The UV-visible spectrum of SN@CuO NPs formed using Solanum nigrum leaf extract has been depicted in Fig. 1. It contains specific absorbance band at 320 nm belonging to cinnamoyl rings of phenolic compounds present in the Solanum nigrum leaf extract and nearly at 225 nm related to benzoyl system of phytochemicals (Nasrollahzadeh et al. 2019). This spectrum consists of a sharp band at 368 nm suggesting the formation of SN@CuO NPs at room temperature. The surface plasmon resonance shown by SN@CuO NPs is due to the spherical shape and small size of these nanostructures (Naika et al. 2015).
The FTIR investigation was done to recognize the nature and class of phytochemicals present in Solanum nigrum leaf extract in charge of fabrication of SN@CuO NPs and adsorbed on the surface of nanospheres as outshining or stabilizing agents. The FTIR spectrum of Solanum nigrum leaf extract (Fig. 2a) involved three leading peaks nearly at 3309 cm-1, 2125 cm-1, and 1634 cm-1. The broad and strongest peak at 3309 cm-1 commonly correlates to N-H stretching, and the second peak in spectra at 1634 cm-1 indicates the stretching frequency of the carbonyl (C = O) group of secondary amides. It also consists of minor peaks between 1240–1260 cm-1, which indicates an amide band of random coils of protein molecules (Ramesh et al. 2015). The broad and intense peak between 3250–3450 cm-1 is due to O-H and N-H stretching of phenols,(Larsson 2014) amines, and amide linkages of enzymes or proteins (Shankar et al. 2003). In the FTIR spectrum of Solanum nigrum leaf extract outshined CuO nanoparticles (Fig. 2b) peak at 3309 cm-1 due to N-H stretching remains unvaried; however, C = O stretching frequency of secondary amides has been somewhat conveyed to the lower frequency at 1626 cm-1. This designates that the interplay of secondary amides of Solanum nigrum leaf extract with CuII ions occurs through C = O (Cu-C = O). The FTIR results demonstrated that Solanum nigrum leaf extract participated in two main purpose, i.e., in the formation of SN@CuO NPs and their stabilization by adsorbing or outshining on nanoparticle surface.
The synthesized CuO nanoparticles capped by Solanum nigrum leaf extract were dried in oven @80°C, and an XRD pattern was recorded to detect the phase and crystallinity of SN@CuO NPs. In the XRD pattern (Fig. 3) appearance of sharp, intense, and broadened peaks specified the preparation of well crystalline, mesoporous nanospheres. The significant peaks were brought to light at 2θ values 32.1, 35.4, 38.2, 46.4, 48.3, 53.9, 57.6, 61.8, 65.7, 67.6, 72.6, 74.7 and 80.8° representing the reflection planes (Miller indices; (hkl)) of (110), (002), (111), (112), (-202), (020), (202), (-113), (-311), (-220), (311) and (-222)), respectively, which indicated the formation of CuO (JCPDS card no: 45–0937), as reported earlier. The mean size of Solanum nigrum leaf extract mediated CuO nanoparticles was calculated using the Debye Scherrer formula, Eq. (1).
$$D=\frac{0.9\lambda }{\beta cos\theta }$$
1
Where D is the mean crystallite size, λ signifies wavelength of X-ray (0.1546 nm) used, β stands for FWHM (full width at half the maximum) in radians, while θ for the Bragg’s angle. The calculated mean size of the SN@CuO NPs was found to be 9.5 nm using (111) XRD peak.
The morphology, size and elemental composition of fabricated SN@CuO NPs were examined by FE-SEM, TEM and Energy-dispersive X-ray spectroscopy (EDS) analyses as shown in Fig. 4. In the FE-SEM images, 80–120 nm diameter spherical particles are clearly visible (Fig. 4a,b). FE-SEM images further revealed that most of the SN@CuO NPs have indistinguishable shapes and sizes, stipulating uniformity throughout the SN@CuO NPs sample. The EDS spectrum and elemental mapping of Solanum nigrum leaf extract outshined CuO nanoparticles revealed that apart from Cu and O, the presence of P, S and Cl due to presence of phytochemicals on the surface of CuO nanoparticles (Fig. 4c). In order to examine clear morphological view of Solanum nigrum leaf extract outshined CuO nanoparticles, TEM micrographs have been recorded (Fig. 4d-h). TEM images revealed that spherical shapes, observed in the FESEM images, are composed of small sized (5–6 nm) SN@CuO NPs. These small size SN@CuO NPs are regularly arranged and uniformly distributed in bigger size spherical aggregate (80–120 nm diameter) of SN@CuO NPs. The selected area electron diffraction pattern (Fig. 4i) specifies the polycrystalline nature while the ring pattern of the same confirms the formation of monoclinic CuO nanoparticles. The size distribution of CuO nanoparticles has been determined using ImageJ software, which signifies that more than 70% Solanum nigrum capped nanospheres belong to the size range of 4–7 nm (Fig. S1 supporting informations) (Rueden et al. 2017).
Nitrogen adsorption/desorption isotherm plots were analyzed to calculate the fundamental surface properties of synthesized SN@CuO NPs such as pore-size distribution and specific surface area. A type-IV isotherm plot and H1-type hysteresis loop were supported the mesoporosity for synthesized SN@CuO NPs (Fig. 5) (Al Bataineh et al. 2020). The specific surface area, calculated using the standard multi-point Brunauer-Emmett-Teller (BET) method, was 128 m2g− 1. Barrett-Joyner-Halenda (BJH) method was applied to estimate the average pore diameter and was found to be 12 nm (Fig. 5 inset). The higher surface area of SN@CuO NPs indicates the presence of large number of catalytic sites and is thus expected to show high catalytic properties (Abbas et al. 2019).
The surface charge of synthesized SN@CuO NPs was evaluated using Dynamic light scattering (DLS) analysis. The Zeta potential curve of SN@CuO NPs was determined in the range of -150 to + 150 mV. Solanum nigrum leaf extract outshined copper oxide nanoparticles possess positive and high zeta potential value of 56 ± 05 mV (Fig. S2). The high absolute zeta potential value designates a strong force of repulsion among the SN@CuO NPs and prevents the aggregation of particles. Zeta potential analysis result has suggested the highly stable nature of synthesized SN@CuO NPs.
The catalytic activity of SN@CuO NPs has been examined in the reduction of dye Congo red using reductant sodium borohydride as a model reaction. In this experiment, 10 mL of freshly prepared aqueous NaBH4 (0.30 M) solution was poured immediately into 10 mL (1.0 mM) aqueous solution of Congo red dye, thus the pH of resulting solution increased instantaneously from 7 to 10. In the UV-Visible absorption spectrum of pure Congo red solution, the observed peak at 495 nm is associated with azo chromophore (-N = N- ). The reduction of Congo red was also investigated in the presence and in absence of SN@CuO NPs catalyst using UV-visible spectrophotometer (Fig. 6a). In the absence of SN@CuO NPs catalyst, no significant change in the absorption maximum at 495 nm (due to the azo group -N = N-) has been observed even after eight hour time, while in the presence of 5 mg of SN@CuO NPs, the decolorization of Congo red solution taken place rapidly. Time-dependent absorption spectra of reaction solution (Fig. 6b), shows that intensity of absorption band at 495 nm is continuously lowering down and red colour of solution continuously fading with increasing treatment time. This band completely diminished and reaction mixture became colorless within 432 seconds of time. The disappearance of the absorption maximum at 495 nm confirmed the complete catalytic reduction of the -N = N- group of Congo red. Concurrently a new peak appeared at wavelength 285 nm in the UV-visible absorption spectrum of reaction mixture, which can be explained as a result of cleavage of the azo bond (-N = N-) and formation of a new reduction product. The appearance of the absorption peaks at 285 nm is correspond to benzidine on 432 seconds of reaction time (Khan et al. 2017).
In order to further confirm the reduction products after disappearance of azo group (Stage-1), a fraction of reaction mixture was centrifuged at 6000 rpm for 10 minutes, and the SN@CuO NPs catalyst was separated. The supernatant of reaction mixture was extracted thrice using diethyl ether, dried with the help of anhydrous sodium sulfate, and evaporated under reduced pressure. The obtained solid degradation product was examined using FTIR, 1H-NMR, 13C-NMR, and LC-MS. The FTIR spectrum of pure Congo red has shown peaks at 3468 and 1586 cm− 1, which are correspond to -N-H and -N = N- stretching respectively of Congo red (Fig S3). The peak at 1448 cm− 1 is for aromatic C = C stretching vibration, 1357 cm− 1 for C-N bending vibrations, and 1225, 1178, and 1063 cm− 1 are attributed to S = O stretching vibrations of -SO3 group of Congo red (Telke et al. 2010). FTIR spectrum of the reduction products exhibited a peak around 3378 cm− 1 for N-H stretching, 2923 cm− 1 and 2851 cm− 1 for C-H stretching vibration, 1609 cm− 1 for N-H bending, and 1497 cm− 1 for C-H bending. The intensity of peaks corresponds to 1178 cm− 1 is decreased while the peak at 1238 cm− 1 is blue shifted. The peak of -N = N- at 1586 cm− 1 is completely disappeared in the FTIR spectrum of the products indicating the complete reduction azo group (-N = N-) of Congo Red. 1H NMR (400 MHz, DMSO) results of obtained solid product have shown δ at 7.22 (4H, d), 6.61 (4H, d), 6.56 (4H, s), 5.20 (3H, s), 3.47 (6H, s), 2.50 (1H, s) (Fig. S 4). The 13C NMR (100MHz, DMSO) has shown δ values at 146.62 (s), 129.51 (s), 126.53 (s), 115.18 (s) (Fig. S 5). From LC-MS (ESI-TOF) (Fig. 7), the major peaks have m/z values at 185.4 (ppm error − 2.04) which has corresponding molecular formula C12H12N2. Based on the above data, the major reduced product is found as benzidine; (M + H) + on 432 second of reaction time. In the HRMS result, some minor reduction products also found at m/z values 168, 152, 134, and 113 which are correspond to aminobiphenyl, biphenyl, 4-carboxybutanoate and hydroxy hexadienolate, respectively. The apparent rate of catalytic reduction of Congo red using SN@CuO NPs (Stage-1) has been deduced. The absorbance versus time graph was plotted and the kinetic Eq. (2) has been used to calculate the magnitude of apparent rate constant.
$$ln\frac{{A}_{t}}{{A}_{0 }}=ln\frac{{C}_{t}}{{C}_{0 }}= -kt$$
2
where k signifies the pseudo-first-order reaction rate constant. Based on data fitting (R2 = 0.9720) the rate of reduction of azo group (absorption peak at 495 nm) of Congo red is found to be 0.468 min− 1.
After complete disappearance of UV-visible peak at 495 nm, the reaction continued for elongated time for complete disappearance of remaining peaks of UV-visible spectrum of Congo red at 350, 285 and 247 nm and degradation has been monitored by recording UV-visible absorbance of reaction mixture as function of time. Figure 8a shows that all remaining peaks; at 247, 285 and 350 nm have completely disappeared after additional 300 minutes of reaction time (Stage-2). The rate of degradation of 285, 247 and 350 nm UV-visible peaks (Stage-2) has been determined using Eq. (2). The data fitting coefficient R2 value found as 0.9931 and the apparent rate constant for the degradation of above peaks has been determined as 0.0189 min− 1 (Fig. 8b).
In the TLC measurement of resulting reaction mixture, no spot has been detected. The SN@CuO NPs has been separated from reaction mixture by centrifugation. The adsorbed degradation products at the SN@CuO NPs surface have been extracted in acetonitrile and LC-MS of extract has been recorded. In the LC-MS spectrum (Fig. 9) the major peaks observed at m/z value of 145, along with few minor peaks at m/z values 116, 127, 156, and 223 which are correspond to α-naphthol, succinic acid, biphenyl, naphthyldiazene, and naphthionic acid, respectively.
From above results it is found that the as produced Solanum nigrum mediated CuO nanoparticles are efficient catalyst for degradation of Congo red. In the first stage, only 5 mg of nanocatalyst enabled to reduce azo group, present in 10 mL of 1.0 mM dye solution using of 10 mL of 0.30 M NaBH4 solution (Fig. 6) within 7.2 minutes. In the second stage, reaction mixture (reduced products of Congo red) has completely been free from Congo red dye content within 300 minutes. Thus the reaction mixture become completely free from dye content within total 307.2 minutes of time (Fig. 6,8). Catalytic efficacy of SN@CuO NPs in this study has been compared with that of some latest reported nanocatalysts (Chen et al. 2014, Indana et al. 2016, Jayapriya &Arulmozhi 2021, Narasaiah &Mandal 2020, Singh et al. 2017, Subair et al. 2016, Umamaheswari et al. 2018). From Table 1 it is obvious that SN@CuO NPs is better catalyst for degradation of Congo red. In these reports costly materials and hazardous chemicals have been used to prepare catalysts while in present study, less expensive SN@CuO NPs synthesized via sustainable route to degrade Congo red rapidly than above catalysts.
Table 1
S.No.
|
Catalyst
|
Reducing Agent
|
Time (minutes)
|
Rate constant (min− 1)
|
Reference
|
1
|
Cubic Gold Nanorattles
|
NaBH4
|
15 min
|
0.362
|
Chen et al., 2014
|
2
|
Au NPs-Dp-p(EI) membrane
|
NaBH4
|
29
|
0.209
|
Indana et al., 2016
|
3
|
Fe3O4@PANI@Au
|
NaBH4
|
20
|
--------
|
Jayapriya and Arulmozhi, 2021
|
4
|
Ag NPs
|
NaBH4
|
15 min
|
0.148
|
Narasaiah and Mandal, 2020
|
5
|
Ag/TiO2
|
NaBH4
|
20 min
|
--------
|
Singh et al., 2017
|
6
|
Au NPs
|
NaBH4
|
10 min
|
0.27
|
Subair et al., 2016
|
7
|
Pd Nps
|
NaBH4
|
14 min
|
0.2387
|
Umamaheswari et al., 2018
|
8
|
SN@CuO NPs
|
NaBH4
|
7.2min
|
0.526
|
Present work
|
It is well established that for catalytic reduction of congo red, both reactants, Congo red and BH4− ion, must be adsorbed onto the surface of the particles, which in turn can react, and the products dissociate from the surface of the catalyst. Without SN@CuO NPs, bare NaBH4 could not reduce Congo red even after eight hour of reaction time while in presence of only 5 mg SN@CuO NPs, Congo red solution completely mineralized within 307.2 minutes. The LC-MS results have shown that degraded products are adsorbed on the surface of CuO nanoparticles which indicates that reaction has taken place at the surface of SN@CuO NPs. Several factors, such as porosity/surface area, pH of reaction mixture and electrostatic attraction between adsorbate dye and adsorbent SN@CuO NPs are responsible for adsorption of Congo red at the surface of SN@CuO NPs (Labanda et al. 2009). Congo red is a dipolar molecule which exists in anionic form in alkaline medium having deep red color, while in acidic medium protonation of azo group takes place, resulting appearance of deep blue color of dye solution. In presence NaBH4 solution the pH of reaction mixture become 10, thus Congo red exists in anionic form. TEM and BET results revealed that SN@CuO NPs have very small size, large surface area and mesoporous structure while zeta potential value has shown high positive charge at the surface of SN@CuO NPs. Due to positive charge on SN@CuO NPs surface, the reactants; negatively charged Congo red and BH4 − 1 attracted to positively charged SN@CuO NPs and adsorbed on free catalytic sites, which are present in large number. Rajesh et al. reported that borohydride ions (BH4 − 1) generated from the dissociation of NaBH4 and molecules of congo red readily diffused from an aqueous reaction solution to the surface of nanoparticles occupying active sites (Rajesh et al. 2014). During reduction of Congo red, nanoparticles work as a relay center for transfer of an electron between borohydride ions (BH4 − 1) ions and congo red molecules in which borohydride ions (BH4 − 1) act as electron donor whereas congo red molecules act as electron acceptors. It is supposed that BH4− interacted with SN@CuO NPs to form Cu tetrahydroborates dihydrogen bonded (DHB) tetrahedral L2Cu(ɳ2-BH4) complexes. The DHB complexes involving both bridging and terminal hydride hydrogens are proved to be effective reducing agents (Belkova et al. 2017). Further, FTIR spectrum of leaf extract mediated CuO nanoparticles (Fig. 2b) has shown that interaction of secondary amides of Solanum nigrum leaf extract with Cu ions take place through C = O (Cu-C = O), leaving –NHR of secondary amide free for interaction with sulphonate group (CR-SO3−) of Congo red. Now congo red molecules transfer electrons to nanoparticles, which activate azo bonds of congo red molecules, whereas borohydride ions, in the form of L2Cu(ɳ2-BH4) complexes, act as dihydrogen sources. Hydrogen released from L2Cu(ɳ2-BH4) complexes attack on activated azo bonds. Due to conversion of (–N = N–) to (–HN–NH–) and finally breakage of (–HN-NH-) bond due to uptake of dihydrogen through nano-catalyst (L2Cu(ɳ2-BH4) complexes), congo red starts fading rapidly and becomes colorless after completion of reduction reaction within 432 seconds (Jia et al. 2014). The plausible mechanism for the degradation of Congo red has been presented in Scheme I.
After reduction of -N = N- group within 432 seconds of reaction time, benzidine is the main degradation product in this experiment, however, some other degradation products have also been detected in LC-MS result (Fig. 7, Scheme-I). When reaction further proceeded for an elongated time 300 minutes, then the reaction mixture become completely free from Congo red and its degradation products, like benzidine and other small fractions, and the UV-visible peaks at 285, 247 and 350 nm completely disappeared. However, degraded products, which are mainly α-naphthol and some other molecules in small fraction, are still found to be adsorbed on the surface of SN@CuO NPs. This indicates that SN@CuO NPs act as support and provide suitable medium and works as relay center for movement of electrons between reactants. Due to this interplay of electrons, degradation products (after disappearance of azo group), like benzidine (major) and other molecules (minor), degraded to final products α-naphthol (major) and some other molecules (minor) in small fraction, present at the surface of SN@CuO NPsin adsorbed form (Scheme-II).