 2p spectrum (Fig. 3b) were observed. The XPS peaks of NCBI-2 exhibit a negative shift, which implies an increase in electron density. This finding further validates the strong interaction and electron transfer between NCO-OV and BiOI. In Fig. 3c, the Co 2p3/2 and Co 2p1/2 doublets appears at 781.5 and 806.3 eV, and also the two satellite peaks at 785.7 and 798.4 eV, indicating Co2+ state. When compared with NCO-OV, NCBI-2 exhibits a towards higher binding energies (positive shift).
Two peaks, situated at 159.5 and 164.8 eV in the high resolution Bi 4f XPS spectrum of BiOI (Fig. 3d), are attributed to the Bi 4f7/2 and Bi 4f5/2 in the Bi3+ chemical states, correspondingly. According to I 3d spectrum of BiOI (Fig. 3e), the peaks at 619.3 and 630.8 eV, correspondingly, coincide with the I 3d5/2 and 3d3/2 of the I- chemical state. Furthermore, the binding energies of I 3d in BiOI are not significantly different from those observed in NCBI-2 hybrid. Moreover, O1s spectra (Fig. 3f) displayed three distinct peaks at 530.2, 531.8, and 533.0 eV, corresponding to lattice oxygen, oxygen vacancies, and surface adsorbed oxygen, correspondingly. When compared to NCO, the O1s spectra of NCBI-2 exhibited a slight shift towards higher binding energies, resulting in a drop in electron density. Also, compared to NCO-OV, NCBI-2 shows a notable reduction in the M–O peak, indicating that the OV defects primarily result from the removal of surface oxygen atoms from the M–O layers. Absorbed hydroxyl groups tend to dehydrate and react with the released Ni and the Bi substrate, forming Ni–O–Bi bonds. This bond serves as an efficient charge transport channel between BiOI and NCO-OV, significantly impacting the photocatalytic performance of the NCBI nanocomposites. This shift indicates the generation of more OV’s after formation of heterojunction between NCO-OV and BiOI. Notably, an increase in electron density is accompanied by a negative shift in binding energy, and vice versa. Hence, the shifting nature of XPS binding energy can be used to assess the electron transport mechanism in the composite. In this case, it appears that NCO-OV served as an e- acceptor in the NCBI composites.
UV-vis DRS was used to assess the optical characteristics of the fabricated materials. Figure 4a presents the absorption spectra of BiOI, NCO, NCO-OV and NCBI-2 hybrid where the absorption edges are determined to be 672, 750, 686 and 712 eV, respectively. Further investigation reveals that there was a red shift in the absorption edges of NCO-OV along with a decrease in the bandgap. When compared with pure NCO, the long absorption tail of NCO-OV demonstrated the existence of high-density surface defects in the material. From the Kubelka-Munk plots (Bavani et al. 2023; Sasikala et al. 2023; Bavani et al. 2022) (Fig. 4b), the bandgaps of NCO and NCO-OV, BiOI and NCBI-2 were determined to be 1.65, 2.05, 2.19 and 2.16 eV, respectively, [32–34]. The rod-on-rose-like 1D/2D binary NCO-OV/BiOI composite has been revealed to be photoresponsive throughout a broad solar spectrum, ranging from UV to NIR, by combining the advantageous properties of both BiOI and NCO-OV. This combination can lead to generate more e−/h+ pairs and taking part in the reactions, thereby improving the efficiency of photocatalytic processes.
The transfer capability and separation of charges is a crucial factor in assessing the performance of photocatalysts. Therefore, PL measurements are essential for this evaluation. Typically, a lower PL intensity signifies better electron-hole separation efficiency. The PL spectra of NCO, BiOI, NCO-OV and NCBI nanohybrid was shown in Fig. 4c. As illustrated in Fig. 5c, all samples exhibit a broad emission peak in the 500–600 nm range. Due to the abundance of oxygen vacancies on the surface of NCBI, the wavelength range of the hybrid has red shift from pure NCO, which can act as trapping sites for photogenerated charge carriers. Moreover, BiOI has the highest peak intensity, indicating the highest rate of charge carrier recombination. However, when BiOI is coupled with NCO-OV, its PL intensity decreases significantly, suggesting that the recombination rate is effectively reduced. Among the samples, NCBI-2 has the smallest PL intensity, representing the reduced electron-hole reconnection rate. As a result, NCBI extend their photogenerated charge carriers’ lifetime and enhances their utilization in photocatalytic reactions.
3.2. Photocatalytic degradation studies
The fabricated samples' photocatalytic activity was assessed by degrading the model pollutant RhB in the presence of light, as shown in Fig. 5a. RhB barely degrades under light without the use of a photocatalyst, so the direct photolysis of RhB is limited. From Fig. 5a, only 33 and 41% of RhB are degraded in 60 min under the influence of NCO-OV and BiOI, correspondingly. The degradation efficiency of NCBI composites with 10, 20, and 30 wt.% of NCO-OV exhibits 62, 89, and 74% degradation efficiency within 60 min of light irradiation. The degradation ability of NCBI-2 nanocomposites is significantly greater compared with pure BiOI and NCO-OV for RhB degradation.
Figure 5b illustrates the linear dependence with time and reveals the photocatalytic RhB decomposition follows apparent pseudo-first-order kinetics, Eq. (1) (Cui et al. 2009; Bavani et al. 2022; Khan et al. 2020; Zulmajdi et al. 2020; Kee et al. 2022),
Where C0 and C are the initial RhB concentration and concentration of the RhB with different time intervals, k is the pseudo-first-order rate constant (min− 1), and t is time (min). The pseudo-first-order kinetics plot with relevant k values is compiled in Fig. 5c; the 0.0073, 0.0092, 0.0148, 0.0279, and 0.0189 min− 1 correspond to BiOI, NCO-OV, NCBI-1, NCBI-2, and NCBI-3 nanocomposites, respectively (Fig. 5c). The k value of the NCBI-2 nanocomposite is around 3.8 and 3.03 times greater than that of BiOI and NCO-OV.
To assess the significant impact of catalyst quantity on the decomposition, the optimized NCBI-2 catalyst was chosen for experimentation. Subsequent experiments were conducted to investigate the photocatalytic degradation of RhB under 60 min of visible light irradiation (VLI), utilizing varying amounts of catalyst ranging from 0.25 to 1.5 g/L. The results obtained, as displayed in Fig. 5d, highlight that the photocatalytic decomposition ability appears to gradually increase up to a catalyst concentration of 1.25 g/L, followed by a progressively reduce thereafter. The initial rise in degradation rate is attributed to the increased accessibility of numerous active sites accessible on the surface of the photocatalyst. However, the subsequent rise in catalyst content leads to a reduction in the decomposition rate due to the shielding effect of the photocatalysts, which impedes the passage of light through the solution. Further, the recycling ability of the NCBI-2 photocatalyst was shown in Fig. 6a, revealing catalyst stability and reusability even after four consecutive cycles.
3.3. Photocatalytic degradation mechanism
Radical trapping studies have identified that major active radicals, including O2•−, h+, and •OH, play crucial roles in the photocatalytic degradation of RhB. Various quenching agents, such as 1 mmol of benzoquinone (BQ), isopropyl alcohol (IPA), and ammonium oxalate (AO), were used to determine their effects (Worku et al. 2024). The results shown in Fig. 6b indicate that the existence of AO (h+) and BQ (O2•−) does not significantly alter the degradation rate compared to when no quenchers are present. However, IPA (•OH) significantly hinders the degradation efficiency. These findings imply that h+ and O2•−radicals are the primary the active species in charge of the degradation by photocatalysis of RhB (Fig. 6b).
The strategy for the photocatalytic degradation mechanism of the NCBI nanocomposite is proposed based on the attained results. Figure 7 illustrates the facilitated charge transfer mechanism for the photodecomposition of RhB on the NCBI system under VLI. To improve photocatalytic performance, the heterojunction formation between the two different types of photocatalysts at different energies is employed as an advantageous method, expediting separation capability, and reducing reconnection rates. Energy levels play a crucial role in this mechanism, and thus the valence band (VB) and conduction band (CB) potentials of the fabricated photocatalysts are determined employing the following equations (2) and (3) (Kong et al. 2019):
EVB = X - Ee + 0.5Eg (2)
ECB = EVB - Eg (3)
Herein, ECB and EVB represent the CB and VB potentials, Ee denotes the energy of free electrons on the hydrogen scale (~ 4.5 eV), X signifies the electronegativity of the semiconductor, and Eg represents the bandgap energy value of the semiconductor. Figure 7 depicts the determined values of the VB and CB potentials of BiOI as 2.14 and − 0.05 eV, respectively, while the CB and VB potentials of NCO are noted as -0.615 eV and 2.26 eV, correspondingly.
When NCO-OV and BiOI come into contact to form NCBI composite, their individual semiconductor properties interact to achieve equilibrium (Fig. 7). Initially, NCO-OV, due to its oxygen vacancies, acts as an n-type semiconductor with its Fermi level (Ef) closer to the CB, while BiOI behaves as a p-type semiconductor with its Ef closer to the VB. The presence of oxygen vacancies in NCO-OV facilitates the formation of active sites for surface reactions, promoting the migration of charge carriers and extending their lifetime. Upon contact, electrons from the n-type NCO-OV diffuse into the p-type BiOI, and holes from the BiOI diffuse into the NCO-OV. This diffusion continues until the Ef of both materials align, creating a depletion region at the interface where no free charge carriers exist [40]. The band structure bends near the junction, establishing an internal electric field that reaches equilibrium when the Fermi levels are uniform across the NCBI composite.
Under light irradiation, the composite of NCO-OV and BiOI exhibits distinct behavior due to the generation of electron-hole pairs by photons with energy greater than the bandgap of the materials. Photons generate electron-hole pairs in both NCO-OV and BiOI as well as in the depletion region. The built-in electric field at the junction separates these photogenerated carriers, driving electrons to the n-type NCO-OV and h+ towards the p-type BiOI. Notably, the e− migrate to the oxygen defect state and significantly impact the electronic properties and reactivity. Apart from that, these electrons also act as recombination centers for photogenerated charge carriers. However, they can participate in redox reactions, aiding RhB degradation into harmless byproducts. Consequently, the resulting potential difference and movement of charge carriers in NCBI composites enhances the efficiency of photocatalytic processes by utilizing the broad solar spectrum.