Munisha Mahajana,
Sanjeev Kumara,
Jyoti Gaurb,
Sandeep Kaushal*c,
Jasvir Dalal
*d,
Gurjinder Singhe,
Mrinmoy Misra
*f and
Dharamvir Singh Ahlawat
g
aDepartment of Physics, Chandigarh University, Gharuan, Mohali, 140413, India
bSchool of Basic and Applied Sciences, RIMT University, Mandi Gobindgarh, 147301, India
cRegional Institute of Education, NCERT, Ajmer, Rajasthan 305004, India. E-mail: kaushalsandeep33@gmail.com
dDepartment of Physics, Rajdhani College, University of Delhi, Delhi, 110015, India. E-mail: jasvirdalal2012@gmail.com
eDepartment of Electrical and Electronics and Communication Engineering, DIT University, Dehradun, 248009, India
fMechatronics Engineering Department, School of Automobile, Mechanical and Mechatronics, Manipal University, Jaipur, India. E-mail: mrinmoy.mishra@jaipur.manipal.edu
gDepartment of Physics, Chaudhary Devi Lal University, Sirsa, Haryana 125055, India
First published on 29th January 2025
Achieving the smallest crystallite/particle size of zinc oxide nanoparticles (ZnO NPs) reported to date, measuring 5.2/12.41 nm with Justicia adhatoda (J. adhatoda) leaf extract, this study introduces a facile green synthesis. Utilizing aqueous J. adhatoda leaf extract as both a reducing and stabilizing agent, the method leverages the plant's rich phytochemical composition to produce highly crystalline and morphologically controlled ZnO NPs. This precise particle size control highlights the effectiveness of the synthesis process in morphological tuning. The synthesized NPs were thoroughly characterized using XRD, UV-vis spectroscopy, FTIR, FESEM, and HRTEM, which collectively revealed superior crystallinity, controlled morphology, and unique surface properties conferred by phytochemical bio-capping. The photocatalytic performance of these biogenic ZnO NPs was evaluated for the degradation of two model pollutants: malachite green (MG), a synthetic dye, and 4-nitrophenol (4-NP), a toxic organic compound. The NPs exhibited exceptional photocatalytic efficiency, achieving 99.8% MG degradation within 180 minutes and demonstrating a rapid photocatalytic reduction of 4-NP to 4-aminophenol with a reaction rate constant of 0.245 min−1 under UV and sunlight irradiation. Mechanistic studies attributed this high performance to reactive oxygen species (ROS) generation and electron–hole pair interactions, supported by improved charge separation and high surface area. This work not only establishes the potential of J. adhatoda-mediated ZnO NPs in addressing persistent organic pollutants but also sets a benchmark for size-controlled NPs synthesis. By delivering scalable and eco-friendly water remediation technologies, this study advances green nanotechnology.
Extensive advancements in nanoscience and nanotechnology have driven the development of innovative materials and advanced processes for water purification, particularly targeting the removal of toxic dyes from industrial effluents. Materials such as titanium dioxide (TiO2), zinc oxide (ZnO), graphitic carbon nitride (g-C3N4), and bismuth vanadate (BiVO4) have demonstrated outstanding photocatalytic properties, making them leading candidates in wastewater treatment.9 Advanced oxidation processes (AOPs), which include ozonation, sonochemical degradation, electrochemical remediation, heterogeneous photocatalysis, and Fenton oxidation, have emerged as transformative techniques due to their ability to mineralize organic pollutants into environmentally safe byproducts like carbon dioxide and water.10 These processes operate by generating reactive oxygen species (ROS), including hydroxyl radicals (˙OH) and superoxide anions (O2−˙), which effectively break down complex dye molecules.11
ZnO exhibits several exceptional characteristics that distinguish it from other top-performing photocatalytic materials such as TiO2, g-C3N4, BiVO4, ZnS, and CeO2.12 With a wide direct bandgap of ∼3.37 eV and a high exciton binding energy of 60 meV, ZnO ensures efficient charge carrier generation and prolonged exciton stability, outperforming TiO2 (∼4 meV), BiVO4 (∼20 meV), g-C3N4 (∼30 meV), ZnS (∼28 meV), and CeO2 (∼19 meV) in minimizing electron–hole recombination.13,14 This property is critical for maintaining photocatalytic activity under UV-light irradiation. Moreover, ZnO's electron mobility, ranging from ∼115–155 cm2 V−1 s−1, is significantly higher than that of TiO2 (∼0.1–4 cm2 V−1 s−1), BiVO4 (∼0.01–0.1 cm2 V−1 s−1), g-C3N4 (∼0.1–1 cm2 V−1 s−1), ZnS (∼5–20 cm2 V−1 s−1), and CeO2 (∼4–6 cm2 V−1 s−1).15 This high electron mobility facilitates faster charge transport and supports ROS generation during photocatalytic reactions. In addition to its optical and electronic properties, ZnO demonstrates remarkable thermal and chemical stability, remaining stable up to ∼1975 °C and across a wide pH range. This durability contrasts with materials prone to photo-corrosion or limited aqueous stability, such as ZnS and BiVO4.16 While CeO2 excels in oxygen storage capacity and redox properties, its relatively narrow bandgap (∼3.2 eV) can lead to undesirable secondary reactions, making ZnO a more efficient choice for photodegradation applications.17
In recent years, green synthesis has emerged as a cost-effective and environmentally friendly approach for developing advanced nanomaterials, particularly for applications such as photocatalytic wastewater treatment.18 By utilizing bioactive phytochemicals present in plant extracts, this method eliminates the need for toxic chemicals and energy-intensive processes, making it a sustainable alternative to conventional techniques like sol–gel, co-precipitation, and chemical vapor deposition.19–21 Among the numerous plant-based resources explored, J. adhatoda stands out as a promising candidate for NPs synthesis due to its rich phytochemical composition and medicinal properties.22 The extract of J. adhatoda contains a high concentration of alkaloids (e.g., vasicine and vasicinone), flavonoids (3.5 mg g−1 dry weight), polyphenols (42 mg g−1 gallic acid equivalent), terpenoids, and tannins, which act as reducing, stabilizing, and capping agents during the synthesis of ZnO NPs.23,24 These bioactive compounds facilitate the reduction of zinc ions (Zn2+) into ZnO, while the capping effect of the phytochemicals ensures particle stabilization and prevents aggregation. The hydroxyl and carboxyl groups of phenolic compounds present in the extract play a pivotal role in metal ion reduction, donating electrons to convert zinc ions into ZnO while simultaneously stabilizing the NPs through strong interactions.25,26 The pH-adjusted synthesis process, typically conducted at pH 8, ensures optimal nucleation and growth, yielding ZnO NPs with an average size of 12.4 nm as confirmed by microscopy analysis.27 Like studies on other plants such as Moringa oleifera, Linum usitatissimum, and Salvadora persica, the extract of J. adhatoda demonstrates the capability to produce high-quality ZnO NPs with significant photocatalytic potential. The plant's phytochemical diversity not only supports the synthesis process but also offers opportunities for tuning NPs properties to meet specific application requirements, further emphasizing its relevance in sustainable nanotechnology.28,29
This study is centered on the development of biogenic ZnO NPs synthesized using J. adhatoda, with a focus on their application in the photocatalytic degradation of malachite green (MG) and the photocatalytic reduction of 4-nitrophenol (4-NP). The novelty of this work lies in the dual evaluation of ZnO NPs for addressing distinct classes of pollutants—an industrial dye and a toxic organic compound—under UV-light irradiation. By employing a sustainable synthesis approach, this research highlights the potential of J. adhatoda-mediated ZnO NPs to integrate eco-friendly nanotechnology with effective wastewater treatment solutions. The ability of the synthesized ZnO NPs to achieve high degradation efficiency for both pollutants emphasizes their versatility and efficacy. Additionally, this study provides a comprehensive mechanistic understanding of the photocatalytic processes involved, shedding light on the interplay between NPs properties and pollutant degradation pathways. The work emphasizes the critical role of particle size, crystallinity, and reactive site density in enhancing photocatalytic performance. By demonstrating the successful degradation of challenging pollutants like MG and the photocatalytic reduction of 4-NP, this study sets a precedent for future research into green nanotechnology-based solutions for environmental remediation, offering a pathway toward sustainable and scalable water purification technologies.
The progress of the extraction was monitored using UV-visible spectroscopy, which confirmed the stability of the phytochemical content throughout the process. The consistent absorbance spectrum of the extract indicated effective phytochemical dissolution. The pH of the extract was measured at room temperature and found to be approximately 5. The freshly prepared extract was stored in an airtight container at 4 °C to preserve its chemical stability and bioactivity. This extract was subsequently used as a reducing and capping agent for the green synthesis of ZnO NPs.
The reaction mixture was allowed to cool to room temperature and left undisturbed for 12 hours to ensure complete precipitation of the NPs. The resulting precipitate was separated by filtration using Whatman filter paper with a pore size of 1.2 μm. It was then washed three times with deionized water and twice with ethanol to remove unreacted precursors and residual phytochemicals. After each washing step, the solution was centrifuged at 10000 rpm for 10 minutes to recover the purified NPs. The washed ZnO NPs were dried in a hot air oven at 80 °C for 12 hours to yield a fine powder. Approximately 400 mg of ZnO NPs were obtained, corresponding to a yield of 8 mg NPs mL−1 of extract used. The green synthesis of ZnO NPs involves several key steps. Initially, zinc acetate dissociates in the aqueous medium, releasing zinc ions (Zn2+), which act as precursors for NPs formation. Next, bioactive compounds such as flavonoids, alkaloids, and terpenoids, present in the plant extract, act as reducing agents. These compounds donate electrons to reduce zinc ions, facilitating the formation of ZnO NPs. During the process, hydroxyl ions supplied by the aqueous medium and phytochemical content form an intermediate compound, zinc hydroxide. Upon heating, zinc hydroxide undergoes dehydration, resulting in the formation of ZnO NPs. The bioactive compounds also stabilize the NPs, preventing agglomeration and ensuring uniform size distribution.
Aliquots of 5 mL were withdrawn at regular time intervals (20, 40, 60, 80… minutes), followed by centrifugation to separate the NPs. The concentration of MG dye in the clear supernatant was analyzed using a UV-visible spectrophotometer by measuring the absorbance at 617 nm, the characteristic wavelength of MG dye. The degradation efficiency was calculated by comparing the initial concentration of MG dye at the start of the reaction (time 0) with the concentration at a specific time interval during the process.14,30 The percentage of degradation was determined based on the reduction in dye concentration over time.
The photocatalytic activity of ZnO NPs was evaluated by studying the reduction of 4-nitrophenol (4-NP) in the presence of sodium borohydride (NaBH4). A reaction mixture was prepared by dissolving 4 mg of 4-NP in 100 mL of deionized water to obtain a 40 ppm solution. Subsequently, 1 mL of freshly prepared NaBH4 solution (0.1 M) and 50 mg of ZnO NPs were added. The mixture was stirred gently to ensure uniform dispersion of the catalyst. The reaction mixture was placed in a quartz reactor and exposed to natural sunlight, with an intensity ranging from 800–1000 W m−2. The reactor was positioned at an angle to maximize light absorption, and the temperature of the reaction system was maintained at 30 °C.
The progress of the photocatalytic reduction of 4-NP was monitored using a UV-vis spectrophotometer. Aliquots (3 mL) of the reaction mixture were withdrawn at specific intervals and immediately centrifuged at 10000 rpm for 10 minutes to separate the NPs. The absorbance of the clear supernatant was measured in the wavelength range of 250–550 nm.
Sr. no. | Synthesis method | Extract used | Crystallite size (nm) | Crystal structure | Key findings | References |
---|---|---|---|---|---|---|
1 | Leaf extract-mediated green synthesis | Dolichos lablab | 29 | Hexagonal wurtzite | Demonstrated photocatalytic and bactericidal properties with phase-pure ZnO | 31 |
2 | Plant-mediated synthesis | Hibiscus subdariffa | 12–18 | Hexagonal wurtzite | Highlighted temperature-dependent synthesis | 32 |
3 | Aqueous extract-mediated synthesis | Punica granatum | 10–45 | Spherical, well-ordered | Reported crystallite size with antimicrobial and catalytic efficacy | 33 |
4 | Plant-mediated synthesis | Thyme leaves | 35.2–243.3 | Hexagonal wurtzite | Emphasized calcination effects | 34 |
5 | Plant-mediated synthesis | Hibiscus cannabinus | 15–20 | Hexagonal wurtzite | Exhibited potential in wastewater purification and antibacterial activity | 35 |
6 | Microbial-mediated synthesis | Streptomyces baarnensis | <12 | Crystalline structure | Enhanced antibacterial efficacy using microbial-derived ZnO | 36 |
7 | Fruit extract-mediated synthesis | Myristica fragrans | 41.23 | Hexagonal wurtzite | Highlighted antifungal and photocatalytic applications | 37 |
8 | Plant-mediated synthesis | Polyalthia longifolia | Varied | Hexagonal wurtzite | Validated diverse synthesis conditions | 38 |
9 | Biomediated synthesis | J. adhatoda | 5.2 (avg.) | Hexagonal wurtzite | Achieved smallest average crystallite size, strong c-axis orientation, and minimal lattice strain | Current study |
The interplanar spacing (d-spacing) was calculated using Bragg's law:
nλ = 2d![]() ![]() | (1) |
The average crystallite size (D) of the ZnO NPs was determined using the Debye–Scherrer equation:39
D = Kλ/β![]() ![]() | (2) |
Sr. no | 2θ (degree) | FWHM (β) (degree) | Lattice planes (hkl) | Inter-planner spacing (d) (Å) | Crystallite size (D) (nm) |
---|---|---|---|---|---|
1 | 31.93 | 1.22809 | (100) | 2.80082 | 7.03 |
2 | 34.46 | 2.29841 | (002) | 2.60050 | 3.78 |
3 | 36.24 | 1.77468 | (101) | 2.47692 | 4.92 |
4 | 47.54 | 2.88408 | (102) | 1.91125 | 3.14 |
5 | 56.79 | 1.26491 | (110) | 1.61979 | 7.46 |
6 | 62.88 | 2.67535 | (103) | 1.47681 | 3.64 |
7 | 68.09 | 2.06760 | (112) | 1.37594 | 4.84 |
8 | 77.96 | 1.45530 | (004) | 1.22459 | 7.33 |
D (average) = 5.2 nm |
To further analyze the effects of size and strain on peak broadening, Williamson–Hall (W–H) analysis was employed. The W–H method separates the size-induced and strain-induced broadening contributions based on the following equation:40
β![]() ![]() ![]() ![]() | (3) |
The ZnO NPs synthesized in this study using J. adhatoda extract exhibits an average crystallite size of 5.2 nm, which is notably smaller than those reported in similar studies (Table 1). A study provides insights into the green synthesis of ZnO NPs using Dolichos lablab leaf extract via a leaf extract-mediated approach. The synthesized NPs exhibited a crystallite size of approximately 29 nm and a hexagonal wurtzite structure. The study demonstrated their excellent photocatalytic activity and bactericidal properties.31 In another research, Hibiscus subdariffa was used for the plant-mediated synthesis of ZnO NPs, resulting in crystallite sizes ranging from 12 to 18 nm with a hexagonal wurtzite crystal structure. This study emphasized the temperature-dependent synthesis process.32 Another study highlights the use of an aqueous extract of Punica granatum to synthesize ZnO NPs and crystallite sizes between 10 and 45 nm. The research reported significant antimicrobial and catalytic efficacy of the NPs, demonstrating their potential in environmental and biomedical applications.33 Research using thyme leaves for the green synthesis of ZnO NPs emphasized the role of calcination temperature in determining their structural properties. The crystallite sizes varied from 35.2 to 243.3 nm, with a hexagonal wurtzite crystal structure.34 A study using Hibiscus cannabinus for plant-mediated synthesis produced ZnO NPs with a crystallite size ranging from 15 to 20 nm and a hexagonal wurtzite structure.35 In microbial-mediated synthesis, Streptomyces baarnensis was used to produce ZnO NPs with crystallite sizes below 12 nm.36 Another study utilized Myristica fragrans fruit extract for the synthesis of ZnO NPs, resulting in a crystallite size of 41.23 nm with a hexagonal wurtzite structure. This research demonstrated the NPs' antifungal and photocatalytic properties.37 In another research, Polyalthia longifolia was employed for plant-mediated synthesis, resulting in ZnO NPs with varied crystallite sizes and a hexagonal wurtzite structure.38 Although their study achieved moderate crystallinity and smaller sizes than most chemical methods, the particles were still larger than those produced in the current work.
The reduced crystallite size of the ZnO NPs can be attributed to the bio-template role of the J. adhatoda extract, which contains alkaloids, flavonoids, and other phytochemicals. These biomolecules act as capping and stabilizing agents, preventing particle agglomeration and promoting the formation of well-crystallized, nanoscale particles. Furthermore, the use of J. adhatoda extract enhances the ecological sustainability of the synthesis process by eliminating the need for toxic chemical stabilizers.
Notable structural differences are observed between the estimated and standard lattice values, likely caused by synthesis-induced effects. While the c-axis value (5.2 Å) closely matches the standard (5.2066 Å), indicating intact symmetry along this direction, the computed a = b value (4.8 Å) is significantly larger than the standard (3.2498 Å), suggesting probable lattice expansion due to strain, defects, or doping (Table 3). Minimal hexagonal distortion is evidenced by the optimal c/a ratio of 1.0, which reinforces the structural stability of the wurtzite phase. The unit cell volume, calculated at 192.3 Å3 compared to the standard 47.62 Å3, indicates significant volumetric expansion, likely due to doping or lattice strain. These variations underscore the impact of synthesis parameters on structural characteristics, which can influence material performance in applications such as optoelectronics and photocatalysis.
Lattice constants | Standard values (Å) | Calculated values (Å) | c/a ratio | Standard volume (Å)3 | Calculated volume (Å)3 |
---|---|---|---|---|---|
a = b | 3.2498 | 4.8 | 1.0 | 47.62 | 192.3 |
c | 5.2066 | 5.2 |
Thus, the XRD and W–H analyses confirm the successful synthesis of hexagonal wurtzite ZnO NPs with nanoscale dimensions, strain effects, and high crystallinity. The synergistic effect of phytochemical-mediated synthesis ensures the controlled growth of ZnO NPs with superior structural and morphological properties, making them suitable for advanced applications in catalysis, optoelectronics, and biomedicine.
The UV absorption observed in this study aligns with findings from prior green synthesis research. The comparative analysis presented in Table 4 highlights the superior optical properties of ZnO NPs synthesized using J. adhatoda extract in this study. With an absorption peak at 353 nm and a bandgap energy of 3.09 eV, the synthesized NPs achieve an optimal balance between bandgap tuning and quantum confinement effects, surpassing several prior reports. For instance, while Pandiyan (2019) reported a bandgap of 3.11 eV for J. adhatoda-derived ZnO NPs, the present study achieves a slightly reduced bandgap, indicating enhanced optical efficiency. Similarly, compared to NPs derived from Moringa oleifera (3.08 eV) and Aloe vera (3.10 eV), the current synthesis process yields ZnO NPs with tailored bandgap energies, making them more suitable for applications requiring precise UV absorption and efficient charge carrier generation.
Order | Plant extract used | Absorption peak (nm) | Bandgap energy (eV) | Reference |
---|---|---|---|---|
1 | Justicia adhatoda | 355 | 3.11 | 42 |
2 | Moringa oleifera | 358 | 3.08 | 43 |
3 | Aloe vera | 200–450 | 3.10 | 44 |
4 | Azadirachta indica | 200–450 | 3.12 | |
5 | Amaranthus dubius | 200–450 | 3.07 | |
6 | Calendula officinalis | 355 and 370 | 2.986 | 45 |
7 | Algerian date syrup | NA | 3.19 | 46 |
8 | Eucalyptus globulus | 464 | 2.67 | 47 |
9 | Allium cepa (onion bulb) | 445 | 2.66–2.79 | 48 |
10 | Mimosa pudica | 300 | 3.50 | 49 |
11 | J. adhatoda | 353 | 3.09 | Present study |
In contrast, plant-mediated syntheses using Eucalyptus globulus (2.67 eV) and Allium cepa (2.66–2.79 eV) exhibit significantly lower bandgap energies, limiting their suitability for high-energy applications. Meanwhile, Mimosa pudica-derived NPs, with a bandgap of 3.50 eV, display excessive quantum confinement, reducing their absorption range. These results underscore the superiority of the J. adhatoda-mediated synthesis presented here, achieving a desirable combination of a blue-shifted absorption peak and a moderately reduced bandgap. This makes the synthesized ZnO NPs particularly suitable for advanced optoelectronic, photocatalytic, and UV-blocking applications. The findings emphasize the effectiveness of the synthesis method employed in this study in producing highly crystalline, uniform ZnO NPs with enhanced optical properties.
The analysis of Fig. 3(b) was conducted to gain deeper insights into the optical properties of the ZnO NPs. Tauc plot, which correlates optical absorbance data with photon energy (hν), was used to determine the bandgap energy (Eg). By plotting (αhν)2 as a function of hν and extrapolating the linear portion of the curve to the x-axis, the bandgap energy of the synthesized ZnO NPs was calculated to be 3.09 eV, smaller than the bulk ZnO bandgap (3.37 eV). This reduction can be attributed to the influence of surface defects, phytochemical interactions, and mild synthesis conditions. Surface-bound phytochemicals may introduce localized energy states within the bandgap, effectively narrowing the measured bandgap. Additionally, defect states, such as oxygen vacancies and particle aggregation, could further contribute to this deviation from quantum confinement effects. This behaviour aligns with prior studies of green-synthesized ZnO, which highlight the complexity of interpreting optical properties in phytochemically mediated syntheses.
The decreased bandgap energy is advantageous for applications requiring higher UV absorption, such as photocatalytic degradation of organic pollutants, optoelectronic devices, and UV-blocking materials.50 This bandgap enhancement ensures better utilization of UV light for generating charge carriers, a critical feature for improving the efficiency of photocatalysts.
Therefore, the results from Fig. 3(a) and (b) validate that the green synthesis of ZnO NPs using J. adhatoda extract produces NPs with tailored optical properties, including a blue-shifted absorption edge and an enhanced bandgap energy. These findings highlight the influence of the phytochemical-mediated synthesis process in achieving nanoscale effects, rendering the material suitable for advanced applications in environmental remediation, optoelectronics, and antimicrobial treatments.
The FTIR spectrum of the J. adhatoda/ZnO NPs (Fig. 4(b)) reveals significant shifts and the emergence of new peaks, indicating the successful formation of NPs and the interaction between phytochemicals and the ZnO surface. The stretching vibrations of the –OH and –NH groups of polyols and amides are responsible for a high absorption at 3451 cm−1.54 The peak at 2354 cm−1 represents CN stretching, while the bands at 1695 cm−1 and 1551 cm−1 are attributed to C
O stretching and N–H bending, respectively. Notably, the sharp peak at 1028 cm−1 corresponds to the Zn–O stretching vibration, confirming the formation of ZnO NPs. The peak at 668 cm−1 further corroborates the Zn–O bond, indicating the successful synthesis of ZnO NPs.55
The shifts in peak positions and changes in intensity between the spectra of the extract and the ZnO NPs demonstrate the active involvement of phytochemicals in reducing zinc ions to ZnO and stabilizing the NPs.56 These findings align with reports in the literature, where bioactive compounds such as alkaloids and flavonoids have been shown to act as natural reducing and capping agents, ensuring the stability and uniformity of NPs. The use of J. adhatoda in green synthesis provides a biocompatible and eco-friendly approach, with potential applications in photocatalysis, antimicrobial activity, and environmental remediation.57 This FTIR analysis highlights the role of phytochemicals in mediating NPs synthesis and highlights the functional integrity of the synthesized ZnO NPs.
![]() | ||
Fig. 5 SEM micrographs of J. adhatoda@ZnO NPs at varying magnifications, highlighting their agglomerated nanostructure (a–e), weight and atomic percentages of Zn and O (f), EDX analysis (g). |
Fig. 5(a) presents a low-magnification FESEM micrograph, showing the agglomeration of ZnO NPs into irregularly shaped clusters. This aggregation, characteristic of green-synthesized ZnO NPs, is primarily driven by van der Waals forces and hydrogen bonding between phytochemicals and NPs during synthesis. The hierarchical assembly of ZnO NPs suggests strong interactions between individual particles, enhancing surface area and porosity. Unlike conventional ZnO synthesis, the biogenic approach introduces organic capping agents that prevent uncontrolled crystal growth while facilitating the formation of functional microstructures. Fig. 5(b) provides a higher magnification view (60k×) of the ZnO NPs, revealing rough surface textures and smaller sub-clusters within the aggregated structures. These sub-clusters range in size from 50 to 100 nm. The observed roughness suggests potential for enhanced light scattering and interaction with reactants in photocatalytic applications. This level of porosity and roughness, rarely achieved through purely chemical synthesis, demonstrates the role of bioactive compounds from J. adhatoda in shaping NPs surfaces. The intricate microstructure supports efficient charge carrier transfer during photocatalytic processes, reducing recombination losses.
Fig. 5(c), captured at 75k× magnification, highlights individual NPs with well-defined boundaries within the agglomerates. The observed size variation (25–200 nm) reflects the dynamic nature of the biogenic synthesis process. The presence of smaller particles within larger aggregates suggests a dual nucleation mechanism, wherein some particles grow independently while others coalesce into larger clusters. This distribution contributes to the hierarchical architecture of the nanostructures, offering multiple reactive sites. Such hierarchical structuring, a hallmark of green synthesis methods, is advantageous for multifunctional applications like photocatalysis and antimicrobial treatments. Fig. 5(d) provides a close-up view (75k× magnification) of the ZnO NPs, emphasizing the quasi-spherical shape of individual particles. These particles exhibit pronounced surface irregularities, likely arising from the interactions of phytochemical capping agents during synthesis. Such irregularities enhance the surface reactivity of the NPs, creating active sites for catalytic reactions. Unlike traditional methods that produce uniform, smooth particles, the bio-mediated approach introduces morphological diversity, critical for improving adsorption and pollutant reactivity.
Fig. 5(e) reveals sub-25 nm particles dispersed within larger agglomerates. These ultra-fine particles significantly increase the material's overall surface area. The phytochemical capping agents from J. adhatoda stabilize these smaller particles, preventing their growth into larger crystals. These fine particles provide additional active sites for light absorption and ROS generation, enhancing photocatalytic performance.
Fig. 5(f) presents the elemental composition of the synthesized ZnO NPs, as determined by EDX spectroscopy. The weight percentages of Zn (86.78%) and O (13.22%) confirm the stoichiometric formation of ZnO. The Zn:
O atomic percentage ratio (61.64
:
38.36) indicates high purity, with minimal contamination from other elements. This purity underscores the efficiency of the J. adhatoda-mediated synthesis process, which uses bioactive compounds to reduce zinc ions while preventing the incorporation of impurities. Fig. 5(g) provides the EDX spectrum, where the dominant Zn and O peaks corroborate the elemental composition results. The absence of additional peaks confirms the lack of significant impurities, ensuring the suitability of the ZnO NPs for sensitive applications such as biomedical devices and environmental remediation. The presence of multiple Zn peaks suggests the material's crystallinity, while the O peak verifies the formation of ZnO bonds.
The SEM and EDX analyses collectively demonstrate the structural and compositional advantages of J. adhatoda-mediated ZnO NPs. The hierarchical nanostructures, high purity, and unique surface features underscore their potential for advanced applications, including photocatalysis, antimicrobial activity, and environmental remediation. This bio-mediated synthesis method ensures eco-friendly production while introducing morphological features rarely achieved through conventional approaches, emphasizing its superiority for nanomaterial development.
Although some agglomeration is observed, individual NPs remain discernible within the clusters, indicating effective stabilization by bioorganic compounds in the J. adhatoda extract. The phytochemicals, identified through FTIR as hydroxyl (–OH), amine (–NH), and carbonyl (CO) groups, not only reduce zinc ions but also cap the particles, minimizing excessive aggregation. Sub-clusters within the agglomerates (Fig. 6(a)) reflect weak van der Waals interactions, preserving surface-active sites critical for catalytic applications. This behaviour demonstrates the fine balance between particle stabilization and retention of functional properties. Fig. 6(b) highlights the uniform size distribution of the NPs, with regions marked by dashed circles and squares emphasizing consistent particle shapes and sizes. This uniformity is quantified in the histogram (Fig. 6(e)), which reveals particle sizes ranging from 10 to 18 nm, with an average size of 12.41 nm and a narrow standard deviation (SD) of 0.19. The narrow distribution reflects the reproducibility of the synthesis process and the phytochemicals' ability to regulate particle growth. The slightly larger particle size observed in HRTEM compared to the crystallite size (∼5.2 nm) determined by XRD is attributed to the controlled clustering of primary crystallites into secondary particles. This phenomenon, far from being a limitation, ensures enhanced stability and tuneable optical properties. The phytochemical capping, confirmed by FTIR, is instrumental in achieving this controlled clustering, making the NPs advantageous for photocatalytic applications.
Fig. 6(c) provides a high-magnification view of the lattice fringes, with an interplanar spacing (d-spacing) of 0.36 nm corresponding to the (101) plane of hexagonal wurtzite ZnO (JCPDS 00-021-1272). The well-defined and continuous lattice fringes reflect excellent crystallinity and minimal structural defects. This high degree of crystalline order, facilitated by bioactive compounds during nucleation, ensures efficient charge transfer and optical performance-key for applications in optoelectronics and photocatalysis. The uniform alignment of atoms in the lattice fringes suggests strain-free growth, likely due to selective interactions between phytochemicals and zinc ions during synthesis. Anisotropic growth patterns observed in the lattice fringes indicate preferential development along specific crystallographic planes, enhancing the NPs' active surface areas.
The SAED pattern (Fig. 6(d)) confirms the polycrystalline nature of the ZnO NPs. The well-defined concentric rings are indexed to the (100), (002), (101), (102), (110), (103), and (112) planes of ZnO, characteristic of the hexagonal wurtzite structure. Bright diffraction spots superimposed on these rings signify high crystalline quality and the presence of multiple crystal orientations, enhancing the optical and catalytic properties of the NPs. The slight elongation of some diffraction spots suggests localized strain effects within the crystalline structure. Rather than a drawback, such strain can introduce beneficial surface defects, enhancing reactivity in photocatalytic and antimicrobial applications. These unique crystallographic features directly result from the controlled growth facilitated by the bioactive compounds in J. adhatoda extract.
The HRTEM analysis highlights the superior morphological and structural features of ZnO NPs synthesized using J. adhatoda. The high crystallinity, uniform size distribution, polycrystalline nature, and unique growth patterns underscore the effectiveness of the green synthesis method. The combined insights from TEM and FTIR analyses offer a comprehensive understanding of the synthesis process, demonstrating how bioactive compounds in J. adhatoda facilitate controlled growth and stabilization. These attributes make the NPs highly promising candidates for applications in photocatalysis, antimicrobial treatments, and optoelectronics.
![]() | ||
Fig. 9 (a) Zeta potential of biogenic synthesized ZnO NPs at different pH values, (b) energy band diagram for biogenic ZnO NPs. |
The stability and dispersion capability of the ZnO NPs are further attributed to the capping molecules derived from J. adhatoda extract. These bioorganic compounds, rich in functional groups such as hydroxyl (–OH) and carboxyl (–COOH), form a stabilizing layer on the NP's surface, preventing agglomeration. The observed negative surface charge at higher pH values reinforces the role of these phytochemicals in maintaining the stability of the ZnO NPs, making them highly suitable for applications in aqueous environments. The zeta potential analysis confirms that the biogenically synthesized ZnO NPs are stable and well-dispersed across a wide pH range. This stability, combined with their tuneable surface charge properties, positions them as effective candidates for applications requiring high colloidal stability, such as photocatalysis and environmental remediation.58
The calculated band edge positions align with the requirements for efficient photocatalysis. The positive EVB value indicates strong oxidative potential, enabling the generation of hydroxyl radicals (OH) through water oxidation. Simultaneously, the negative ECB value ensures sufficient reduction potential for the formation of superoxide radicals (O2˙−) via electron transfer to dissolved oxygen. This dual capability enhances the photocatalytic degradation of pollutants such as MG dye. The band edge determination also provides insights into the compatibility of ZnO NPs with visible and UV light absorption. The precise positioning of the VB and CB edges confirms the suitability of these biogenic ZnO NPs for environmental remediation applications. Additionally, the role of phytochemical capping agents from J. adhatoda extract, which potentially influence electronic properties by modulating the NPs surface, further underscores the utility of green synthesis routes.
![]() | ||
Fig. 10 Photocatalytic degradation of the MG dye with various dosages of biogenic NPs: (a) 50 mg L−1, and (b) removal efficiency. |
In Fig. 10(b), the % removal efficiency of MG dye is analyzed over time for two different ZnO NP dosages: 50 mg L−1 and 100 mg L−1. The plot shows a significant difference in degradation performance based on catalyst concentration. At a dosage of 50 mg L−1, ZnO NPs achieved 96.5% removal of MG dye within 180 minutes, demonstrating substantial photocatalytic activity even at lower concentrations. However, increasing the catalyst dosage to 100 mg L−1 significantly accelerated the degradation process, achieving 99.8% removal within the same time frame. This enhanced performance is attributed to the increased availability of active sites and a higher generation of reactive oxygen species (ROS), such as hydroxyl radicals (˙OH) and superoxide radicals (O2˙−), at the higher concentration. The greater density of ZnO NPs in the solution improves dye adsorption, photon absorption, and sustained ROS generation, all of which contribute to faster and more efficient MG degradation.
The initial stages of the reaction (0–40 minutes) at 100 mg L−1 show a steep increase in removal efficiency, indicating rapid dye degradation due to the abundant active sites and higher ROS availability. In comparison, the degradation rate at 50 mg L−1 is slower, as evidenced by the gentler curve during the same period, reflecting the limited catalyst concentration. As the reaction progresses, the difference in degradation rates becomes less pronounced, as most dye molecules have been degraded, and the remaining intermediates are effectively mineralized by the excess ROS at both dosages.
Table 5 presents a comprehensive comparison of ZnO NPs synthesized via various methods, including hydrothermal and sol–gel techniques, as reported in prior studies, with the green synthesis approach adopted in this work using J. adhatoda leaf extract. The current study achieved significantly smaller particle sizes (5.2 nm crystallite size, 12.41 nm particle size) compared to previously reported values (∼30–40 nm), with a quasi-spherical to polygonal morphology. The band gap energy of 3.09 eV demonstrates enhanced optical properties suitable for advanced photocatalytic applications. In contrast to the photocatalytic degradation of methylene blue and antibacterial activity emphasized in earlier studies, the ZnO NPs synthesized in this work exhibited dual functionality, effectively degrading malachite green and reducing 4-nitrophenol. Furthermore, the eco-friendly synthesis method utilizing phytochemicals highlights the unique advantages of this study, such as improved crystallinity, superior particle size control, and sustainable NPs production.
Order | Aspect | Ref. 62 | Ref. 63 | Present work |
---|---|---|---|---|
1 | Synthesis method | Hydrothermal method | Sol–gel method | Green synthesis using J. adhatoda leaf extract |
2 | Particle size | ∼40 nm | ∼30 nm | 5.2 nm (crystallite size), 12.41 nm (particle size) |
3 | Morphology | Spherical | Rod-like | Quasi-spherical to polygonal |
4 | Band gap energy (eV) | 3.1 | 3.0 | 3.09 |
5 | Applications | Photocatalytic degradation of methylene blue | Antibacterial activity | Photocatalytic degradation of malachite green and 4-nitrophenol reduction |
6 | Unique advantage | Moderate crystallinity and size uniformity | Enhanced size control with post-synthesis treatment | Achieved smallest reported size, dual pollutant degradation, eco-friendly synthesis |
In the absence of any quenching agent (control), the ZnO NPs achieved a degradation efficiency of 85%, indicating their high photocatalytic potential. Upon the addition of tertiary butanol, a known hydroxyl radical (OH) scavenger, the degradation efficiency decreased significantly to 45%, highlighting the critical role of hydroxyl radicals in breaking down MG dye molecules. Similarly, the addition of benzoquinone, a scavenger for superoxide radicals (˙O2−), resulted in a reduction of degradation efficiency to 60%, confirming the contribution of superoxide radicals to the photocatalytic process. Sodium azide, a singlet oxygen (1O2) quencher, reduced the degradation efficiency to 70%, indicating that singlet oxygen plays a secondary but supportive role in the mechanism.
The results indicate that hydroxyl radicals are the primary ROS responsible for MG degradation, followed by superoxide radicals and singlet oxygen. The high degradation efficiency in the absence of quenchers (control) highlights the synergistic action of these reactive species, driven by the ZnO NPs' photocatalytic activity. This quenching study confirms that biogenic ZnO NPs generate a diverse range of ROS, with hydroxyl radicals playing a dominant role in breaking down the aromatic structure of MG dye. The interaction of photogenerated electrons and holes with dissolved oxygen and water molecules under UV light enables continuous ROS production, making this system highly effective for photocatalytic applications.
Under UV light irradiation, ZnO NPs absorb energy, exciting electrons (e−) from the valence band (VB) to the conduction band (CB), leaving behind holes (h+) in the VB:64
ZnO + UV light → e− + h+ | (4) |
The phytochemical capping agents act as stabilizing agents that mitigate electron–hole recombination, sustaining the availability of reactive charge carriers.65 These photogenerated charge carriers initiate the generation of reactive oxygen species (ROS), as follows:
The photogenerated electrons in the conduction band reduce dissolved oxygen (O2) into superoxide radicals (˙O2−):
e− + O2 → ˙O2− | (5) |
The superoxide radicals further react with water (H2O) or hydroxide ions (OH−), producing hydroxyl radicals (˙OH), which are highly reactive and capable of degrading complex organic molecules:
˙O2− + H2O → ˙OH + OH− | (6) |
These ROS, including superoxide and hydroxyl radicals, target the aromatic structure of MG dye, inducing oxidative cleavage and subsequent degradation into smaller, less toxic intermediates. The photogenerated holes (h+) in the valence band of ZnO actively participate in the degradation process.66 They directly oxidize MG dye molecules, particularly those adsorbed on the ZnO surface:
h+ + MG → MG+ → degradation products. | (7) |
This direct hole oxidation complements the ROS-mediated degradation pathway, ensuring a dual-mode mechanism for breaking down the complex structure of MG dye. The phytochemical constituents of J. adhatoda, including flavonoids and alkaloids, impart surface functionalization to ZnO NPs. These functional groups create active sites that enhance the adsorption of MG dye molecules onto the catalyst surface.67 The close interaction between the dye and the NPs surface facilitates efficient electron transfer, accelerating the degradation reactions:
MG− + phytochemical-ZnO → MG-CTAB-Co3O4 complex. | (8) |
This increased adsorption further promotes ROS activity and enables the rapid degradation of MG dye into non-hazardous products. The J. adhatoda-mediated synthesis of ZnO NPs not only ensures eco-friendly fabrication but also enhances their photocatalytic properties. The bioactive compounds stabilize the NPs, mitigate aggregation, and optimize charge transfer dynamics.67 These capping agents effectively integrate adsorption, stabilization, and charge separation mechanisms into a unified catalytic system.68 The biogenic ZnO NPs exhibit a multifaceted mechanism for MG degradation, involving synergistic pathways of ROS generation, direct oxidation, and enhanced adsorption. This holistic mechanism ensures efficient and rapid photocatalytic degradation of MG dye, highlighting the potential of green-synthesized ZnO NPs for sustainable environmental remediation.
Sr. no. | % Removal | Photo-catalyst dose (mg L−1) | Pseudo-zero-order | Pseudo-first-order | Pseudo-second order | |||
---|---|---|---|---|---|---|---|---|
k0 (mg L−1 min−1) | R2 | k1 (min−1) | R2 | k2 (L mg−1 min−1) | R2 | |||
1 | 96.5% | 50 | 0.006 | 0.81 | 0.021 | 1 | 0.165 | 0.80 |
2 | 99.8% | 100 | 0.005 | 0.64 | 0.035 | 1 | 1.826 | 0.64 |
Analysis of the regression coefficients (R2) indicates that the pseudo-first-order model best fits the degradation data for MG (Fig. 13b), suggesting that the rate depends on the concentration of remaining MG. Rate constants k1 were 0.021 min−1 for 50 mg L−1 and 0.035 min−1 for 100 mg L−1, with high R2 values. The increase in the rate constant with higher catalyst dosage demonstrates a positive correlation between catalyst concentration and degradation rate.
![]() | ||
Fig. 13 Kinetics study of degradation of MG dye: (a) pseudo-zero-order, (b) pseudo-first-order, (c) pseudo-second-order, and (d) reusability J. adhatoda/ZnO NPs for the MG dye. |
The ability of J. adhatoda-mediated ZnO NPs to be reused effectively is crucial for sustainable applications. Fig. 13(d) demonstrates that the photocatalyst maintained over 90% of its dye removal efficiency after five reuse cycles, highlighting its stability and potential for long-term applications in wastewater treatment.
![]() | ||
Fig. 14 Intermediate produced during the photodegradation of MG dye over biogenic synthesized ZnO NPs. |
The degradation of MG dye proceeds via two distinct yet converging pathways – pathway (a): N-demethylation and pathway (b): radical-mediated oxidative cleavage – both of which culminate in the mineralization of MG dye into non-hazardous molecules. The first pathway involves the sequential N-demethylation of MG dye, producing intermediates such as mono-demethylated (m/z = 315), di-demethylated (m/z = 301), and further oxidized species (m/z = 273). These intermediates undergo additional deamination and oxidative cleavage reactions, ultimately resulting in the formation of smaller fragments such as m/z = 195. This pathway highlights the stepwise dismantling of MG's chromophore structure, culminating in the complete mineralization of the dye into non-toxic end-products. In the second pathway, radical-mediated oxidative cleavage plays a dominant role, where reactive oxygen species (ROS), such as hydroxyl radicals (˙OH) and superoxide radicals (O2˙−), target the MG chromophore directly. This process generates intermediates such as Michler's ketone (m/z = 329) and 4-(dimethylamino) phenol (m/z = 137). Michler's ketone undergoes subsequent N-demethylation, producing fragments like m/z = 226, which are further degraded into mineralized products.
The degradation process is primarily driven by the photocatalytic activation of ZnO NPs under UV light, which generates charge carriers (electron–hole pairs). The holes (h+) in the valence band oxidize water molecules, generating hydroxyl radicals that attack the MG dye and its intermediates. Simultaneously, the electrons (e−) in the conduction band reduce dissolved oxygen, forming superoxide radicals that facilitate oxidative cleavage of the chromophore. These synergistic redox processes ensure the systematic breakdown of MG dye into environmentally benign molecules, as evidenced by the absence of persistent peaks in the HRMS spectrum after complete degradation.
The rapid disappearance of the m/z = 329 peak, along with the identification of multiple intermediates, demonstrates the high efficiency and versatility of ZnO NPs in photocatalytic degradation. The dual pathways—N-demethylation and oxidative cleavage—not only emphasize the robustness of the photocatalytic mechanism but also underscore the ability of ZnO NPs to simultaneously target different functional groups within the dye molecule. The formation of mineralized end-products confirms the environmental safety of this approach, making biogenic ZnO NPs a promising solution for the remediation of dye-laden wastewater.
In this context, we investigated the photocatalytic activity of ZnO NPs synthesized using J. adhatoda extract for the reduction of 4-nitrophenol (4-NP) in the presence of sodium borohydride (NaBH4) under direct sunlight irradiation. The biogenic synthesis of ZnO NPs provides a sustainable and eco-friendly approach, ensuring high surface area and active sites for photocatalytic activity. Additionally, the nanoscale size of the particles enhances their interaction with 4-NP, leading to improved photocatalytic performance. The photocatalytic reduction mechanism was monitored using UV-visible spectroscopy by observing the characteristic absorbance peak of 4-NP. Initially, the spectrum of 4-nitrophenol displayed a prominent absorption peak at 317 nm. Upon adding freshly prepared NaBH4 solution, a significant red shift to approximately 400 nm was observed, corresponding to the formation of p-nitrophenolate ions (Fig. 15(a)).71 Fig. 15 (b) shows the UV-visible spectra after 120 minutes of reaction without a catalyst, illustrating that NaBH4 alone cannot reduce 4-NP, as indicated by the persistence of nitrophenolate ions. However, when ZnO NPs were introduced and exposed to sunlight, a rapid decrease in absorbance at 400 nm occurred, along with an increase in absorbance at 300 nm (Fig. 15(c)). The complete disappearance of the 400 nm peak within just 15 minutes confirmed the efficient photocatalytic conversion of 4-nitrophenol to 4-aminophenol.72
The relationship between absorbance and concentration allowed us to express the initial concentration as absorbance A0 (at t = 0) and the concentration at time t as absorbance At. The reaction rate constant (k) was determined from the plot of ln(Ct/C0) against time, yielding a value of 0.245 min−1 (Fig. 15(d)), indicating the high efficiency of ZnO NPs in facilitating the photocatalytic reduction process. The novelty of using biogenic ZnO NPs lies in their unique properties, including their high surface area, crystallinity, and enhanced surface charge dynamics, which promote efficient adsorption of 4-NP molecules. This leads to increased reactive site availability, thereby accelerating the photocatalytic degradation of 4-NP. Additionally, the phytochemicals from J. adhatoda used during synthesis play a critical role in stabilizing the ZnO NPs and maintaining their dispersion, ensuring sustained catalytic performance.
The overall photocatalytic reduction process can be represented by the following chemical equation
In this reaction, sodium borohydride (NaBH4) acts as a reducing agent, facilitating the conversion of 4-nitrophenol (4-NP) into 4-aminophenol (4-AP), a less toxic and more biodegradable compound. The ZnO NPs play a dual role in this process by adsorbing the reactant molecules and generating reactive species under sunlight. ZnO is well-known for its ability to generate electron–hole pairs upon UV irradiation, which are critical for initiating and sustaining photocatalytic reactions. These electron–hole pairs drive the reduction of 4-NP into 4-AP through electron transfer reactions. The tailored bandgap of ZnO NPs synthesized via green methods enables efficient light absorption in the UV region, thus enhancing their photocatalytic efficiency under natural sunlight. Furthermore, their structural configuration promotes effective charge separation and minimizes electron–hole recombination, critical for achieving high photocatalytic performance. In summary, the unique properties of biogenic ZnO NPs make them a promising material for the photocatalytic degradation of persistent pollutants like 4-nitrophenol, offering an innovative and sustainable solution for environmental remediation.
This journal is © The Royal Society of Chemistry 2025 |