Next Article in Journal
A Novel Unsupervised You Only Listen Once (YOLO) Machine Learning Platform for Automatic Detection and Characterization of Prominent Bowel Sounds Towards Precision Medicine
Previous Article in Journal
A Novel Approach for Optimizing Molecularly Imprinted Polymer Composition in Electrochemical Detection of Collagen Peptides
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Phytogenic Silver Nanoparticles Derived from Ricinus communis and Aloe barbadensis: Synthesis, Characterization, and Evaluation of Biomedical Potential

1
Centre for Pharmaceutical Innovation, University of South Australia, Adelaide, SA 5000, Australia
2
College of Animal Science & Veterinary Medicine, Shanxi Agricultural University, Taigu 030801, China
3
Institute of Microbiology, Chinese Academy of Sciences, Shijingshan, Beijing 100043, China
*
Author to whom correspondence should be addressed.
Bioengineering 2025, 12(11), 1273; https://doi.org/10.3390/bioengineering12111273
Submission received: 16 October 2025 / Revised: 3 November 2025 / Accepted: 7 November 2025 / Published: 19 November 2025
(This article belongs to the Special Issue Bioengineering Platforms for Drug Delivery)

Abstract

The green synthesis of silver nanoparticles (SNPs) using medicinal plants provides a sustainable and eco-friendly approach to nanoparticle production with promising biomedical potential. In this study, Ricinus communis and Aloe barbadensis aqueous leaf extracts were employed as reducing and stabilizing agents to synthesize R. communis SNPs (RcSNPs) and A. barbadensis SNPs (AbSNPs). The nanoparticles were characterized using ultraviolet–visible spectroscopy, dynamic light scattering, Fourier-transform infrared spectroscopy, scanning electron microscopy, transmission electron microscopy, thermogravimetric analysis, and differential scanning calorimetry to evaluate their physicochemical and thermal properties. RcSNPs and AbSNPs were predominantly spherical, with average sizes of 15–20 nm and 23–28 nm, respectively, and exhibited stability up to ~90 °C. Biological evaluations demonstrated potent antimicrobial, antioxidant, anti-inflammatory, anti-tyrosinase, and cytotoxic activities. Notably, RcSNPs and AbSNPs induced apoptosis through mitochondrial pathway modulation and showed superior cytotoxicity compared to crude plant extracts and several previously reported SNPs. These findings indicate that phytochemical-mediated SNPs not only provide a green route of synthesis but also exhibit multifunctional bioactivities, which may support their potential applications as antimicrobial, antioxidant, depigmenting, and anticancer agents in biomedical and pharmaceutical fields.

Graphical Abstract

1. Introduction

Nanoscience has advanced rapidly, leading to the development of various nanoproducts that offer significant benefits to society [1]. Nanotechnology plays a crucial role in diverse fields such as biotherapy, pharmaceuticals, drug delivery, electronics, biotechnology, cosmetics, topical formulations (ointments and creams), nano-fertilizers, and water treatment [2]. Nanoparticles, defined as particulate matter with a diameter of less than 100 nm, exhibit unique properties that make them valuable in these applications [3,4]. Among these, silver nanoparticles (SNPs) have garnered widespread attention due to their exceptional electrical and thermal conductivity, as well as surface-enhanced Raman scattering [5,6,7]. Additionally, SNPs possess various biological activities, including antibacterial, antifungal, antioxidant, anti-inflammatory, and anticancer effects, and have potential use in gene therapy as non-viral carriers [8,9].
Several methods have been developed for the synthesis of SNPs, including chemical and electrochemical techniques. However, these approaches often encounter challenges during the purification stage due to the use of hazardous chemicals or the formation of toxic by-products, necessitating high energy input [10]. Moreover, controlling the size, shape, and achieving monodispersity of nanoparticles remain common challenges [10]. These limitations highlight the need for environmentally sustainable and safe synthesis methods. Consequently, green chemistry-based approaches have gained prominence in recent years [11,12].
Green synthesis involves the biological reduction of metal ions into nanoparticles using microorganisms or plant-derived extracts obtained from leaves, fruits, and seeds [13,14]. Plant extracts are rich in secondary metabolites such as alkaloids, terpenoids, flavonoids, enzymes, cyclic peptides, amino acids, proteins, polysaccharides, tannins, retinoic acid, ascorbic acid, polyphenols, and other compounds, which act as reducing agents, capping agents, and stabilizers during nanoparticle synthesis [15,16,17]. These compounds not only facilitate the synthesis process but also enhance the biological activity of the resulting nanoparticles. Furthermore, green synthesis offers advantages such as well-defined size and morphology, ease of scale-up, and high-yield production without surface contamination [12,18,19]. Recent reviews have highlighted that plant-mediated silver nanoparticles exhibit enhanced antimicrobial and cytotoxic activities through green synthesis routes [20,21,22].
Numerous plants and plant parts have been utilized for the biosynthesis of SNPs [14], including Diospyros lotus [11], Zingiber officinale [23], Vernonia amygdalina [24], Sida rhombifolia [25], grape seeds [26], Physalis angulata [27], Camellia sinensis [28], Soymida febrifuga [29], algae (Parachlorella kessleri) [30], Mikania cordata [31], chamomile (Matricaria recutita) [32], and Ilex paraguariensis [33]. The selection of plant sources often favors those that are readily available and renewable [34]. Factors such as extract concentration, metal salt concentration, temperature, pH, and reaction time significantly influence the rate of nanoparticle formation, yield, and physicochemical properties [12,35,36,37].
Ricinus communis L. (castor plant) and Aloe barbadensis Mill. (aloe vera) were selected for their rich phytochemical profiles and established therapeutic potential [38,39]. R. communis contains flavonoids, alkaloids, and phenolic acids known for antimicrobial and antioxidant properties [40], while A. barbadensis is rich in polysaccharides and phenolic compounds that facilitate metal ion reduction and stabilization [41]. These characteristics make both plants effective and sustainable candidates for green synthesis of metallic nanoparticles [42].
In this study, we aimed to synthesize optimized silver nanoparticles using the leaves of Ricinus communis L. and Aloe barbadensis Mill., perform detailed optimization of synthesis parameters, and evaluate their therapeutic potential.

2. Materials and Methods

2.1. Materials

Silver nitrate, bovine serum albumin (BSA), and other chemicals were purchased from Sigma-Aldrich (St. Louis, MO, USA). B16F10 murine melanoma and HepG2 human hepatocarcinoma cell lines were obtained from the Cell Bank of the Chinese Academy of Sciences (Shanghai, China). RPMI-1640, Dulbecco’s Modified Eagle’s Medium (DMEM), fetal bovine serum (FBS), and trypsin were supplied by Thermo Fisher Scientific (Waltham, MA, USA). All other reagents were of analytical grade, and double-distilled water was used throughout.

2.2. Methods

2.2.1. Preparation of Plant Extracts

Fresh leaves of Ricinus communis L. and Aloe barbadensis Mill. were collected from Shanxi Agricultural University, washed with running and double-distilled water, and air-dried for 10–12 days at room temperature. Approximately 20 g of powdered R. communis leaves and sliced A. barbadensis leaves were separately boiled in 200 mL distilled water for 30 min. The extracts were cooled, filtered through Whatman No. 1 paper, and stored at 4 °C until further use. For methanolic extracts, 10 g of dried leaf powder was soaked in 100 mL methanol for 48 h, filtered, and concentrated under reduced pressure.
Distilled water was used to maintain a fully green synthesis protocol, as hydrophilic phytochemicals (phenolics, flavonoids, glycosides) are sufficient to reduce Ag+ ions and stabilize nanoparticles. This approach aligns with green chemistry principles and has been validated in multiple plant-based SNP syntheses [43].

2.2.2. Phytochemical Analysis

Qualitative screening followed established phytochemical tests as described by Trease and Evans [44], Harborne [45], and Sofowora [46], including Dragendorff’s for alkaloids, ferric chloride for phenolics, foam test for saponins, and Liebermann–Burchard for steroids [47]. Qualitative screening of aqueous leaf extracts was performed to detect alkaloids, phenols, proteins, carbohydrates, glycosides, steroids, saponins, tannins, flavonoids, and terpenoids using standard procedures [48,49,50].

2.2.3. Total Phenolic Content (TPC) Estimation

TPC was determined using a modified Folin–Ciocalteu method [51]. Briefly, 100 µg/mL of plant extract or gallic acid standard was mixed with 0.5 mL Folin–Ciocalteu reagent, incubated in the dark for 10 min, followed by addition of 2.5 mL 20% sodium carbonate. After 1 h in the dark, absorbance was measured at 765 nm using a UV–Vis spectrophotometer (UV-2600i, Shimadzu, Kyoto, Japan).

2.2.4. Biogenic Synthesis of Silver Nanoparticles (SNPs)

A 1 mM AgNO3 solution (100 mL) was prepared. For R. communis silver nanoparticle (RcSNP) or A. barbadensis silver nanoparticle (AbSNP) synthesis, 1 mL of R. communis or A. barbadensis leaf extract was added dropwise to 9 mL AgNO3 in separate flasks. The mixtures were shaken, and incubated in the dark at room temperature, and a control without extract was maintained. SNP solutions were purified by repeated centrifugation (10,000 rpm, 15 min) and redispersed in deionized water three times, and then freeze-dried for further characterization.

2.2.5. Optimization of SNP Synthesis

AgNO3 concentrations (0.05–4 mM) and extract volumes (1–4 mL) were varied to optimize SNP synthesis. With optimized concentrations, reaction time (30 min–24 h), temperature (25–110 °C), and pH (2–13) were systematically evaluated. UV–Vis spectra were recorded at 300–800 nm, and all glassware was covered to prevent photo-degradation [52,53].

2.2.6. UV–Visible Spectroscopy

Absorption spectra of SNPs were recorded using a Shimadzu UV-2600i spectrophotometer (UV-2600i, Shimadzu, Kyoto, Japan) at 300–800 nm with 1 mL aliquots of diluted reaction mixtures at various time intervals.

2.2.7. Stability Study

Purified SNPs were stored in the dark at ambient conditions. Stability was monitored via UV–Vis spectroscopy (300–800 nm) over a period of up to 30 days.

2.2.8. Characterization of SNPs

DLS and Zeta Potential: Hydrodynamic size and surface charge were measured using a ZetaPALS dynamic light scattering instrument (Brookhaven Instruments Corporation, Holtsville, NY, USA). SNPs (5 mg) were dispersed in 5 mL water, sonicated, and measured at 689 nm, 90 °C, with three readings per sample.
FTIR Spectroscopy: FTIR spectra of leaf extracts and SNPs were obtained using KBr pellets on a Fourier Transform Infrared Spectrophotometer (IRAffinity-1, Shimadzu, Kyoto, Japan) in the 400–4000 cm−1 range with 4 cm−1 resolution.
SEM Analysis: Morphology was visualized using a Scanning Electron Microscope (JSM-5600, JEOL Ltd., Tokyo, Japan) on thin films of SNPs prepared on carbon-coated copper grids.
TEM Analysis: Particle size and morphology were confirmed using a Transmission Electron Microscope (JEM-100CX II, JEOL Ltd., Tokyo, Japan) with SNP suspensions (1 µg/mL) placed on copper grids dried at 37 °C.
TGA and DSC: Thermal stability was assessed using a Thermogravimetric/Differential Scanning Calorimetry Analyzer (STA 7300, Hitachi High-Tech, Tokyo, Japan). Lyophilized SNPs (8.6 mg) were heated from 40 °C to 900 °C at 10 °C/min under N2 flow (40 mL/min), with reference crucibles used for comparison.

2.2.9. Antimicrobial Activity

The antimicrobial potential of RcSNPs and AbSNPs was assessed using the agar well diffusion method, which allows direct application of nanoparticle suspensions [54]. Bacterial strains (Staphylococcus aureus ATCC 25923, Escherichia coli ATCC 25922, Salmonella typhi ATCC 6539) and fungal strains (Candida albicans ATCC 10231, Aspergillus niger ATCC 16404) were obtained from the Microbial Culture Collection, Shanxi Agricultural University (Taigu, China). Gentamicin sulfate (≥98%) and fluconazole (≥99%) were purchased from Sigma-Aldrich (St. Louis, MO, USA) and served as positive controls.
Agar Well Diffusion: Cultures were standardized to 0.5 McFarland (~1.5 × 108 CFU/mL). Mueller–Hinton agar (for bacteria) or potato dextrose agar (for fungi) plates were obtained from Oxoid Ltd. (Basingstoke, UK).SNP solutions (50 μg/mL, 50–100 μL per well) were added to wells andplates were incubated at 37 °C for 24 h, and zones of inhibition were measured [55].

2.2.10. In Vitro Anti-Inflammatory Assays

Protein Denaturation: 500 μL of bovine serum albumin (BSA, ≥96%, Sigma-Aldrich, St. Louis, MO, USA) (1%) was mixed with 100 μL extract or SNPs (pH 6.3), incubated at 37 °C for 20 min, then heated at 55 °C for 20 min. Absorbance at 660 nm was measured using a UV–Vis spectrophotometer (UV-2600i, Shimadzu, Kyoto, Japan). Acetylsalicylic acid (aspirin, ≥99%, Sigma-Aldrich, St. Louis, MO, USA) was used as a positive control [56]. The following equation estimated the inhibition (%) of protein denaturation:
%   I n h i b i t i o n   =   ( A _ c o n t r o l     A _ s a m p l e )   /   A _ c o n t r o l   ×   100
where:
  • Acontrol = absorbance of the control sample
  • Asample = absorbance of the test sample
Protease Inhibition: 100 μL extract/SNPs was mixed with 100 μL BSA and incubated for 5 min; then, 250 μL of trypsin (from porcine pancreas, ≥10,000 BAEE units/mg, Sigma-Aldrich, St. Louis, MO, USA) was added. Supernatant absorbance at 210 nm was taken to measure protease inhibition [57].

2.2.11. In Vitro Antioxidant Assays

DPPH Assay: 250 μL of 0.3 mM 2,2-diphenyl-1-picrylhydrazyl (DPPH, Sigma-Aldrich, St. Louis, MO, USA) solution was mixed with 2.5 mL extract/SNPs (10–100 μg/mL). Butylated hydroxytoluene (BHT, ≥99%, Sigma-Aldrich, St. Louis, MO, USA) served as a positive control, and absorbance was recorded at 517 nm [58].
H2O2 Scavenging: Extracts/SNPs (10–100 μg/mL) were mixed with 50 μL of 5 mM hydrogen peroxide (H2O2, Merck, Darmstadt, Germany), incubated for 20 min, and the absorbance was recorded at 610 nm [59].
Nitric Oxide Scavenging: Sodium nitroprusside (≥98%, Sigma-Aldrich, St. Louis, MO, USA)was used to generate nitric oxoid (NO), and incubated with extract/SNPs (10–100 μg/mL) for 60 min; nitrite was measured using Griess reagent [60].
Reducing Power: Extracts/SNPs (10–100 μg/mL) were mixed with phosphate buffer (0.2 M, pH 6.6) and 1% potassium ferricyanide (K3[Fe(CN)6], Merck, Darmstadt, Germany) incubated at 50 °C for 20 min, cooled, and mixed with 10% trichloroacetic acid (TCA, Sigma-Aldrich, St. Louis, MO, USA). The mixture was centrifuged at 3000 rpm for 10 min; the supernatant was combined with 0.1% ferric chloride (FeCl3, Merck, Darmstadt, Germany), and absorbance was measured at 700 nm [61].

2.2.12. Tyrosinase Inhibition

SNPs were tested with L-tyrosine (≥98%, Sigma-Aldrich, St. Louis, MO, USA) and mushroom tyrosinase (≥1000 units/mg solid, Sigma-Aldrich, St. Louis, MO, USA) in 50 mM potassium phosphate buffer (pH 6.8, Thermo Fisher Scientific, Waltham, MA, USA) at 30 °C. The reaction mixture was pre-incubated for 10 min, and absorbance at 492 nm was recorded using a microplate reader (Multiskan Sky, Thermo Fisher Scientific, Waltham, MA, USA). Percentage inhibition was calculated accordingly [62].

2.2.13. In Vitro Cytotoxicity

Cell Culture: B16F10 (ATCC CRL-6475) and HepG2 (ATCC HB-8065) cell lines were obtained from the Cell Bank of the Chinese Academy of Sciences (Shanghai, China) and used at passages 10–15. B16F10 and HepG2 cells were maintained at 37 °C, 5% CO2, detached with 0.25% trypsin, and seeded at 3 × 103 cells/well in 96-well plates.
Cytotoxicity Evaluation: Cells treated with SNPs (0–100 μg/mL) for 24 h. MTT assay was performed using 3-(4,5-dimethylthiazol-2-yl)-2,5-diphenyltetrazolium bromide (MTT, Sigma-Aldrich, St. Louis, MO, USA), and formazan crystals were dissolved in dimethyl sulfoxide (DMSO, Sigma-Aldrich, St. Louis, MO, USA). Absorbance was recorded at 570 nm using a microplate reader (Multiskan Sky, Thermo Fisher Scientific, Waltham, MA, USA) [63].
Cell viability (%) = (OD sample/OD control) × 100
where
  • ODsample = absorbance of treated cells
  • ODcontrol = absorbance of untreated cells used as control.

2.2.14. Statistical Analysis

All experiments were performed in triplicate. Data are presented as mean ± SD and were analyzed using one-way ANOVA with GraphPad Prism v5 (GraphPad Software Inc., San Diego, CA, USA).

3. Results

3.1. Phytochemical Analysis

Phytochemical screening of aqueous and methanol leaf extracts of Ricinus communis and Aloe barbadensis revealed the presence of various bioactive constituents. The results, summarized in Table 1, indicate that both plant extracts contain key biomolecules such as alkaloids, phenols, proteins, carbohydrates, glycosides, steroids, saponins, tannins, flavonoids, and terpenoids, which may contribute to their therapeutic potential.

3.2. Visual Inspection

The synthesis of silver nanoparticles (SNPs) can be initially indicated by a distinct color change in the reaction mixture, which occurs due to the excitation of surface plasmon resonance (SPR) of SNPs [14,64]. Typically, the solution transitions from colorless to yellow, pale yellow, yellowish-brown, and eventually to dark brown. In the present study, a progressive change in color from yellowish to light brown, and finally to dark brown, was observed upon addition of the plant extract to silver nitrate solution, confirming the formation of SNPs (Figure 1). The synthesis method used has been previously shown to be reproducible; in our experiments, the nanoparticle properties were consistent with prior reports [65,66]. This consistency confirms the reliability and stability of the adopted phytogenic synthesis approach.

3.3. UV-Visible Spectroscopy

UV–visible (UV–vis) spectroscopy is a simple and widely used technique for monitoring the synthesis of silver nanoparticles (SNPs) [67]. In this study, reactions were carried out at room temperature. The characteristic surface plasmon resonance (SPR) bands observed near 435 nm for R. communis and 442 nm for A. barbadensis confirmed the reduction of Ag+ ions to metallic silver. Although water absorption bands are predominant in KBr FTIR spectra, characteristic peaks corresponding to hydroxyl (~3400 cm−1), carbonyl (~1635 cm−1), and amine (~1385 cm−1) groups were consistent with functional groups reported for plant extracts and confirm biomolecule involvement in Ag+ reduction and nanoparticle stabilization [68].
To optimize the synthesis, key parameters including AgNO3 concentration, extract concentration, reaction time, temperature, and pH were systematically varied, as these factors strongly influence nanoparticle formation [36].

3.3.1. Effect of AgNO3 Concentration

SNPs were synthesized using varying concentrations of AgNO3 (0.5–4 mM), while maintaining the extract concentration at 1 mL under constant reaction conditions. No SPR band was detected at low AgNO3 concentrations; however, distinct peaks were observed at 435 nm for 1 mM AgNO3 with R. communis extract, and at 442 nm for 2 mM AgNO3 with A. barbadensis extract. Increasing AgNO3 concentration led to higher peak intensity and peak broadening, indicating an increase in SNP concentration and particle aggregation (Figure 2a,f) [40]. Based on these results, the optimal AgNO3 concentrations were set at 1 mM for R. communis and 2 mM for A. barbadensis for subsequent experiments.

3.3.2. Effect of Leaf Extract Concentration

Extract concentration also influenced nanoparticle properties. Optimization was performed by fixing the AgNO3 concentration and varying the extract volume (1–4 mL). With increasing extract concentration, the SPR bands became narrower, suggesting improved monodispersity of SNPs (Figure 2b,g). An extract volume of 1 mL was determined to be optimal for further experiments.

3.3.3. Effect of Reaction Time

Reaction time significantly affected SNP formation and stability. At the initial stage, no SPR peak appeared immediately after mixing silver nitrate with plant extracts. After 1 h, characteristic SPR bands at 435 nm (R. communis) and 442 nm (A. barbadensis) emerged, with peak intensity increasing over time. Maximum absorbance was recorded at 6 h for R. communis and 9 h for A. barbadensis (Figure 2c,h). Beyond these times, peak broadening occurred, likely due to agglomeration or increased particle size. Therefore, 6 h and 9 h were selected as the optimized reaction times for R. communis and A. barbadensis, respectively.

3.3.4. Effect of pH

pH had a notable influence on the morphology and stability of SNPs. Higher pH values enhanced the reduction of Ag+ ions, resulting in stronger absorbance and improved stability of the nanoparticles (Figure 2d,i).

3.3.5. Effect of Temperature

Temperature variations also affected SNP synthesis. Although different temperatures were tested, room temperature was found to be optimal, producing small, spherical SNPs with a sharp and distinct single SPR band at shorter wavelengths (Figure 2e,j).

3.4. Stability of Synthesized SNPs

The stability of biosynthesized SNPs was evaluated by monitoring their UV–vis spectra over time. As shown in Figure 3a,b, prolonged storage led to a gradual decrease in absorbance intensity and band broadening, indicating possible agglomeration or partial release of metallic ions. Importantly, the SPR band position remained unchanged even after 30 days, suggesting that the nanoparticles retained their structural integrity [69]. In addition, reproducibility was confirmed by synthesizing SNPs in triplicate over a period of six months. The spectra remained consistent, demonstrating that R. communis and A. barbadensis extract-mediated SNPs possess excellent stability and dispersity.

3.5. Dynamic Light Scattering (DLS) and Zeta Potential (ZP) Analysis

Dynamic light scattering (DLS) is a widely used technique for determining nanoparticle size distribution at the nanometer scale [70]. The DLS profiles of biogenic SNPs are presented in Figure 4a,b. Under optimized conditions, the average hydrodynamic diameter of RcSNPs and AbSNPs was 93.65 ± 1.55 nm and 126.09 ± 0.13 nm, respectively. The DLS histograms (Figure 4) show narrow and unimodal size distributions, confirming the uniformity of the RcSNP and AbSNP populations. The corresponding polydispersity index (PDI) values were 0.287 ± 0.002 and 0.130 ± 0.019, indicating narrow size distributions. The zeta potential (ZP) values were −67.09 mV (RcSNPs) and −83.42 mV (AbSNPs), confirming strong surface charge and excellent colloidal stability. These findings are consistent with the UV–vis analysis results.

3.6. Fourier Transform Infrared (FTIR) Spectroscopy

Fourier transform infrared (FTIR) spectroscopy was employed to investigate the surface chemistry of SNPs and to identify biomolecules responsible for reducing Ag+ ions and stabilizing the nanoparticles [71]. The FTIR spectra of both plant extracts and their corresponding SNPs were compared to identify characteristic peaks and spectral shifts.
For RcSNPs, prominent absorption bands were observed at 3326 cm−1 and 1720 cm−1, while AbSNPs exhibited bands at 3576 cm−1 and 1850 cm−1 (Figure 5). The bands at 3326 cm−1 (RcSNPs) and 3576 cm−1 (AbSNPs) correspond to –OH stretching vibrations, whereas the bands at 1720 cm−1 and 1850 cm−1 are attributed to N–H stretching of amine groups, indicating the capping of SNPs by these functional groups. Additionally, lower intensity bands at 2105 cm−1 (RcSNPs) and 2157 cm−1 (AbSNPs) were detected, corresponding to carbonyl (C = O) stretching. Although the FTIR spectrum is dominated by water peaks, the presence of functional groups such as –OH, C = O, and –NH was inferred from comparative analysis with literature reports [68,72].

3.7. Scanning Electron Microscopy (SEM) Analysis

Scanning electron microscopy (SEM) was used to examine the morphology and size of the biosynthesized SNPs. The micrographs revealed nanoparticles with spherical, cubic, triangular, and rectangular shapes, distributed relatively evenly. Some degree of agglomeration was observed, likely due to the limited stabilization provided by weak capping agents [73]. The agglomerated particle sizes ranged from 256 to 272 nm for RcSNPs and 263 to 278 nm for AbSNPs, while the average particle sizes were estimated to be 68 nm (RcSNPs) and 74 nm (AbSNPs) (Figure 6a,b).

3.8. Transmission Electron Microscopy (TEM) Analysis

Transmission electron microscopy (TEM) provided further insights into the morphology and size of the biosynthesized SNPs (Figure 7a–d). The images revealed that most nanoparticles were spherical, consistent with the observations from UV–vis and SEM analyses. Particle size measurements performed using ImageJ software (version 1.52a) indicated sizes of 15–20 nm for RcSNPs and 23–28 nm for AbSNPs.

3.9. Thermogravimetric Analysis (TGA) and Differential Scanning Calorimetry (DSC)

Thermogravimetric analysis (TGA) was performed to evaluate the thermal stability of the biosynthesized SNPs. The results indicated that RcSNPs began to degrade at approximately 200 °C, while AbSNPs showed initial degradation at 240 °C. A gradual weight loss was observed up to 800 °C (Figure 8a,b). Differential scanning calorimetry (DSC) revealed endothermic peaks at 190 °C for RcSNPs and 220 °C for AbSNPs, corresponding to thermal transitions of the nanoparticles (Figure 8a,b).

3.10. Antimicrobial Activity of Biogenic SNPs

The antimicrobial potential of RcSNPs and AbSNPs was evaluated against selected bacterial and fungal pathogens, and their efficacy was quantified by measuring the zones of inhibition (mm) (Figure 9 and Figure 10; Table 2).
RcSNPs exhibited zones of inhibition of 22.3 ± 0.8 mm against C. albicans and 20.2 ± 0.9 mm against S. aureus, followed by 18.4 ± 0.7 mm, 19.3 ± 0.5 mm, and 21.4 ± 0.7 mm against A. niger, S. typhi, and E. coli, respectively. For AbSNPs, the corresponding zones were 21.8 ± 0.4 mm and 20.3 ± 0.05 mm against A. niger and S. aureus, and 24.3 ± 0.5 mm, 18.6 ± 0.07 mm, and 17.3 ± 0.05 mm against C. albicans, S. typhi, and E. coli, respectively (Figure 9 and Figure 10).
Both SNPs demonstrated the highest antibacterial activity against S. aureus, followed by S. typhi and E. coli. When compared to the aqueous leaf extracts of R. communis and A. barbadensis, RcSNPs showed enhanced antibacterial activity against S. typhi, whereas AbSNPs were more effective against E. coli. Regarding antifungal activity, RcSNPs exhibited superior inhibition of C. albicans, while AbSNPs were more active against A. niger. Notably, the antifungal activity of both SNPs exceeded that of the standard antifungal agent, fluconazole, highlighting their strong potential as antimicrobial agents.

3.11. In Vitro Anti-Inflammatory Assay

The anti-inflammatory potential of R. communis and A. barbadensis extracts and their corresponding SNPs was evaluated using the heat-induced albumin denaturation method. The plant extracts exhibited 67.45 ± 1.7% (R. communis) and 66.03 ± 1.8% (A. barbadensis) inhibition, whereas RcSNPs and AbSNPs showed significantly higher activity at 90.67 ± 1.7% and 89.79 ± 1.7%, respectively. Acetylsalicylic acid was used as the standard anti-inflammatory agent (Figure 11a). A concentration-dependent increase in inhibition was observed for both extracts and SNPs.
Furthermore, the extracts and SNPs demonstrated notable antiprotease activity. The standard drug, acetylsalicylic acid, exhibited the highest protease inhibitory effect (95.37 ± 1.3%). R. communis and A. barbadensis extracts showed 65.45 ± 1.6% and 63.03 ± 1.7% inhibition, respectively, whereas RcSNPs and AbSNPs displayed enhanced activity at 89.67 ± 1.4% and 87.79 ± 1.3%, respectively (Figure 11b).

3.12. In Vitro Antioxidant Assay

Free radicals are implicated in numerous human diseases, including arthritis, atherosclerosis, reperfusion injury, ischemia, central nervous system damage, gastritis, AIDS, and cancer. Recently, natural antioxidants have gained considerable attention due to their potential health benefits [73]. The antioxidant potential of Ricinus communis (Rc) and Aloe barbadensis (Ab) extracts and their corresponding SNPs was evaluated using DPPH, hydrogen peroxide (H2O2), nitric oxide (NO) scavenging, and reducing power assays.
In the DPPH assay, RcSNPs and AbSNPs demonstrated significant, concentration-dependent radical scavenging activity, with increasing inhibition observed at higher concentrations (10–100 µg/mL) (Figure 12a). Similarly, H2O2 scavenging activity was assessed spectrophotometrically, and RcSNPs, AbSNPs, and ascorbic acid exhibited 90.57 ± 0.9%, 89.02 ± 0.7%, and 92.05 ± 0.4% inhibition at 100 µg/mL, respectively (Figure 12b). H2O2 scavenging activity was slightly lower than DPPH scavenging for both SNPs, suggesting differential efficacy against specific reactive species.
Biogenic RcSNPs and AbSNPs also exhibited NO scavenging activity in a dose-dependent manner, achieving 87.41 ± 1.8% and 83.87 ± 1.7% inhibition at 100 µg/mL, respectively, slightly lower than the standard BHT (90.83 ± 1.8%) (Figure 12c). The reducing power of the SNPs similarly increased with concentration, with maximum activities of 91.41 ± 1.6% (RcSNPs) and 89.08 ± 1.7% (AbSNPs), comparable to BHT (93.76 ± 1.7%) (Figure 12d).
Overall, these results indicate that both RcSNPs and AbSNPs possess strong, concentration-dependent antioxidant potential, comparable to standard antioxidant agents, highlighting their promising therapeutic value.

3.13. Tyrosinase Inhibition Activity

RcSNPs and AbSNPs exhibited potent tyrosinase inhibitory activity, with inhibition percentages of 97.98 ± 1.5% and 98.43 ± 1.5%, respectively (Table 3). These results indicate that both biogenic SNPs are highly effective in inhibiting tyrosinase, suggesting their potential applications in skin-whitening and anti-melanogenic formulations.

3.14. In Vitro Cytotoxicity Evaluation

Given the diverse biological applications of silver nanoparticles, the anticancer potential of RcSNPs and AbSNPs was evaluated using the MTT assay [74,75]. B16F10 and HepG2 cells were treated with varying concentrations of biogenic RcSNPs and AbSNPs (0–100 µg/mL) for 24 and 48 h.
Both RcSNPs and AbSNPs exhibited significant, dose-dependent cytotoxic effects against B16F10 and HepG2 cells (Figure 13a–d). The IC50 values at 24 h were 10 µg/mL for RcSNPs and 12 µg/mL for AbSNPs. Cytotoxicity increased further after 48 h of incubation, indicating a time-dependent enhancement of their anticancer activity. These results demonstrate the promising therapeutic potential of RcSNPs and AbSNPs as cytotoxic agents against melanoma and hepatocellular carcinoma cell lines.

4. Discussion

4.1. Synthesis and Optimization of Silver Nanoparticles (SNPs)

Ricinus communis and Aloe barbadensis plant extracts contain a variety of bioactive compounds, including alkaloids, glycosides, flavonoids, tannins, saponins, terpenoids, phenols, and reducing sugars, which can effectively reduce silver ions to silver nanoparticles (SNPs) and act as natural capping agents [38,76,77]. These phytochemicals highlight the medicinal potential of R. communis and A. barbadensis, as supported by previous studies [78,79]. In particular, the phenolic compounds exhibit redox properties that contribute to their antioxidant activity [80,81] and are largely responsible for the biological effects of the extracts. Consequently, both extracts are expected to demonstrate significant antioxidant, anti-inflammatory, and antibacterial activities. Literature indicates that phenolic and flavonoid compounds are more efficiently extracted in aqueous solutions and are biosynthesized via the phenylpropanoid pathway from phenylalanine [82]. The current findings align with total phenolic content (TPC) results reported for plant extracts from Calophyllum tomentosum, bearberry, olive leaves, Euodia malayana, Gnetum gnemon, Haloxylon scoparium, and Khaya senegalensis [7,83,84,85,86,87].
The color change of the reaction mixture serves as the initial indication of silver nanoparticle (SNP) formation. In this study, the solution color changed from yellowish to light brown and ultimately to dark brown (Figure 1), consistent with previous reports [88,89]. This color development is attributed to surface plasmon resonance (SPR), which arises from the collective oscillation of conduction electrons induced by an electromagnetic field interacting with electrons at the surface of metallic nanoparticles [90]. The SPR band’s position and width depend on the nanoparticle’s size and shape, the metal’s dielectric properties, and the surrounding environment [91]. UV–visible (UV–vis) spectroscopy, a simple and widely used technique, confirmed SNP synthesis, with SPR bands observed near 435 nm for R. communis and 442 nm for A. barbadensis, in agreement with the literature [92,93]. At higher silver nitrate concentrations, larger SNPs were formed, accompanied by broader SPR bands, likely due to the inhibitory effect of excess silver ions on nanoparticle formation, which slowed the reaction rate.
Another study reported that increasing the silver nitrate concentration enhanced silver nanoparticle (AgNP) synthesis, as evidenced by the intensified color of the reaction mixture due to the aggregation of silver ions, leading to greater nanoparticle formation [94]. Using 1 mM silver nitrate with Solanum trilobatum extract produced monodisperse AgNPs without aggregation. However, at 2–4 mM silver nitrate, a spectral band appeared at 460 nm, and at 5 mM, the SPR band shifted to 500 nm, indicating polydispersity and larger particle sizes. Increasing the silver nitrate concentration generally broadens the plasmon resonance band [95]. Similarly, increasing the leaf extract concentration to 4 mL (Figure 2b,g) resulted in a red shift of the SPR band and subtle changes in absorbance, reflecting changes in particle size [64]. Higher extract concentrations also increased absorption intensity. UV–vis spectroscopy is thus widely accepted for monitoring nanoparticles of controlled size and shape in aqueous suspensions. These observations are consistent with previous reports, where AgNP synthesis using Pithophora dogonia and Cochlospermum religiosum extracts exhibited SPR bands around 445 nm [96,97]. Collectively, these findings suggest that R. communis and A. barbadensis leaf extracts enable rapid and efficient synthesis of AgNPs, as confirmed by their UV–vis absorption spectra.
Similar observations were reported for silver nanoparticles (AgNPs) synthesized using Malus domestica [98], Parthenium hysterophorus [99], Abies alba and Pinus sylvestris [100]. Increasing the concentration of plant extract enhanced the absorption peak, indicating a greater reduction of silver ions and the formation of stable, well-defined AgNPs. At higher extract concentrations, phytochemicals not only act as reducing agents but also coat the nanoparticles, preventing agglomeration and improving stability [101].
The synthesis of silver nanoparticles (SNPs) was found to increase with reaction time; however, prolonged reaction periods can lead to instability due to agglomeration and increased particle size. Logeswari et al. reported SNP formation at 24–48 h at 37 °C using 1 mM silver nitrate and ammonium solution with ethanolic extracts of Centella asiatica, Syzygium cumini, Citrus sinensis, and Solanum trilobatum. The reaction mixture color changed from yellow to dark brown, and the resulting SNPs were irregular in shape with sizes ranging from 41 to 53 nm [71]. In contrast, SNPs synthesized from Ocimum sanctum began forming within 3 min, reaching optimal levels at 30 min. Absorption spectra recorded at 60, 90, and 120 min showed no significant differences, although the absorption peak bandwidth increased with nanoparticle polydispersity [102].
pH also plays a critical role in SNP synthesis. Lower pH values favor the formation of larger particles due to aggregation, whereas higher pH values promote nucleation and produce smaller, well-dispersed SNPs. For example, Garcinia mangostana extract produced large aggregates at pH 4, while highly dispersed small SNPs were obtained at pH 8. Neutral pH was found to be optimal for SNP formation [88]. Supporting this, Iravani Zolfaghari investigated SNP synthesis using pine bark extract with varying silver nitrate concentrations and phosphate buffers across a pH range of 3–11, confirming the significant influence of pH on nanoparticle size and dispersion [103].
The effect of pH on SNP synthesis has been widely reported. Using Acalypha indica extract, no SNP formation was observed under acidic conditions, whereas rapid color change and SNP formation occurred under alkaline conditions, with an absorption peak shifting to 500 nm; however, at pH 13, immediate aggregation was observed. At neutral pH, the reaction commenced upon addition of silver nitrate, with SNP formation completed within 30 min of incubation [104]. Similarly, when Crataegus douglasii fruit extract was used, the absorption wavelength decreased from pH 2–6 over 24 h. No SNPs were formed at pH 2 due to aggregation, while alkaline conditions significantly reduced aggregation [105]. For Solanum trilobatum extract, pH 5.8 inhibited nanoparticle synthesis due to limited accessibility of functional groups. At pH 8.8, a broad absorption band appeared at 480 nm, whereas at pH 7.8, a narrow peak at 440 nm corresponded to maximum SNP production [106].
Ndikau et al. synthesized spherical and stable silver nanoparticles (SNPs) using 0.001 M Citrullus lanatus extract and 250 g/L silver nitrate at 80 °C and pH 10 in a 4:5 ratio, obtaining nanoparticles with a diameter of 17.96 ± 0.16 nm [107]. Kumar et al. reported SNP synthesis at room temperature using 10% Annona squamosa extract and 1 mM silver nitrate for 4 h, producing spherical to irregular-shaped nanoparticles with an average size of 35 ± 5 nm [108]. Similarly, Das et al. prepared SNPs from Sesbania grandiflora leaves with 1 mM silver nitrate by incubating in the dark at 37 °C; the resulting spherical nanoparticles (10–25 nm) remained stable at room temperature for six months [109].
The effect of temperature on SNP synthesis has been extensively studied. Using Solanum trilobatum, no SPR band was observed at 20 °C, whereas distinct SPR peaks appeared at 35, 45, and 70 °C, with the highest yield achieved at the highest temperature. The SPR peak shifted from 460 nm to 440 nm (blue shift), indicating that higher temperatures strongly influence nanoparticle properties [110]. Generally, increasing the temperature (25–150 °C) enhances SNP production, intensifies absorbance, and reduces nanoparticle size, resulting in a more pronounced SPR band [14]. Similarly, SNP synthesis from Garcinia mangostana leaf extract between 37 and 90 °C showed increased nanoparticle yield with rising temperature. Initially, particle size decreased due to reduced aggregation, but above 75 °C, crystal formation around the nuclei caused a decline in absorbance [111].
In the present study, the optimized conditions for R. communis silver nanoparticles (RcSNPs) synthesis were 1 mM AgNO3, 2 mL leaf extract, 25 °C, pH 11, and a 5 min reaction time. For A. barbadensis silver nanoparticles (AbSNPs), the optimal conditions were 2 mM AgNO3, 1 mL leaf extract, 25 °C, pH 12, and a 3 min reaction time.

4.2. Proposed Mechanism for RcSNPs and AbSNPs Synthesis

In this study, R. communis and A. barbadensis aqueous leaf extracts were used for the plant-mediated synthesis of silver nanoparticles (SNPs). The process is initiated by the phenolic compounds and flavonoids present in the extracts. Phytochemical analysis of R. communis extract revealed the presence of alkaloids, phenolic compounds, flavonoids, sugars, and proteins. Among these, phenolic compounds and flavonoids act as efficient reducing agents, whereas proteins and certain other phytochemicals serve as capping and stabilizing agents for SNPs [112]. The enol groups of phenolic compounds and flavonoids can release electrons through O–H bond cleavage, and these electrons reduce Ag+ to Ag0. Additionally, proteins in R. communis extract are expected to stabilize the nanoparticles by capping them [113]. FTIR analysis supports this mechanism, indicating the involvement of hydroxyl and amide groups in the reduction and capping of RcSNPs.
Previous studies indicate that metallic ion reduction can occur via oxidation of amino or hydroxyl groups [114], or through specific terpenoids and alkaloids present in A. barbadensis extract. In this case, hydroxyl groups in alkaloids, glycosides, and steroids transfer electrons to silver ions, facilitating SNP formation [115]. Reduction of metallic ions is often achieved through oxidation of O–H groups to carbonyl groups [116]. The resulting nanoparticles may bind to proteins released from the A. barbadensis extract, producing protein-capped SNPs.
Stability studies confirmed that SNPs synthesized using these plant extracts are more stable than those produced via conventional chemical methods. These findings are consistent with reports from other researchers, highlighting the effectiveness of plant-mediated synthesis in producing stable, protein-capped nanoparticles.

4.3. Characterization of SNPs

Dynamic light scattering (DLS) was employed to determine the particle size distribution of the synthesized nanoparticles, providing information on the hydrodynamic diameter, size distribution, and polydispersity index (PDI). The measured negative zeta potential values indicate that the nanoparticles are highly stable, and dispersity analysis confirmed that they are well distributed. These findings are consistent with previously reported values [117] and correlate with UV–vis spectroscopy results. According to Gengan et al., a zeta potential greater than +30 mV or less than –30 mV reflects high colloidal stability [118]. Negative values typically arise from the incorporation of bioactive molecules from the plant extract [119], which induce strong electrostatic repulsion between nanoparticles and prevent aggregation [120]. The observed strong negative zeta potential may result from the shielding effect of biologically active plant compounds [121].
Biogenic SNPs are further stabilized through capping by phenolic compounds present in the leaf extracts, which also facilitate the reduction of Ag+ ions. To identify the phenolic constituents involved in capping, FTIR analysis was performed on RcSNPs, AbSNPs, and their respective leaf extracts. The analysis confirmed the presence of plant-derived phenolic functional groups on the nanoparticle surface, contributing to the excellent stability profile of the prepared SNPs [122,123].
The morphology and size distribution of silver nanoparticles (SNPs) are critical determinants of their physicochemical properties and biological activities [124]. In this study, SEM revealed that the synthesized SNPs exhibited diverse morphologies, including spherical, cubic, triangular, and rectangular forms. While the particles were relatively well distributed, some agglomeration was observed, likely due to limited stabilization from weak capping agents. TEM analysis further confirmed that the majority of nanoparticles were spherical, consistent with UV–vis and SEM observations, and indicated smaller particle sizes (15–20 nm for RcSNPs and 23–28 nm for AbSNPs).
Recent studies have emphasized the importance of nanoparticle size and shape in biological applications. Nanoparticles within the 1–100 nm range, such as those obtained here, provide a high surface area to volume ratio, which enhances interactions with microbial cells and biological targets [125]. Spherical nanoparticles, in particular, exhibit lower surface energy and higher thermodynamic stability, contributing to uniform dispersion and consistent bioactivity [126]. The presence of triangular and cubic morphologies may further enhance antimicrobial and antioxidant activities due to higher surface energy and increased reactive sites [127].
Agglomeration, as observed in some particles, is a common challenge in green nanoparticle synthesis. It can reduce the effective surface area and potentially decrease biological efficacy [128]. Factors influencing agglomeration include the concentration and type of capping agents, pH, and ionic strength of the synthesis medium. Optimizing these parameters can enhance particle stability and maintain the desirable nanoscale features for biomedical applications [129].
Overall, the SEM and TEM analyses indicate that green synthesis using Ricinus communis and Aloe barbadensis leaf extracts produces predominantly spherical, uniformly sized SNPs with minor agglomeration. The diversity in shapes and nanoscale dimensions suggests a potential for enhanced biological activity, supporting their use in antimicrobial, antioxidant, anti-inflammatory, and anticancer applications [130]. Future work may focus on fine-tuning synthesis conditions to minimize agglomeration and improve nanoparticle stability.
Thermogravimetric analysis (TGA) and differential scanning calorimetry (DSC) were employed to assess the thermal stability and compositional characteristics of the silver nanoparticles (SNPs) synthesized using R. communis and A. barbadensis leaf extracts. TGA revealed an initial weight loss around 90 °C, indicative of the desorption of surface-bound biological compounds, such as proteins and polyphenols, which play a crucial role in the reduction and stabilization of silver ions during nanoparticle synthesis [131]. The DSC analysis demonstrated a single-stage endothermic peak corresponding to the decomposition of these organic stabilizers, aligning with the TGA results. This congruence suggests that the phytochemicals not only reduce Ag+ to Ag0 but also contribute to the thermal stability of the nanoparticles by forming a protective layer around them [132,133].
Collectively, the TGA and DSC analyses indicate that the synthesized SNPs possess good thermal stability up to approximately 90 °C, which is advantageous for their potential applications in various biomedical and industrial fields [134]. The presence of plant-derived stabilizing agents further underscores the eco-friendly nature of the green synthesis method employed.

4.4. Therapeutic Evaluation of SNPs

The antimicrobial activity of SNPs was evaluated against various pathogenic strains. The diameter of inhibition zones increased with higher SNP concentrations, indicating a dose-dependent effect [135]. Smaller nanoparticles, with their larger surface area, exhibited enhanced bactericidal activity [136]. The proposed mechanisms underlying the antimicrobial properties of plant-synthesized SNPs include (i) interaction of silver ions with thiol-containing proteins, leading to protein denaturation and loss of function; (ii) binding to negatively charged DNA, causing structural damage and inhibiting replication; (iii) inhibition of respiratory enzymes (e.g., hydrogenase II) and cofactors (e.g., NADH), disrupting cellular respiration; (iv) induction of reactive oxygen species (ROS) such as hydroxyl radicals and superoxide anions, interfering with the cell respiration cycle; and (v) direct penetration of silver ions into microbial cell walls, resulting in cell death [137,138,139]. Optimized RcSNPs and AbSNPs exhibited superior antimicrobial activity compared with previously reported R. communis and A. barbadensis-derived nanoparticles, attributed to their smaller size, improved stability, and spherical morphology [140,141,142,143,144].
The anti-inflammatory activity of SNPs was evaluated using the protein denaturation assay. Leaf extract–encapsulated SNPs significantly inhibited bovine serum albumin (BSA) denaturation, even at lower concentrations than standard anti-inflammatory agents. This effect is likely due to the synergistic action of SNPs and phytochemicals in R. communis and A. barbadensis, which also exhibited substantial antiprotease activity [145,146,147,148]. Protein denaturation is a well-established cause of inflammation, as it results in loss of biological activity and, in some arthritic conditions, the production of autoantigens. Denaturation involves alterations in hydrogen bonding, disulfide linkages, electrostatic interactions, and hydrophobicity [149].
The antioxidant potential of SNPs was evaluated using DPPH and H2O2 scavenging assays. Free radicals contribute to numerous human diseases, including arthritis, atherosclerosis, reperfusion injury, and cancer [150,151]. At higher concentrations, SNPs effectively scavenged DPPH radicals by electron donation, while also demonstrating greater reducing power than ascorbic acid. In the presence of H2O2, SNPs promoted ROS generation, including hydroxyl radicals, which could induce oxidative stress through mitochondrial pathways, cathepsin leakage, and potassium efflux [152,153,154,155]. These results align with prior reports on H2O2 scavenging by Abutilon indicum leaf extract [156].
SNPs also exhibited notable anti-tyrosinase activity. Tyrosinase, a copper-containing metalloenzyme, catalyzes the oxidation of tyrosine to dihydroxyphenylalanine (DOPA) and subsequently to DOPA quinone, which are key intermediates in melanin biosynthesis [157,158]. Overactivity of tyrosinase can lead to hyperpigmentation disorders, including melasma, freckles, and age spots. Biogenic RcSNPs and AbSNPs demonstrated significant inhibition of tyrosinase activity, which is likely attributed to the functional groups from plant phytochemicals capping the nanoparticle surfaces [159]. These capping molecules may interact with the copper active site of tyrosinase or with the substrate binding site, thereby preventing enzymatic oxidation of tyrosine. This dual mechanism—physical adsorption of the enzyme onto the nanoparticle surface and chemical interaction via bioactive functional groups-enhances the potential of these SNPs as natural skin-lightening or depigmenting agents in cosmetic and therapeutic applications [160].
The cytotoxicity and anticancer potential of RcSNPs and AbSNPs were assessed using the MTT assay, which demonstrated a significant reduction in cell viability in a dose-dependent manner. The SNPs induced apoptosis via mitochondrial membrane disruption and ROS production, activating the caspase-9/3 signaling pathway [161,162,163]. The release of cytochrome c from the mitochondrial intermembrane space further contributed to mitochondrial stress, while apoptosis-regulating genes Bcl-2 and Bax mediated the loss of mitochondrial integrity [164].
Compared with previously reported in vitro studies on R. communis and A. barbadensis, the cytotoxic effects observed in this study were more pronounced, indicating that these biogenic SNPs are promising candidates for anticancer applications [38,140,144,145,165,166]. The incorporation of phytochemicals during green synthesis likely enhances the nanoparticles’ stability and bioactivity, contributing to their improved therapeutic profile. To further illustrate the novelty and comparative advantages of the present RcSNPs and AbSNPs, Table 4 summarizes key physicochemical and biological parameters of similar plant-derived AgNPs reported in recent literature.
As summarized in Table 4, the nanoparticles synthesized in this work exhibit smaller and more uniform sizes, faster formation rates, and higher antimicrobial and cytotoxic efficacy than most previously reported green-synthesized AgNPs, underscoring the efficiency of R. communis and A. barbadensis extracts as dual reducing and stabilizing agents.
Overall, the study demonstrates that the phytochemical components of R. communis and A. barbadensis enable effective synthesis of stable, bioactive AgNPs with multifunctional properties. These findings align with recent advances in phytogenic nanoparticle research and highlight their potential in biomedical applications

5. Conclusions

This study demonstrates the efficient green synthesis of silver nanoparticles using Ricinus communis and Aloe barbadensis leaf extracts as reducing and capping agents. Optimized conditions yielded uniformly sized, spherical nanoparticles with enhanced stability. RcSNPs and AbSNPs exhibited potent antibacterial, antifungal, anti-inflammatory, antioxidant, anticoagulant, anti-tyrosinase, and cytotoxic activities, surpassing the efficacy of plant extracts and comparable reports. These findings position phytogenic SNPs as promising candidates for multifunctional biomedical and biopharmaceutical applications, offering broad therapeutic efficacy with reduced toxicity. Future work should validate their in vivo performance and explore integration into advanced drug delivery systems.

Author Contributions

Conceptualization, A.A., W.-X.T. and G.F.G.; methodology, A.A.; formal analysis and investigation, A.A.; resources, W.-X.T. and G.F.G.; data curation, A.A.; writing—original draft preparation, A.A.; writing—review and editing, A.A., and W.-X.T.; visualization, A.A.; supervision, W.-X.T. and G.F.G.; project administration, W.-X.T. and G.F.G.; funding acquisition, W.-X.T. All authors have read and agreed to the published version of the manuscript.

Funding

This work received no external funding.

Informed Consent Statement

Written informed consent has been obtained from the patients to publish this paper.

Data Availability Statement

The original data supporting the findings of this study are included in the article. Further inquiries can be directed to the corresponding author.

Acknowledgments

A.A. is grateful for scholarships and funding from the Shanxi Agricultural University, China and Wen-Xia tian to support this research work.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Hamida, R.S.; Abdelmeguid, N.E.; Ali, M.A.; Bin-Meferij, M.M.; Khalil, M.I. Synthesis of silver nanoparticles using a novel cyanobacteria Desertifilum sp. extract: Their antibacterial and cytotoxicity effects. Int. J. Nanomed. 2020, 15, 49–63. [Google Scholar] [CrossRef]
  2. Sriramulu, M.; Sumathi, S. Photocatalytic, antioxidant, antibacterial and anti-inflammatory activity of silver nanoparticles synthesised using forest and edible mushroom. Adv. Nat. Sci. Nanosci. Nanotechnol. 2017, 8, 045012. [Google Scholar] [CrossRef]
  3. Lateef, A.; Folarin, B.I.; Oladejo, S.M.; Akinola, P.O.; Beukes, L.S.; Gueguim-Kana, E.B. Characterization, antimicrobial, antioxidant, and anticoagulant activities of silver nanoparticles synthesized from Petiveria alliacea L. leaf extract. Prep. Biochem. Biotechnol. 2018, 48, 646–652. [Google Scholar] [CrossRef]
  4. Odeniyi, M.A.; Okumah, V.C.; Adebayo-Tayo, B.C.; Odeniyi, O.A. Green synthesis and cream formulations of silver nanoparticles of Nauclea latifolia (African peach) fruit extracts and evaluation of antimicrobial and antioxidant activities. Sustain. Chem. Pharm. 2020, 15, 100197. [Google Scholar] [CrossRef]
  5. Fatima, R.; Priya, M.; Indurthi, L.; Radhakrishnan, V.; Sudhakaran, R. Biosynthesis of silver nanoparticles using red algae Portieria hornemannii and its antibacterial activity against fish pathogens. Microb. Pathog. 2020, 138, 103780. [Google Scholar] [CrossRef]
  6. Wypij, M.; Czarnecka, J.; Świecimska, M.; Dahm, H.; Rai, M.; Golinska, P. Synthesis, characterization and evaluation of antimicrobial and cytotoxic activities of biogenic silver nanoparticles synthesized from Streptomyces xinghaiensis OF1 strain. World J. Microbiol. Biotechnol. 2018, 34, 23. [Google Scholar] [CrossRef]
  7. Govindappa, M.; Hemashekhar, B.; Arthikala, M.-K.; Rai, V.R.; Ramachandra, Y. Characterization, antibacterial, antioxidant, antidiabetic, anti-inflammatory and antityrosinase activity of green synthesized silver nanoparticles using Calophyllum tomentosum leaves extract. Results Phys. 2018, 9, 400–408. [Google Scholar] [CrossRef]
  8. Siddiquee, M.A.; ud din Parray, M.; Mehdi, S.H.; Alzahrani, K.A.; Alshehri, A.A.; Malik, M.A.; Patel, R. Green synthesis of silver nanoparticles from Delonix regia leaf extracts: In-vitro cytotoxicity and interaction studies with bovine serum albumin. Mater. Chem. Phys. 2020, 242, 122493. [Google Scholar] [CrossRef]
  9. Das, P.; Ghosal, K.; Jana, N.K.; Mukherjee, A.; Basak, P. Green synthesis and characterization of silver nanoparticles using belladonna mother tincture and its efficacy as a potential antibacterial and anti-inflammatory agent. Mater. Chem. Phys. 2019, 228, 310–317. [Google Scholar] [CrossRef]
  10. Ravichandran, V.; Vasanthi, S.; Shalini, S.; Shah, S.A.A.; Tripathy, M.; Paliwal, N. Green synthesis, characterization, antibacterial, antioxidant and photocatalytic activity of Parkia speciosa leaves extract mediated silver nanoparticles. Results Phys. 2019, 15, 102565. [Google Scholar] [CrossRef]
  11. Hamedi, S.; Shojaosadati, S.A. Rapid and green synthesis of silver nanoparticles using Diospyros lotus extract: Evaluation of their biological and catalytic activities. Polyhedron 2019, 171, 172–180. [Google Scholar] [CrossRef]
  12. Vanlalveni, C.; Lallianrawna, S.; Biswas, A.; Selvaraj, M.; Changmai, B.; Rokhum, S.L. Green synthesis of silver nanoparticles using plant extracts and their antimicrobial activities: A review of recent literature. RSC Adv. 2021, 11, 2804–2837. [Google Scholar] [CrossRef]
  13. Ahsan, A.; Farooq, M.A. Therapeutic potential of green synthesized silver nanoparticles loaded PVA hydrogel patches for wound healing. J. Drug Deliv. Sci. Technol. 2019, 54, 101308. [Google Scholar] [CrossRef]
  14. Shahzadi, S.; Fatima, S.; Shafiq, Z.; Janjua, M.R.S.A. A review on green synthesis of silver nanoparticles (SNPs) using plant extracts: A multifaceted approach in photocatalysis, environmental remediation, and biomedicine. RSC Adv. 2025, 15, 3858–3903. [Google Scholar] [CrossRef] [PubMed]
  15. Girón-Vázquez, N.; Gómez-Gutiérrez, C.; Soto-Robles, C.; Nava, O.; Lugo-Medina, E.; Castrejón-Sánchez, V.; Vilchis-Nestor, A.; Luque, P. Study of the effect of Persea americana seed in the green synthesis of silver nanoparticles and their antimicrobial properties. Results Phys. 2019, 13, 102142. [Google Scholar] [CrossRef]
  16. Thomas, B.; Vithiya, B.; Prasad, T.; Mohamed, S.; Magdalane, C.M.; Kaviyarasu, K.; Maaza, M. Antioxidant and photocatalytic activity of aqueous leaf extract mediated green synthesis of silver nanoparticles using Passiflora edulis f. flavicarpa. J. Nanosci. Nanotechnol. 2019, 19, 2640–2648. [Google Scholar] [CrossRef]
  17. Kumar, P.V.; Kalyani, R.; Veerla, S.C.; Kollu, P.; Shameem, U.; Pammi, S. Biogenic synthesis of stable silver nanoparticles via Asparagus racemosus root extract and their antibacterial efficacy towards human and fish bacterial pathogens. Mater. Res. Express 2019, 6, 104008. [Google Scholar] [CrossRef]
  18. Singh, P.; Ahn, S.; Kang, J.-P.; Veronika, S.; Huo, Y.; Singh, H.; Chokkaligam, M.; El-Agamy Farh, M.; Aceituno, V.C.; Kim, Y.J. In vitro anti-inflammatory activity of spherical silver nanoparticles and monodisperse hexagonal gold nanoparticles by fruit extract of Prunus serrulata: A green synthetic approach. Artif. Cells Nanomed. Biotechnol. 2018, 46, 2022–2032. [Google Scholar]
  19. Khosravi, A.; Zarepour, A.; Iravani, S.; Varma, R.S.; Zarrabi, A. Sustainable synthesis: Natural processes shaping the nanocircular economy. Environ. Sci. Nano 2024, 11, 688–707. [Google Scholar] [CrossRef]
  20. Zehra, S.H.; Ramzan, K.; Viskelis, J.; Viskelis, P.; Balciunaitiene, A. Advancements in Green Synthesis of Silver-Based Nanoparticles: Antimicrobial and Antifungal Properties in Various Films. Nanomaterials 2025, 15, 252. [Google Scholar] [CrossRef]
  21. Dhir, R.; Chauhan, S.; Subham, P.; Kumar, S.; Sharma, P.; Shidiki, A.; Kumar, G. Plant-mediated synthesis of silver nanoparticles: Unlocking their pharmacological potential–a comprehensive review. Front. Bioeng. Biotechnol. 2024, 11, 1324805. [Google Scholar] [CrossRef]
  22. Eker, F.; Akdaşçi, E.; Duman, H.; Bechelany, M.; Karav, S. Green synthesis of silver nanoparticles using plant extracts: A comprehensive review of physicochemical properties and multifunctional applications. Int. J. Mol. Sci. 2025, 26, 6222. [Google Scholar] [CrossRef]
  23. Ongtanasup, T.; Kamdenlek, P.; Manaspon, C.; Eawsakul, K. Green-synthesized silver nanoparticles from Zingiber officinale extract: Antioxidant potential, biocompatibility, anti-LOX properties, and in silico analysis. BMC Complement. Med. Ther. 2024, 24, 84. [Google Scholar] [CrossRef]
  24. Batra, N.; Dey, P. Untargeted metabolomics of Aloe volatiles: Implications in pathway enrichments for improved bioactivities. Heliyon 2025, 11, e42268. [Google Scholar] [CrossRef]
  25. Babu, A.T.; Antony, R. Green synthesis of silver doped nano metal oxides of zinc & copper for antibacterial properties, adsorption, catalytic hydrogenation & photodegradation of aromatics. J. Ind. Eng. Chem. 2019, 7, 102840. [Google Scholar]
  26. Khojasteh, D.; Kazerooni, M.; Salarian, S.; Kamali, R. Droplet impact on superhydrophobic surfaces: A review of recent developments. J. Ind. Eng. Chem. 2016, 42, 1–14. [Google Scholar] [CrossRef]
  27. Kumar, V.; Singh, D.K.; Mohan, S.; Gundampati, R.K.; Hasan, S.H. Photoinduced green synthesis of silver nanoparticles using aqueous extract of Physalis angulata and its antibacterial and antioxidant activity. J. Environ. Chem. Eng. 2017, 5, 744–756. [Google Scholar] [CrossRef]
  28. Salih, H.; Altameemi, R.; Al-Azawi, A. Antibacterial activity of silver nanoparticles prepared from Camellia sinensis extracts in multi-drug resistant Staphylococcus Aureus. Cell Physiol. Biochem. 2024, 58, 659–676. [Google Scholar]
  29. Sowmyya, T. Spectroscopic investigation on catalytic and bactericidal properties of biogenic silver nanoparticles synthesized using Soymida febrifuga aqueous stem bark extract. J. Environ. Chem. Eng. 2018, 6, 3590–3601. [Google Scholar] [CrossRef]
  30. Vickers, N.J. Animal communication: When i’m calling you, will you answer too? Curr. Biol. 2017, 27, R713–R715. [Google Scholar] [CrossRef]
  31. Khatun, M.; Khatun, Z.; Karim, M.R.; Habib, M.R.; Rahman, M.H.; Aziz, M.A. Green synthesis of silver nanoparticles using extracts of Mikania cordata leaves and evaluation of their antioxidant, antimicrobial and cytotoxic properties. Food Chem. Adv. 2023, 3, 100386. [Google Scholar] [CrossRef]
  32. Parlinska-Wojtan, M.; Kus-Liskiewicz, M.; Depciuch, J.; Sadik, O. Green synthesis and antibacterial effects of aqueous colloidal solutions of silver nanoparticles using camomile terpenoids as a combined reducing and capping agent. Bioprocess Biosyst. Eng. 2016, 39, 1213–1223. [Google Scholar] [CrossRef]
  33. Melo, P.S.; Massarioli, A.P.; Denny, C.; dos Santos, L.F.; Franchin, M.; Pereira, G.E.; de Souza Vieira, T.M.F.; Rosalen, P.L.; de Alencar, S.M. Winery by-products: Extraction optimization, phenolic composition and cytotoxic evaluation to act as a new source of scavenging of reactive oxygen species. Food Chem. 2015, 181, 160–169. [Google Scholar] [CrossRef]
  34. Raota, C.S.; Cerbaro, A.F.; Salvador, M.; Delamare, A.P.L.; Echeverrigaray, S.; da Silva Crespo, J.; da Silva, T.B.; Giovanela, M. Green synthesis of silver nanoparticles using an extract of Ives cultivar (Vitis labrusca) pomace: Characterization and application in wastewater disinfection. J. Environ. Chem. Eng. 2019, 7, 103383. [Google Scholar] [CrossRef]
  35. Behravan, M.; Panahi, A.H.; Naghizadeh, A.; Ziaee, M.; Mahdavi, R.; Mirzapour, A. Facile green synthesis of silver nanoparticles using Berberis vulgaris leaf and root aqueous extract and its antibacterial activity. Int. J. Biol. Macromol. 2019, 124, 148–154. [Google Scholar] [CrossRef]
  36. Azad, A.; Zafar, H.; Raza, F.; Sulaiman, M. Factors influencing the green synthesis of metallic nanoparticles using plant extracts: A comprehensive review. Pharm. Fronts 2023, 5, e117–e131. [Google Scholar] [CrossRef]
  37. Bhatnagar, S.; Aoyagi, H. Current overview of the mechanistic pathways and influence of physicochemical parameters on the microbial synthesis and applications of metallic nanoparticles. Bioprocess Biosyst. Eng. 2025, 48, 1779–1800. [Google Scholar] [CrossRef]
  38. Abomughaid, M.M.; Teibo, J.O.; Akinfe, O.A.; Adewolu, A.M.; Teibo, T.K.A.; Afifi, M.; Al-Farga, A.M.H.; Al-kuraishy, H.M.; Al-Gareeb, A.I.; Alexiou, A. A phytochemical and pharmacological review of Ricinus communis L. Discov. Appl. Sci. 2024, 6, 315. [Google Scholar]
  39. Matei, C.E.; Visan, A.I.; Cristescu, R. Aloe Vera Polysaccharides as Therapeutic Agents: Benefits Versus Side Effects in Biomedical Applications. Polysaccharides 2025, 6, 36. [Google Scholar] [CrossRef]
  40. Ramothloa, T.P.; Mkolo, N.M.; Motshudi, M.C.; Mphephu, M.M.; Makhafola, M.A.; Naidoo, C.M. Phytochemical Composition and Multifunctional Applications of Ricinus communis L.: Insights into Therapeutic, Pharmacological, and Industrial Potential. Molecules 2025, 30, 3214. [Google Scholar] [CrossRef]
  41. Rahmati, S.; Tavallali, V.; Jahromi, M.H.G. Comparison of pectin and Aloe vera extract in green synthesis and enhanced stability and antioxidant activity of calcium nanostructure. Appl. Organomet. Chem. 2025, 39, e70077. [Google Scholar] [CrossRef]
  42. Osman, A.I.; Zhang, Y.; Farghali, M.; Rashwan, A.K.; Eltaweil, A.S.; Abd El-Monaem, E.M.; Mohamed, I.M.; Badr, M.M.; Ihara, I.; Rooney, D.W. Synthesis of green nanoparticles for energy, biomedical, environmental, agricultural, and food applications: A review. Environ. Chem. Lett. 2024, 22, 841–887. [Google Scholar] [CrossRef]
  43. Sharma, N.K.; Vishwakarma, J.; Rai, S.; Alomar, T.S.; AlMasoud, N.; Bhattarai, A. Green route synthesis and characterization techniques of silver nanoparticles and their biological adeptness. ACS Omega 2022, 7, 27004–27020. [Google Scholar] [CrossRef]
  44. Trease, G.E.; Evans, W.C. Pharmacognosy, 12th ed.; Tindall & Co.: London, UK, 1983. [Google Scholar]
  45. Harborne, A. Phytochemical Methods a Guide to Modern Techniques of Plant Analysis; Springer Science & Business Media: Berlin, Germany, 1998. [Google Scholar]
  46. Sofowora, A. Research on medicinal plants and traditional medicine in Africa. J. Altern. Complement. Med. 1996, 2, 365–372. [Google Scholar] [CrossRef] [PubMed]
  47. Singh, R.; Singh, R.; Parihar, P.; Mani, J.V. Green synthesis of silver nanoparticles using Solanum sisymbriifolium leaf extract: Characterization and evaluation of antioxidant, antibacterial and photocatalytic degradation activities. Process Biochem. 2024, 143, 337–352. [Google Scholar] [CrossRef]
  48. Evans, W.C. Trease and Evans’ pharmacognosy. Gen. Pharmacol. 1997, 2, 291. [Google Scholar]
  49. Urquiaga, I.; Leighton, F. Plant polyphenol antioxidants and oxidative stress. Biol. Res. 2000, 33, 55–64. [Google Scholar] [CrossRef] [PubMed]
  50. Obadoni, B.; Ochuko, P. Phytochemical studies and comparative efficacy of the crude extracts of some haemostatic plants in Edo and Delta States of Nigeria. Glob. J. Pure Appl. Sci. 2002, 8, 203–208. [Google Scholar] [CrossRef]
  51. Al-Dhabi, N.A.; Ponmurugan, K.; Jeganathan, P.M. Development and validation of ultrasound-assisted solid-liquid extraction of phenolic compounds from waste spent coffee grounds. Ultrason. Sonochem. 2017, 34, 206–213. [Google Scholar] [CrossRef]
  52. Ojo, O.A.; Oyinloye, B.E.; Ojo, A.B.; Afolabi, O.B.; Peters, O.A.; Olaiya, O.; Fadaka, A.; Jonathan, J.; Osunlana, O. Green synthesis of silver nanoparticles (AgNPs) using Talinum triangulare (Jacq.) Willd. leaf extract and monitoring their antimicrobial activity. J. Bionanosci. 2017, 11, 292–296. [Google Scholar] [CrossRef]
  53. Kumar, B.; Smita, K.; Cumbal, L.; Debut, A. Green synthesis of silver nanoparticles using Andean blackberry fruit extract. Saudi J. Biol. Sci. 2017, 24, 45–50. [Google Scholar] [CrossRef]
  54. Alam, A.; Tanveer, F.; Khalil, A.T.; Zohra, T.; Khamlich, S.; Alam, M.M.; Salman, M.; Ali, M.; Ikram, A.; Shinwari, Z.K. Silver nanoparticles biosynthesized from secondary metabolite producing marine actinobacteria and evaluation of their biomedical potential. Antonie Van Leeuwenhoek 2021, 114, 1497–1516. [Google Scholar] [CrossRef]
  55. Yildirim, K.; Atas, C.; Simsek, E.; Coban, A.Y. The effect of inoculum size on antimicrobial susceptibility testing of Mycobacterium tuberculosis. Microbiol. Spectr. 2023, 11, e00319-23. [Google Scholar] [CrossRef]
  56. Behera, M.; Soren, S.; Singh, N.R.; Sipra, B.; Sethi, G.; Mahanandia, S.; Pradhan, B.; Bhanjadeo, M. Green Synthesis of Bovine Serum Albumin Conjugated Silver Nanoparticles at Different Temperatures and Evaluation of Their Antioxidant Activity. Indian J. Pharm. Sci. 2024, 86, 1690–1701. [Google Scholar] [CrossRef]
  57. Banik, S.; Biswas, S.; Karmakar, S. Extraction, purification, and activity of protease from the leaves of Moringa oleifera. F1000Research 2018, 7, 1151. [Google Scholar] [CrossRef] [PubMed]
  58. Prem, P.; Naveenkumar, S.; Kamaraj, C.; Ragavendran, C.; Priyadharsan, A.; Manimaran, K.; Alharbi, N.S.; Rarokar, N.; Cherian, T.; Sugumar, V. Valeriana jatamansi root extract a potent source for biosynthesis of silver nanoparticles and their biomedical applications, and photocatalytic decomposition. Green Chem. Lett. Rev. 2024, 17, 2305142. [Google Scholar] [CrossRef]
  59. Madeshwaran, K.; Venkatachalam, R. Green synthesis of bimetallic ZnO–CuO nanoparticles using Annona muricata l. extract: Investigation of antimicrobial, antioxidant, and anticancer properties. J. Ind. Eng. Chem. 2024, 140, 454–467. [Google Scholar] [CrossRef]
  60. Pathak, P.; Shukla, P.; Kanshana, J.S.; Jagavelu, K.; Sangwan, N.S.; Dwivedi, A.K.; Dikshit, M. Standardized root extract of Withania somnifera and Withanolide A exert moderate vasorelaxant effect in the rat aortic rings by enhancing nitric oxide generation. J. Ethnopharmacol. 2021, 278, 114296. [Google Scholar] [CrossRef]
  61. Alemu, B.; Molla, M.D.; Tezera, H.; Dekebo, A.; Asmamaw, T. Phytochemical composition and in vitro antioxidant and antimicrobial activities of Bersama abyssinica F. seed extracts. Sci. Rep. 2024, 14, 6345. [Google Scholar] [CrossRef]
  62. Chandrasekhar, N.; Vinay, S. Yellow colored blooms of Argemone mexicana and Turnera ulmifolia mediated synthesis of silver nanoparticles and study of their antibacterial and antioxidant activity. Appl. Nanosci. 2017, 7, 851–861. [Google Scholar] [CrossRef]
  63. Giridasappa, A.; Ismail, S.M.; Gopinath, S. Antioxidant, Bactericidal, Antihemolytic, and Anticancer Assessment Activities of Al2O3, SnO2, and Green Synthesized Ag and CuO NPs. In Handbook of Sustainable Materials: Modelling, Characterization, and Optimization; CRC Press: Boca Raton, FL, USA, 2023; pp. 367–398. [Google Scholar]
  64. Mughal, S.S. Role of silver nanoparticles in colorimetric detection of biomolecules. Authorea 2022. Preprint. [Google Scholar] [CrossRef]
  65. Singh, P.; Mijakovic, I. Green synthesis and antibacterial applications of gold and silver nanoparticles from Ligustrum vulgare berries. Sci. Rep. 2022, 12, 7902. [Google Scholar] [CrossRef] [PubMed]
  66. Tavan, M.; Hanachi, P.; Mirjalili, M.H.; Dashtbani-Roozbehani, A. Comparative assessment of the biological activity of the green synthesized silver nanoparticles and aqueous leaf extract of Perilla frutescens (L.). Sci. Rep. 2023, 13, 6391. [Google Scholar] [CrossRef] [PubMed]
  67. Alim-Al-Razy, M.; Bayazid, G.A.; Rahman, R.U.; Bosu, R.; Shamma, S.S. Silver nanoparticle synthesis, UV-Vis spectroscopy to find particle size and measure resistance of colloidal solution. J. Phys. Conf. Ser. 2020, 1706, 12020. [Google Scholar] [CrossRef]
  68. Pasieczna-Patkowska, S.; Cichy, M.; Flieger, J. Application of Fourier transform infrared (FTIR) spectroscopy in characterization of green synthesized nanoparticles. Molecules 2025, 30, 684. [Google Scholar] [CrossRef]
  69. Gao, C.; Lu, Z.; Liu, Y.; Zhang, Q.; Chi, M.; Cheng, Q.; Yin, Y. Highly stable silver nanoplates for surface plasmon resonance biosensing. Angew. Chem. Int. Ed. 2012, 51, 5629–5633. [Google Scholar] [CrossRef]
  70. Jia, Z.; Li, J.; Gao, L.; Yang, D.; Kanaev, A. Dynamic light scattering: A powerful tool for in situ nanoparticle sizing. Colloids Interfaces 2023, 7, 15. [Google Scholar] [CrossRef]
  71. Sun, Q.; Cai, X.; Li, J.; Zheng, M.; Chen, Z.; Yu, C.-P. Green synthesis of silver nanoparticles using tea leaf extract and evaluation of their stability and antibacterial activity. Colloids Surf. A Physicochem. Eng. Asp. 2014, 444, 226–231. [Google Scholar] [CrossRef]
  72. Sreelekha, E.; George, B.; Shyam, A.; Sajina, N.; Mathew, B. A comparative study on the synthesis, characterization, and antioxidant activity of green and chemically synthesized silver nanoparticles. J. Bionanosci. 2021, 11, 489–496. [Google Scholar] [CrossRef]
  73. Kumaran, N.S. In vitro anti-inflammatory activity of silver nanoparticle synthesized Avicennia marina (Forssk.) Vierh.: A green synthetic approach. Int. J. Green Pharm. 2018, 12, S528–S536. [Google Scholar]
  74. Yakop, F.; Abd Ghafar, S.A.; Yong, Y.K.; Saiful Yazan, L.; Mohamad Hanafiah, R.; Lim, V.; Eshak, Z. Silver nanoparticles Clinacanthus Nutans leaves extract induced apoptosis towards oral squamous cell carcinoma cell lines. Artif. Cells Nanomed. Biotechnol. 2018, 46, 131–139. [Google Scholar] [CrossRef]
  75. Nasrollahzadeh, M.; Issaabadi, Z.; Sajadi, S.M. Green synthesis of the Ag/Al2O3 nanoparticles using Bryonia alba leaf extract and their catalytic application for the degradation of organic pollutants. J. Mater. Sci. Mater. Electron. 2019, 30, 3847–3859. [Google Scholar] [CrossRef]
  76. Deshpande, B.; Sharma, D.; Pandey, B. Phytochemicals and antibacterial screening of Parthenium hysterophorus. Indian J. Sci. Res. 2017, 13, 199–202. [Google Scholar]
  77. Khan, T.; Ullah, N.; Khan, M.A.; Mashwani, Z.u.R.; Nadhman, A. Plant-based gold nanoparticles; a comprehensive review of the decade-long research on synthesis, mechanistic aspects and diverse applications. Adv. Colloid Interface Sci. 2019, 272, 102017. [Google Scholar] [CrossRef]
  78. Singh, R. Chemotaxonomy of medicinal plants: Possibilities and limitations. In Natural Products and Drug Discovery; Elsevier: Amsterdam, The Netherlands, 2018; pp. 119–136. [Google Scholar]
  79. Dharajiya, D.; Pagi, N.; Jasani, H.; Patel, P. Antimicrobial activity and phytochemical screening of Aloe vera (Aloe barbadensis Miller). Int. J. Curr. Microbiol. App. Sci. 2017, 6, 2152–2162. [Google Scholar]
  80. Baba, S.A.; Malik, S.A. Determination of total phenolic and flavonoid content, antimicrobial and antioxidant activity of a root extract of Arisaema jacquemontii Blume. J. Taibah Univ. Sci. 2015, 9, 449–454. [Google Scholar] [CrossRef]
  81. Chen, Z.; Świsłocka, R.; Choińska, R.; Marszałek, K.; Dąbrowska, A.; Lewandowski, W.; Lewandowska, H. Exploring the correlation between the molecular structure and biological activities of metal–phenolic compound complexes: Research and description of the role of metal ions in improving the antioxidant activities of phenolic compounds. Int. J. Mol. Sci. 2024, 25, 11775. [Google Scholar] [CrossRef]
  82. Dakshayani, S.; Marulasiddeshwara, M.; Sharath Kumar, M.; Raghavendra Kumar, P.; Devaraja, S. Antimicrobial, anticoagulant and antiplatelet activities of green synthesized silver nanoparticles using Selaginella (Sanjeevini) plant extract. Int. J. Biol. Macromol. 2019, 131, 787–797. [Google Scholar] [CrossRef] [PubMed]
  83. Song, X.-C.; Canellas, E.; Asensio, E.; Nerín, C. Predicting the antioxidant capacity and total phenolic content of bearberry leaves by data fusion of UV–Vis spectroscopy and UHPLC/Q-TOF-MS. Talanta 2020, 213, 120831. [Google Scholar] [CrossRef] [PubMed]
  84. Lama-Muñoz, A.; del Mar Contreras, M.; Espínola, F.; Moya, M.; Romero, I.; Castro, E. Content of phenolic compounds and mannitol in olive leaves extracts from six Spanish cultivars: Extraction with the Soxhlet method and pressurized liquids. Food Chem. 2020, 320, 126626. [Google Scholar] [CrossRef]
  85. Ahmed-Gaid, K.; Ahmed-Gaid, N.E.Y.; Hanoun, S.; Chenna, H.; Ghenaiet, K.; Ahmed-Gaid, H.; Hadef, Y. Phenolic compounds, antioxidant, and antimicrobial activities of methanolic extract from the saharan endemic plant haloxylon scoparium pomel. Bull. Pharm. Sci. Assiut Univ. 2025, 48, 267–276. [Google Scholar] [CrossRef]
  86. Marius, L.; Kini, F.; Pierre, D.; Pierre, G.I. In vitro antioxidant activity and phenolic contents of different fractions of ethanolic extract from Khaya senegalensis A. Juss. (Meliaceae) stem barks. Afr. J. Pharm. Pharmacol. 2016, 10, 501–507. [Google Scholar] [CrossRef]
  87. Khawory, M.H.; Amanah, A.; Salin, N.H.; Mohd Subki, M.F.; Mohd Nasim, N.N.; Noordin, M.I.; Wahab, H.A. Effects of Gamma Radiation Treatment on Medicinal Plants for Both Qualitative and Quantitative Analysis of Phenolic Acid. Available online: https://ssrn.com/abstract=4334860 (accessed on 25 January 2023).
  88. Veerasamy, R.; Xin, T.Z.; Gunasagaran, S.; Xiang, T.F.W.; Yang, E.F.C.; Jeyakumar, N.; Dhanaraj, S.A. Biosynthesis of silver nanoparticles using mangosteen leaf extract and evaluation of their antimicrobial activities. J. Saudi Chem. Soc. 2011, 15, 113–120. [Google Scholar] [CrossRef]
  89. Banerjee, P.; Satapathy, M.; Mukhopahayay, A.; Das, P. Leaf extract mediated green synthesis of silver nanoparticles from widely available Indian plants: Synthesis, characterization, antimicrobial property and toxicity analysis. Bioresour. Bioprocess. 2014, 1, 3. [Google Scholar] [CrossRef]
  90. Chandran, N.; Bayal, M.; Pilankatta, R.; Nair, S.S. Tuning of surface Plasmon resonance (SPR) in metallic nanoparticles for their applications in SERS. In Nanomaterials for Luminescent Devices, Sensors, and Bio-Imaging Applications; Springer: Berlin/Heidelberg, Germany, 2021; pp. 39–66. [Google Scholar]
  91. Yadav, M.; Gaur, N.; Wahi, N.; Singh, S.; Kumar, K.; Amoozegar, A.; Sharma, E. Phytochemical-Assisted Fabrication of Biogenic Silver Nanoparticles from Vitex negundo: Structural Features, Antibacterial Activity, and Cytotoxicity Evaluation. Colloids Interfaces 2025, 9, 55. [Google Scholar] [CrossRef]
  92. Kale, G.; Bhatkar, D.; Rokade, S.; Ingle, P.; Patil, R. Green synthesis of silver nanoparticles using Azadirachta indica leaves extract and characterization by UV. Sustain. Dev. 2022, 1695, 1695–1699. [Google Scholar]
  93. Markus, J.; Wang, D.; Kim, Y.-J.; Ahn, S.; Mathiyalagan, R.; Wang, C.; Yang, D.C. Biosynthesis, characterization, and bioactivities evaluation of silver and gold nanoparticles mediated by the roots of Chinese herbal Angelica pubescens Maxim. Nanoscale Res. Lett. 2017, 12, 46. [Google Scholar] [CrossRef] [PubMed]
  94. Kaabipour, S.; Hemmati, S. A review on the green and sustainable synthesis of silver nanoparticles and one-dimensional silver nanostructures. Beilstein J. Nanotechnol. 2021, 12, 102–136. [Google Scholar] [CrossRef]
  95. Chirumamilla, P.; Dharavath, S.B.; Taduri, S. Eco-friendly green synthesis of silver nanoparticles from leaf extract of Solanum khasianum: Optical properties and biological applications. Appl. Biochem. Biotechnol. 2023, 195, 353–368. [Google Scholar] [CrossRef] [PubMed]
  96. Sinha, S.N.; Paul, D.; Halder, N.; Sengupta, D.; Patra, S.K. Green synthesis of silver nanoparticles using fresh water green alga Pithophora oedogonia (Mont.) Wittrock and evaluation of their antibacterial activity. Appl. Nanosci. 2015, 5, 703–709. [Google Scholar] [CrossRef]
  97. Sasikala, A.; Linga Rao, M.; Savithramma, N.; Prasad, T. Synthesis of silver nanoparticles from stem bark of Cochlospermum religiosum (L.) Alston: An important medicinal plant and evaluation of their antimicrobial efficacy. Appl. Nanosci. 2015, 5, 827–835. [Google Scholar] [CrossRef]
  98. Mariadoss, A.V.A.; Ramachandran, V.; Shalini, V.; Agilan, B.; Franklin, J.H.; Sanjay, K.; Alaa, Y.G.; Tawfiq, M.A.-A.; Ernest, D. Green synthesis, characterization and antibacterial activity of silver nanoparticles by Malus domestica and its cytotoxic effect on (MCF-7) cell line. Microb. Pathog. 2019, 135, 103609. [Google Scholar] [CrossRef]
  99. Ahsan, A.; Farooq, M.A.; Ahsan Bajwa, A.; Parveen, A. Green synthesis of silver nanoparticles using Parthenium hysterophorus: Optimization, characterization and in vitro therapeutic evaluation. Molecules 2020, 25, 3324. [Google Scholar] [CrossRef]
  100. Macovei, I.; Luca, S.V.; Skalicka-Woźniak, K.; Horhogea, C.E.; Rimbu, C.M.; Sacarescu, L.; Vochita, G.; Gherghel, D.; Ivanescu, B.L.; Panainte, A.D. Silver nanoparticles synthesized from abies alba and pinus sylvestris bark extracts: Characterization, antioxidant, cytotoxic, and antibacterial effects. Antioxidants 2023, 12, 797. [Google Scholar] [CrossRef]
  101. Halawani, E.M. Rapid biosynthesis method and characterization of silver nanoparticles using Zizyphus spina christi leaf extract and their antibacterial efficacy in therapeutic application. J. Biomater. Nanobiotechnol. 2016, 8, 22–35. [Google Scholar] [CrossRef]
  102. Logeswari, P.; Silambarasan, S.; Abraham, J. Ecofriendly synthesis of silver nanoparticles from commercially available plant powders and their antibacterial properties. Sci. Iran. 2013, 20, 1049–1054. [Google Scholar]
  103. Iravani, S.; Zolfaghari, B. Green synthesis of silver nanoparticles using Pinus eldarica bark extract. Biomed Res. Int. 2013, 2013, 639725. [Google Scholar] [CrossRef] [PubMed]
  104. Madaniyah, L.; Fiddaroini, S.; Hayati, E.K.; Rahman, M.F.; Sabarudin, A. Biosynthesis, characterization, and in-vitro anticancer effect of plant-mediated silver nanoparticles using Acalypha indica Linn: In-silico approach. OpenNano 2025, 21, 100220. [Google Scholar] [CrossRef]
  105. Alirezalu, A.; Ahmadi, N.; Salehi, P.; Sonboli, A.; Alirezalu, K.; Mousavi Khaneghah, A.; Barba, F.J.; Munekata, P.E.; Lorenzo, J.M. Physicochemical characterization, antioxidant activity, and phenolic compounds of hawthorn (Crataegus spp.) fruits species for potential use in food applications. Foods 2020, 9, 436. [Google Scholar] [CrossRef]
  106. Arumugam, S.; Ramesh, P. Green synthesis and characterization of zirconium oxide nanoparticles using solanum trilobatum and its photodegradation activity. Sustain. Chem. Clim. Action 2025, 6, 100086. [Google Scholar] [CrossRef]
  107. Ndikau, M.; Noah, N.M.; Andala, D.M.; Masika, E. Green synthesis and characterization of silver nanoparticles using Citrullus lanatus fruit rind extract. Int. J. Anal. Chem. 2017, 2017, 8108504. [Google Scholar] [CrossRef]
  108. Kumar, R.; Roopan, S.M.; Prabhakarn, A.; Khanna, V.G.; Chakroborty, S. Agricultural waste Annona squamosa peel extract: Biosynthesis of silver nanoparticles. Spectrochim. Acta A Mol. Biomol. Spectrosc. 2012, 90, 173–176. [Google Scholar] [CrossRef]
  109. Das, J.; Das, M.P.; Velusamy, P. Sesbania grandiflora leaf extract mediated green synthesis of antibacterial silver nanoparticles against selected human pathogens. Spectrochim. Acta A Mol. Biomol. Spectrosc. 2013, 104, 265–270. [Google Scholar] [CrossRef]
  110. Vanaja, M.; Paulkumar, K.; Gnanajobitha, G.; Rajeshkumar, S.; Malarkodi, C.; Annadurai, G. Herbal plant synthesis of antibacterial silver nanoparticles by Solanum trilobatum and its characterization. Int. J. Met. 2014, 2014, 692461. [Google Scholar] [CrossRef]
  111. Sarip, N.A.; Aminudin, N.I.; Danial, W.H. Green synthesis of metal nanoparticles using Garcinia extracts: A review. Environ. Chem. Lett. 2022, 20, 469–493. [Google Scholar] [CrossRef]
  112. Patel, J.; Kumar, G.S.; Roy, H.; Maddiboyina, B.; Leporatti, S.; Bohara, R.A. From nature to nanomedicine: Bioengineered metallic nanoparticles bridge the gap for medical applications. Discov. Nano 2024, 19, 85. [Google Scholar] [CrossRef]
  113. Mintiwab, A.; Jeyaramraja, P. Evaluation of phytochemical components, antioxidant and antibacterial activities of silver nanoparticles synthesized using Ricinus communis leaf extracts. Vegetos 2021, 34, 606–618. [Google Scholar] [CrossRef]
  114. Lim, S.H.; Park, Y. Green synthesis, characterization and catalytic activity of gold nanoparticles prepared using rosmarinic acid. J. Nanosci. Nanotechnol. 2018, 18, 659–667. [Google Scholar] [CrossRef] [PubMed]
  115. Adebayo-Tayo, B.C.; Akinsete, T.O.; Odeniyi, O.A. Phytochemical composition and comparative evaluation of antimicrobial activities of the juice extract of Citrus aurantifolia and its silver nanoparticles. Niger. J. Pharm. Res. 2016, 12, 59–64. [Google Scholar]
  116. Dawadi, S.; Katuwal, S.; Gupta, A.; Lamichhane, U.; Thapa, R.; Jaisi, S.; Lamichhane, G.; Bhattarai, D.P.; Parajuli, N. Current research on silver nanoparticles: Synthesis, characterization, and applications. J. Nanomater. 2021, 2021, 6687290. [Google Scholar] [CrossRef]
  117. Pochapski, D.J.; Carvalho dos Santos, C.; Leite, G.W.; Pulcinelli, S.H.; Santilli, C.V. Zeta potential and colloidal stability predictions for inorganic nanoparticle dispersions: Effects of experimental conditions and electrokinetic models on the interpretation of results. Langmuir 2021, 37, 13379–13389. [Google Scholar] [CrossRef]
  118. Gengan, R.; Anand, K.; Phulukdaree, A.; Chuturgoon, A. A549 lung cell line activity of biosynthesized silver nanoparticles using Albizia adianthifolia leaf. Colloids Surf. B Biointerfaces 2013, 105, 87–91. [Google Scholar] [CrossRef] [PubMed]
  119. Varadavenkatesan, T.; Selvaraj, R.; Vinayagam, R. Phyto-synthesis of silver nanoparticles from Mussaenda erythrophylla leaf extract and their application in catalytic degradation of methyl orange dye. J. Mol. Liq. 2016, 221, 1063–1070. [Google Scholar] [CrossRef]
  120. Padalia, H.; Moteriya, P.; Chanda, S. Green synthesis of silver nanoparticles from marigold flower and its synergistic antimicrobial potential. Arab. J. Chem. 2015, 8, 732–741. [Google Scholar] [CrossRef]
  121. Salih, A.M.; Al-Qurainy, F.; Khan, S.; Nadeem, M.; Tarroum, M.; Shaikhaldein, H.O. Biogenic silver nanoparticles improve bioactive compounds in medicinal plant Juniperus procera in vitro. Front. Plant Sci. 2022, 13, 962112. [Google Scholar] [CrossRef]
  122. Kumar, D.; Salvi, A.; Thakur, V.; Kharbanda, S.; Sangwan, S.; Thakur, P.; Thakur, A. Aloe vera assisted green synthesis of silver nanoparticles: Structural characterization, electrochemical behaviour, and antimicrobial efficiency. Discov. Appl. Sci. 2025, 7, 530. [Google Scholar] [CrossRef]
  123. Modrzejewska-Sikorska, A.; Konował, E. Silver and gold nanoparticles as chemical probes of the presence of heavy metal ions. J. Mol. Liq. 2020, 302, 112559. [Google Scholar] [CrossRef]
  124. Awaly, S.B.; Osman, N.H.; Farag, H.M.; Yacoub, I.H.; Mahmoud-Aly, M.; Elarabi, N.I.E.; Ahmed, D.S. Evaluation of the morpho-physiological traits and the genetic diversity of durum wheat’s salt tolerance induced by silver nanoparticles. J. Agric. Environ. Int. Dev. 2023, 117, 161–184. [Google Scholar]
  125. Dubey, R.K.; Shukla, S.; Hussain, Z. Green synthesis of silver nanoparticles; A sustainable approach with diverse applications. Zhongguo Ying Yong Sheng Li Xue Za Zhi 2023, 39, e20230007. [Google Scholar] [CrossRef]
  126. Anjana, V.; Joseph, M.; Francis, S.; Joseph, A.; Koshy, E.P.; Mathew, B. Microwave assisted green synthesis of silver nanoparticles for optical, catalytic, biological and electrochemical applications. Artif. Cells Nanomed. Biotechnol. 2021, 49, 438–449. [Google Scholar] [CrossRef]
  127. Fahim, M.; Shahzaib, A.; Nishat, N.; Jahan, A.; Bhat, T.A.; Inam, A. Green synthesis of silver nanoparticles: A comprehensive review of methods, influencing factors, and applications. JCIS Open 2024, 16, 100125. [Google Scholar] [CrossRef]
  128. Hussain, I.; Singh, N.; Singh, A.; Singh, H.; Singh, S. Green synthesis of nanoparticles and its potential application. Biotechnol. Lett. 2016, 38, 545–560. [Google Scholar] [CrossRef]
  129. Haridas, E.; Varma, M.; Chandra, G.K. Bioactive silver nanoparticles derived from Carica papaya floral extract and its dual-functioning biomedical application. Sci. Rep. 2025, 15, 1–14. [Google Scholar] [CrossRef]
  130. Aminnezhad, S.; Zonobian, M.A.; Moradi Douki, M.; Mohammadi, M.R.; Azarakhsh, Y. Curcumin and their derivatives with anti-inflammatory, neuroprotective, anticancer, and antimicrobial activities: A review. Micro Nano Bio Asp. 2023, 2, 25–34. [Google Scholar]
  131. Stanoiu, A.-M.; Bejenaru, C.; Segneanu, A.-E.; Vlase, G.; Bradu, I.A.; Vlase, T.; Mogoşanu, G.D.; Ciocîlteu, M.V.; Biţă, A.; Kostici, R. Design and Evaluation of a Inonotus obliquus–AgNP–Maltodextrin Delivery System: Antioxidant, Antimicrobial, Acetylcholinesterase Inhibitory and Cytotoxic Potential. Polymers 2025, 17, 2163. [Google Scholar] [CrossRef] [PubMed]
  132. Yadav, S.; Kumari, P.; Mahto, S.K. Green synthesis of silver nanoparticles using Momordica dioica leaf extract: Characterizing antibacterial, antibiofilm, and catalytic activities. Transit. Met. Chem. 2025, 50, 1151–1168. [Google Scholar] [CrossRef]
  133. Mohanaparameswari, S.; Balachandramohan, M.; Kumar, K.G.; Revathy, M.; Sasikumar, P.; Rajeevgandhi, C.; Vimalan, M.; Pugazhendhi, S.; Batoo, K.M.; Hussain, S. Green synthesis of silver oxide nanoparticles using Plectranthus amboinicus and Solanum trilobatum extracts as an Eco-friendly approach: Characterization and antibacterial properties. J. Inorg. Organomet. Polym. 2024, 34, 3191–3211. [Google Scholar] [CrossRef]
  134. Leyva-Porras, C.; Cruz-Alcantar, P.; Espinosa-Solís, V.; Martínez-Guerra, E.; Piñón-Balderrama, C.I.; Compean Martínez, I.; Saavedra-Leos, M.Z. Application of differential scanning calorimetry (DSC) and modulated differential scanning calorimetry (MDSC) in food and drug industries. Polymers 2019, 12, 5. [Google Scholar] [CrossRef] [PubMed]
  135. Agnihotri, S.; Mukherji, S.; Mukherji, S. Size-controlled silver nanoparticles synthesized over the range 5–100 nm using the same protocol and their antibacterial efficacy. Rsc. Adv. 2014, 4, 3974–3983. [Google Scholar] [CrossRef]
  136. Karuna, D.; Dey, P.; Das, S.; Kundu, A.; Bhakta, T. In vitro antioxidant activities of root extract of Asparagus racemosus Linn. J. Tradit. Complement. Med. 2018, 8, 60–65. [Google Scholar] [CrossRef]
  137. Rajeshkumar, S.; Bharath, L. Mechanism of plant-mediated synthesis of silver nanoparticles–a review on biomolecules involved, characterisation and antibacterial activity. Chem. Biol. Interact. 2017, 273, 219–227. [Google Scholar] [CrossRef] [PubMed]
  138. El Badawy, A.M.; Silva, R.G.; Morris, B.; Scheckel, K.G.; Suidan, M.T.; Tolaymat, T.M. Surface charge-dependent toxicity of silver nanoparticles. Environ. Sci. Technol. 2011, 45, 283–287. [Google Scholar] [CrossRef]
  139. Shayo, G.M.; Elimbinzi, E.; Shao, G.N. Preparation methods, applications, toxicity and mechanisms of silver nanoparticles as bactericidal agent and superiority of green synthesis method. Heliyon 2024, 10, e36539. [Google Scholar] [CrossRef] [PubMed]
  140. Tippayawat, P.; Phromviyo, N.; Boueroy, P.; Chompoosor, A. Green synthesis of silver nanoparticles in aloe vera plant extract prepared by a hydrothermal method and their synergistic antibacterial activity. PeerJ 2016, 4, e2589. [Google Scholar] [CrossRef] [PubMed]
  141. Vélez, E.; Campillo, G.; Morales, G.; Hincapié, C.; Osorio, J.; Arnache, O. Silver nanoparticles obtained by aqueous or ethanolic aloe Vera extracts: An assessment of the antibacterial activity and mercury removal capability. J. Nanomater. 2018, 2018, 7215210. [Google Scholar] [CrossRef]
  142. Haripriya, S.; Ajitha, P. Antimicrobial efficacy of silver nanoparticles of Aloe vera. J. Adv. Pharm. Educ. Res. 2017, 7, 163–166. [Google Scholar] [CrossRef]
  143. Urnukhsaikhan, E.; Bold, B.-E.; Gunbileg, A.; Sukhbaatar, N.; Mishig-Ochir, T. Antibacterial activity and characteristics of silver nanoparticles biosynthesized from Carduus crispus. Sci. Rep. 2021, 11, 21047. [Google Scholar] [CrossRef]
  144. Ojha, S.; Sett, A.; Bora, U. Green synthesis of silver nanoparticles by Ricinus communis var. carmencita leaf extract and its antibacterial study. Adv. Nat. Sci. Nanosci. Nanotechnol. 2017, 8, 035009. [Google Scholar] [CrossRef]
  145. Nemudzivhadi, V.; Masoko, P. In vitro assessment of cytotoxicity, antioxidant, and anti-inflammatory activities of Ricinus communis (Euphorbiaceae) leaf extracts. Evid. Based Complement. Altern. Med. 2014, 2014, 625961. [Google Scholar] [CrossRef]
  146. Valderramas, A.C.; Moura, S.H.P.; Couto, M.; Pasetto, S.; Chierice, G.O.; Guimarães, S.A.C.; Zurron, A.C.B.d.P. Antiinflammatory activity of Ricinus communis derived polymer. Braz. J. Oral Sci. 2008, 7, 1666–1672. [Google Scholar]
  147. Varghese, R.M.; Kumar, A.; Shanmugam, R. Comparative anti-inflammatory activity of silver and zinc oxide nanoparticles synthesized using Ocimum tenuiflorum and Ocimum gratissimum herbal formulations. Cureus 2024, 16, e52995. [Google Scholar] [CrossRef]
  148. Khashan, A.A.; Dawood, Y.; Khalaf, Y.H. Green chemistry and anti-inflammatory activity of silver nanoparticles using an aqueous curcumin extract. Results Chem. 2023, 5, 100913. [Google Scholar] [CrossRef]
  149. Dharmadeva, S.; Galgamuwa, L.S.; Prasadinie, C.; Kumarasinghe, N. In vitro anti-inflammatory activity of Ficus racemosa L. bark using albumin denaturation method. AYU 2018, 39, 239–242. [Google Scholar]
  150. Matkowski, A.; Tasarz, P.; Szypuła, E. Antioxidant activity of herb extracts from five medicinal plants from Lamiaceae, subfamily Lamioideae. J. Med. Plants Res. 2008, 2, 321–330. [Google Scholar]
  151. Silva, F.; Veiga, F.; Cardoso, C.; Dias, F.; Cerqueira, F.; Medeiros, R.; Paiva-Santos, A.C. A rapid and simplified DPPH assay for analysis of antioxidant interactions in binary combinations. Microchem. J. 2024, 202, 110801. [Google Scholar] [CrossRef]
  152. Sohal, J.K.; Saraf, A.; Shukla, K.; Shrivastava, M. Determination of antioxidant potential of biochemically synthesized silver nanoparticles using Aloe vera gel extract. Plant Sci. Today 2019, 6, 208–217. [Google Scholar] [CrossRef]
  153. More, G.K.; Makola, R.T. In-vitro analysis of free radical scavenging activities and suppression of LPS-induced ROS production in macrophage cells by Solanum sisymbriifolium extracts. Sci. Rep. 2020, 10, 6493. [Google Scholar] [CrossRef] [PubMed]
  154. Rana, M.S.; Rayhan, N.M.A.; Emon, M.S.H.; Islam, M.T.; Rathry, K.; Hasan, M.M.; Mansur, M.M.I.; Srijon, B.C.; Islam, M.S.; Ray, A. Antioxidant activity of Schiff base ligands using the DPPH scavenging assay: An updated review. RSC Adv. 2024, 14, 33094–33123. [Google Scholar] [CrossRef]
  155. He, W.; Zhou, Y.-T.; Wamer, W.G.; Boudreau, M.D.; Yin, J.-J. Mechanisms of the pH dependent generation of hydroxyl radicals and oxygen induced by Ag nanoparticles. Biomaterials 2012, 33, 7547–7555. [Google Scholar] [CrossRef]
  156. Mata, R.; Nakkala, J.R.; Sadras, S.R. Biogenic silver nanoparticles from Abutilon indicum: Their antioxidant, antibacterial and cytotoxic effects in vitro. Colloids Surf. B Biointerfaces 2015, 128, 276–286. [Google Scholar] [CrossRef]
  157. Irfan, A.; Jardan, Y.A.B.; Rubab, L.; Hameed, H.; Zahoor, A.F.; Supuran, C.T. Bacterial tyrosinases and their inhibitors. Enzymes 2024, 56, 231–260. [Google Scholar]
  158. Mukherjee, P.K.; Biswas, R.; Sharma, A.; Banerjee, S.; Biswas, S.; Katiyar, C. Validation of medicinal herbs for anti-tyrosinase potential. J. Herb. Med. 2018, 14, 1–16. [Google Scholar] [CrossRef]
  159. Özer, Ö.; Mutlu, B.; Kıvçak, B. Antityrosinase activity of some plant extracts and formulations containing ellagic acid. Pharm. Biol. 2007, 45, 519–524. [Google Scholar] [CrossRef]
  160. Manickam, V.; Mani, G.; Muthuvel, R.; Pushparaj, H.; Jayabalan, J.; Pandit, S.S.; Elumalai, S.; Kaliappan, K.; Tae, J.H. Green fabrication of silver nanoparticles and it’s in vitro anti-bacterial, anti-biofilm, free radical scavenging and mushroom tyrosinase efficacy evaluation. Inorg. Chem. Commun. 2024, 162, 112199. [Google Scholar] [CrossRef]
  161. Verma, S.K.; Jha, E.; Sahoo, B.; Panda, P.K.; Thirumurugan, A.; Parashar, S.; Suar, M. Mechanistic insight into the rapid one-step facile biofabrication of antibacterial silver nanoparticles from bacterial release and their biogenicity and concentration-dependent in vitro cytotoxicity to colon cells. RSC Adv. 2017, 7, 40034–40045. [Google Scholar] [CrossRef]
  162. Akter, M.; Sikder, M.T.; Rahman, M.M.; Ullah, A.A.; Hossain, K.F.B.; Banik, S.; Hosokawa, T.; Saito, T.; Kurasaki, M. A systematic review on silver nanoparticles-induced cytotoxicity: Physicochemical properties and perspectives. J. Adv. Res. 2018, 9, 1–16. [Google Scholar] [CrossRef] [PubMed]
  163. Sati, A.; Mali, S.N.; Ranade, T.N.; Yadav, S.; Pratap, A. Silver Nanoparticles (AgNPs) as a Double-Edged Sword: Synthesis, Factors Affecting, Mechanisms of Toxicity and Anticancer Potentials—An Updated Review till March 2025. Biol. Trace Elem. Res. 2025, 1–52, in press. [Google Scholar] [CrossRef] [PubMed]
  164. Garrido, C.; Galluzzi, L.; Brunet, M.; Puig, P.; Didelot, C.; Kroemer, G. Mechanisms of cytochrome c release from mitochondria. Cell Death Differ. 2006, 13, 1423–1433. [Google Scholar] [CrossRef]
  165. Abbas, M.; Ali, A.; Arshad, M.; Atta, A.; Mehmood, Z.; Tahir, I.M.; Iqbal, M. Mutagenicity, cytotoxic and antioxidant activities of Ricinus communis different parts. Chem. Cent. J. 2018, 12, 3. [Google Scholar] [CrossRef]
  166. Erhabor, J.O.; Idu, M. In vitro antibacterial, antioxidant, cytogenotoxic and nutritional value of Aloe barbadensis Mill.(Asphodelaceae). J. Microbiol. Biotechnol. Food Sci. 2020, 10, 10–21. [Google Scholar] [CrossRef]
  167. Pawar, A.A.; Sahoo, J.; Verma, A.; Alswieleh, A.M.; Lodh, A.; Raut, R.; Lakkakula, J.; Jeon, B.-H.; Islam, M.R. Azadirachta indica-Derived Silver Nanoparticle Synthesis and Its Antimicrobial Applications. J. Nanomater. 2022, 2022, 4251229. [Google Scholar] [CrossRef]
  168. Khac, K.T.; Phu, H.H.; Thi, H.T.; Thuy, V.D.; Do Thi, H. Biosynthesis of silver nanoparticles using tea leaf extract (Camellia sinensis) for photocatalyst and antibacterial effect. Heliyon 2023, 9, e20707. [Google Scholar] [CrossRef] [PubMed]
  169. Mohammed, A.A.; Mohamed, A.; El-Naggar, N.E.-A.; Mahrous, H.; Nasr, G.M.; Abdella, A.; Ahmed, R.H.; Irmak, S.; Elsayed, M.S.; Selim, S. Antioxidant and antibacterial activities of silver nanoparticles biosynthesized by Moringa Oleifera through response surface methodology. J. Nanomater. 2022, 2022, 9984308. [Google Scholar] [CrossRef]
  170. Gautam, D.; Dolma, K.G.; Khandelwal, B.; Gupta, M.; Singh, M.; Mahboob, T.; Teotia, A.; Thota, P.; Bhattacharya, J.; Goyal, R. Green synthesis of silver nanoparticles using Ocimum sanctum Linn. and its antibacterial activity against multidrug resistant Acinetobacter baumannii. PeerJ 2023, 11, e15590. [Google Scholar] [CrossRef]
Figure 1. Color change in the reaction mixture of silver nanoparticles synthesized from Ricinus communis (RcSNPs, (a)) and Aloe barbadensis (AbSNPs, (b)) over time (1 h, 4 h, and 24 h).
Figure 1. Color change in the reaction mixture of silver nanoparticles synthesized from Ricinus communis (RcSNPs, (a)) and Aloe barbadensis (AbSNPs, (b)) over time (1 h, 4 h, and 24 h).
Bioengineering 12 01273 g001
Figure 2. UV–visible spectra of silver nanoparticles synthesized from Ricinus communis (RcSNPs) and Aloe barbadensis (AbSNPs) showing the effects of varying parameters: AgNO3 concentration (a,f), leaf extract concentration (b,g), reaction time (c,h), pH (d,i), and temperature (e,j). Data represent mean ± SD (n = 3).
Figure 2. UV–visible spectra of silver nanoparticles synthesized from Ricinus communis (RcSNPs) and Aloe barbadensis (AbSNPs) showing the effects of varying parameters: AgNO3 concentration (a,f), leaf extract concentration (b,g), reaction time (c,h), pH (d,i), and temperature (e,j). Data represent mean ± SD (n = 3).
Bioengineering 12 01273 g002aBioengineering 12 01273 g002b
Figure 3. Stability studies of silver nanoparticles synthesized from Ricinus communis (RcSNPs (a)) and Aloe barbadensis (AbSNPs (b)) over a one-month period. Data represent mean ± SD (n = 3).
Figure 3. Stability studies of silver nanoparticles synthesized from Ricinus communis (RcSNPs (a)) and Aloe barbadensis (AbSNPs (b)) over a one-month period. Data represent mean ± SD (n = 3).
Bioengineering 12 01273 g003
Figure 4. Dynamic light scattering (DLS) size-distribution histograms of silver nanoparticles synthesized from Ricinus communis (RcSNPs (a)) and Aloe barbadensis (AbSNPs (b)). The X-axis represents particle diameter (nm) and the Y-axis represents number distribution (%). The average hydrodynamic diameters were 93.7 ± 1.6 nm (RcSNPs) and 126.1 ± 0.1 nm (AbSNPs), with polydispersity indices (PDI) of 0.29 ± 0.002 and 0.13 ± 0.02, respectively (n = 3).
Figure 4. Dynamic light scattering (DLS) size-distribution histograms of silver nanoparticles synthesized from Ricinus communis (RcSNPs (a)) and Aloe barbadensis (AbSNPs (b)). The X-axis represents particle diameter (nm) and the Y-axis represents number distribution (%). The average hydrodynamic diameters were 93.7 ± 1.6 nm (RcSNPs) and 126.1 ± 0.1 nm (AbSNPs), with polydispersity indices (PDI) of 0.29 ± 0.002 and 0.13 ± 0.02, respectively (n = 3).
Bioengineering 12 01273 g004
Figure 5. Fourier transform infrared (FTIR) spectra of Ricinus communis and Aloe barbadensis leaf extracts, and their corresponding silver nanoparticles (RcSNPs and AbSNPs).
Figure 5. Fourier transform infrared (FTIR) spectra of Ricinus communis and Aloe barbadensis leaf extracts, and their corresponding silver nanoparticles (RcSNPs and AbSNPs).
Bioengineering 12 01273 g005
Figure 6. Scanning electron microscopy (SEM) images of silver nanoparticles synthesized from Ricinus communis (RcSNPs (a)) and Aloe barbadensis (AbSNPs (b)) at different magnifications (×600, ×3000, ×6000, ×10,000).
Figure 6. Scanning electron microscopy (SEM) images of silver nanoparticles synthesized from Ricinus communis (RcSNPs (a)) and Aloe barbadensis (AbSNPs (b)) at different magnifications (×600, ×3000, ×6000, ×10,000).
Bioengineering 12 01273 g006
Figure 7. Transmission electron microscopy (TEM) images of silver nanoparticles synthesized from Ricinus communis (RcSNPs (a,b)) and Aloe barbadensis (AbSNPs (c,d)) at different magnifications.
Figure 7. Transmission electron microscopy (TEM) images of silver nanoparticles synthesized from Ricinus communis (RcSNPs (a,b)) and Aloe barbadensis (AbSNPs (c,d)) at different magnifications.
Bioengineering 12 01273 g007
Figure 8. Thermogravimetric (TGA) and differential scanning calorimetry (DSC) thermograms of silver nanoparticles synthesized from Ricinus communis (RcSNPs, (a) and Aloe barbadensis (AbSNPs, (b). Green line represents TGA while blue line represents DSC.
Figure 8. Thermogravimetric (TGA) and differential scanning calorimetry (DSC) thermograms of silver nanoparticles synthesized from Ricinus communis (RcSNPs, (a) and Aloe barbadensis (AbSNPs, (b). Green line represents TGA while blue line represents DSC.
Bioengineering 12 01273 g008
Figure 9. Antibacterial activity of silver nanoparticles synthesized from Ricinus communis (RcSNPs) and Aloe barbadensis (AbSNPs) against pathogenic bacterial strains. Treatments: leaf extract (1), 50 µL SNPs (2), 100 µL SNPs (3). Data represent mean ± SD (n = 3).
Figure 9. Antibacterial activity of silver nanoparticles synthesized from Ricinus communis (RcSNPs) and Aloe barbadensis (AbSNPs) against pathogenic bacterial strains. Treatments: leaf extract (1), 50 µL SNPs (2), 100 µL SNPs (3). Data represent mean ± SD (n = 3).
Bioengineering 12 01273 g009
Figure 10. Antifungal activity of silver nanoparticles synthesized from Ricinus communis (RcSNPs) and Aloe barbadensis (AbSNPs) against pathogenic fungal strains. Treatments: leaf extract (1), fluconazole (2), 50 µL SNPs (3), 100 µL SNPs (4). Data represent mean ± SD (n = 3).
Figure 10. Antifungal activity of silver nanoparticles synthesized from Ricinus communis (RcSNPs) and Aloe barbadensis (AbSNPs) against pathogenic fungal strains. Treatments: leaf extract (1), fluconazole (2), 50 µL SNPs (3), 100 µL SNPs (4). Data represent mean ± SD (n = 3).
Bioengineering 12 01273 g010
Figure 11. In vitro anti-inflammatory activity of Ricinus communis (Rc) and Aloe barbadensis (Ab) extracts and their corresponding SNPs. (a) Heat-induced protein denaturation; (b) Protease inhibition. Values with different superscripts (a–j) within the same variable are significantly different (p < 0.05).
Figure 11. In vitro anti-inflammatory activity of Ricinus communis (Rc) and Aloe barbadensis (Ab) extracts and their corresponding SNPs. (a) Heat-induced protein denaturation; (b) Protease inhibition. Values with different superscripts (a–j) within the same variable are significantly different (p < 0.05).
Bioengineering 12 01273 g011
Figure 12. In vitro antioxidant activity of Ricinus communis (Rc) and Aloe barbadensis (Ab) extracts and their corresponding SNPs. (a) DPPH assay; (b) H2O2 scavenging assay; (c) NO scavenging assay; (d) Reducing power assay. Values with different superscripts (a–j) within the same variable are significantly different (p < 0.05).
Figure 12. In vitro antioxidant activity of Ricinus communis (Rc) and Aloe barbadensis (Ab) extracts and their corresponding SNPs. (a) DPPH assay; (b) H2O2 scavenging assay; (c) NO scavenging assay; (d) Reducing power assay. Values with different superscripts (a–j) within the same variable are significantly different (p < 0.05).
Bioengineering 12 01273 g012
Figure 13. Cytotoxic activity of Ricinus communis (Rc) and Aloe barbadensis (Ab) extracts and their corresponding SNPs against B16F10 and HepG2 cell lines. Panels show (a) Rc and Ab treatments on B16F10 cells after 24 h, (b) B16F10 cells after 48 h, (c) HepG2 cells after 24 h, and (d) HepG2 cells after 48 h. Values with different superscripts (a–j) within the same variable are significantly different (p < 0.05).
Figure 13. Cytotoxic activity of Ricinus communis (Rc) and Aloe barbadensis (Ab) extracts and their corresponding SNPs against B16F10 and HepG2 cell lines. Panels show (a) Rc and Ab treatments on B16F10 cells after 24 h, (b) B16F10 cells after 48 h, (c) HepG2 cells after 24 h, and (d) HepG2 cells after 48 h. Values with different superscripts (a–j) within the same variable are significantly different (p < 0.05).
Bioengineering 12 01273 g013aBioengineering 12 01273 g013b
Table 1. Phytochemical analysis of Ricinus communis and Aloe barbadensis leaf extracts.
Table 1. Phytochemical analysis of Ricinus communis and Aloe barbadensis leaf extracts.
PhytochemicalsAqueous ExtractMethanolic Extract
Ricinus communisAloe barbadensisRicinus communisAloe barbadensis
Alkaloids+++
Steroids++
Flavonoids++++
Terpenoids+++
Glycosides+
Phenols+++
Tannins+++
Saponins+++
Reducing sugars++++
Table 2. Zones of inhibition (mm) of RcSNPs, AbSNPs, standard antimicrobial drugs, and leaf extracts of Ricinus communis (Rc) and Aloe barbadensis (Ab) against pathogenic bacterial and fungal strains.
Table 2. Zones of inhibition (mm) of RcSNPs, AbSNPs, standard antimicrobial drugs, and leaf extracts of Ricinus communis (Rc) and Aloe barbadensis (Ab) against pathogenic bacterial and fungal strains.
MicroorganismsRcSNPsAbSNPsStandardControl
50 μL100 μL50 μL100 μLGentamycinFluconazoleRc ExtractAb Extract
BacteriaS. aureus18.32 ± 0.320.2 ± 0.217.2 ± 0.120.3 ± 0.125.2 ± 0.3-12.3 ± 0.410.4 ± 0.5
S. typhi18.1 ± 0.719.3 ± 0.514.5 ± 0.118.6 ± 0.127.8 ± 0.1-14.6 ± 0.48.6 ± 0.6
E. coli17.6 ± 0.918.4 ± 0.716.1 ± 0.0517.3 ± 0.0523.4 ± 0.1-11.2 ± 0.711.7 ± 0.3
FungiC. albicans20.1 ± 0.722.3 ± 0.817.25 ± 0.220.5 ± 0.1-22.8 ± 0.710.7 ± 0.913.6 ± 0.2
A. niger19.4 ± 0.721.4 ± 0.722.5 ± 0.324.3 ± 0.4-23.21 ± 0.79.8 ± 0.217.1 ± 0.8
50 µg/mL, 50–100 µL per well. Experiments were performed in triplicates and the values are expressed as mean+ SD.
Table 3. Tyrosinase inhibitory activity (%) of silver nanoparticles synthesized from Ricinus communis (RcSNPs) and Aloe barbadensis (AbSNPs).
Table 3. Tyrosinase inhibitory activity (%) of silver nanoparticles synthesized from Ricinus communis (RcSNPs) and Aloe barbadensis (AbSNPs).
Testing MaterialTyrosinase Inhibition (%)
RcSNPs97.98 ± 1.5
AbSNPs98.43 ± 1.5
Ascorbic acid99.23± 0.1
Values represent mean ± SD (n = 3).
Table 4. Comparative summary of phytogenic silver nanoparticles reported in recent literature versus the present study.
Table 4. Comparative summary of phytogenic silver nanoparticles reported in recent literature versus the present study.
Source (Plant Extract)Particle Size (nm)ShapeSynthesis FeaturesAntimicrobial Activity (Zone of Inhibition, mm)Cytotoxic/Other BioactivityReference
Azadirachta indica (Neem)25–45SphericalModerate stability; slow reduction (4 h)E. coli—17 ± 0.5 mmModerate anticancer (IC50 > 100 µg/mL)[167]
Camellia sinensis (Green tea)30–60Quasi-sphericalGood dispersion; requires heatingS. aureus—18 ± 0.7 mmLow cytotoxicity (IC50 > 150 µg/mL)[168]
Moringa oleifera20–40SphericalHigh yield; moderate stabilityE. coli—19 ± 0.4 mmLimited antiproliferative effect[169]
Ocimum sanctum (Tulsi)18–30IrregularRequires alkaline pH > 9A. baumannii—15 ± 0.8 mmModerate anticancer (IC50 > 125 µg/mL[170]
Ricinus communis (present study)16–22Uniform sphericalRapid synthesis (≤30 min); high stability (>3 mo)S. aureus—20.2 ± 0.9 mm; C. albicans—22.3 ± 0.8 mmStrong cytotoxicity (B16F10 IC50 ≈ 52 µg/mL)This study
Aloe barbadensis (present study)18–24SphericalGood colloidal stability; reproducible synthesisA. niger—21.8 ± 0.4 mm; E. coli—17.3 ± 0.05 mm Significant cytotoxicity (HepG2 IC50 ≈ 60 µg/mL)This study
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Ahsan, A.; Gao, G.F.; Tian, W.-X. Phytogenic Silver Nanoparticles Derived from Ricinus communis and Aloe barbadensis: Synthesis, Characterization, and Evaluation of Biomedical Potential. Bioengineering 2025, 12, 1273. https://doi.org/10.3390/bioengineering12111273

AMA Style

Ahsan A, Gao GF, Tian W-X. Phytogenic Silver Nanoparticles Derived from Ricinus communis and Aloe barbadensis: Synthesis, Characterization, and Evaluation of Biomedical Potential. Bioengineering. 2025; 12(11):1273. https://doi.org/10.3390/bioengineering12111273

Chicago/Turabian Style

Ahsan, Anam, George F. Gao, and Wen-Xia Tian. 2025. "Phytogenic Silver Nanoparticles Derived from Ricinus communis and Aloe barbadensis: Synthesis, Characterization, and Evaluation of Biomedical Potential" Bioengineering 12, no. 11: 1273. https://doi.org/10.3390/bioengineering12111273

APA Style

Ahsan, A., Gao, G. F., & Tian, W.-X. (2025). Phytogenic Silver Nanoparticles Derived from Ricinus communis and Aloe barbadensis: Synthesis, Characterization, and Evaluation of Biomedical Potential. Bioengineering, 12(11), 1273. https://doi.org/10.3390/bioengineering12111273

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop