Home Pharmaceutical properties for green fabricated ZnO and Ag nanoparticle-mediated Borago officinalis: In silico predications study
Article Open Access

Pharmaceutical properties for green fabricated ZnO and Ag nanoparticle-mediated Borago officinalis: In silico predications study

  • Alyaa A. Alkhafaji , Hind M. Ahmed , Yasir B. Fadhil , Mina M. Allos , Majid S. Jabir EMAIL logo , Duha S. Ahmed , Ali Abdullah Issa , Kin Weng Kong , Beng Fye Lau , Suresh Ghotekar and Ayman A. Swelum EMAIL logo
Published/Copyright: March 26, 2025
Become an author with De Gruyter Brill

Abstract

In this study, Borago officinalis flower extract was performed for green synthesis of silver nanoparticles (AgNPs) and Zinc Oxide (ZnONPs). The plant extract played essential roles as a stabilising and reducing agent. The aim of the current study was to evaluate the pharmaceutical applications for green fabricated ZnO NPs and Ag NPs such as antibacterial, antibiofilm properties, antioxidant, and anticancer activity. The resulting nanoparticles were characterised using ultraviolet (UV)–visible, Fourier-transform infrared spectroscopy, X-ray Diffraction, Field emission scanning electron microscopy, and Energy-dispersive X-ray spectroscopy analyses. The findings revealed the formation of spherical Ag NPs (33–41 nm) and ZnO NPs (59–80 nm) with the cubic structure of Ag NPs and the hexagonal structure of ZnO NPs. The antibacterial activity was assessed against Staphylococcus aureus, Streptococcus mutans, Klebsiella pneumoniae, and Escherichia coli using the agar well diffusion method. The antibiofilm potential was evaluated using a crystal violet assay, revealing that ZnO NPs demonstrated slightly superior antibiofilm activity compared to Ag NPs, particularly against K. pneumoniae. The enhanced antibacterial and antibiofilm activities are attributed to the nanoparticles, ability to generate reactive oxygen species and their improved physical interactions with bacterial cells. The cytotoxic effects of prepared NPs against colon cancer cells HT-29 were investigated. The cytotoxicity assay confirmed that the prepared NPs were selectively toxic in cancer cells. Our results showed that the prepared NPs had antioxidant activity when using the DPPH assay. The results improved the potential properties of green synthesised Ag NPs and ZnO NPs as effective antimicrobial agents and prevent biofilm-formation-associated infections. Taken together, the results highlighted the viability of utilising B. officinalis in the eco-friendly biosynthesis of nanoparticles for the application of biomedical.

1 Introduction

The nanotechnology field has been expanded rapidly, particularly nanomaterials, due to their unique features, which enable versatility in many industries and biomedical science applications [1]. Among the nanomaterials, silver nanoparticles (Ag NPs), characterised as nontoxic ions, have been mainly investigated as antibacterial, antifungal, and antiviral agents that can effectively suppress many bacteria and fungi [2]. Similarly, zinc oxide nanoparticles (ZnO NPs) used in biological applications have gained much interest recently due to their unique anticancer, anti-inflammatory, antioxidant, and antibacterial features [3]. The synthesis of Ag and ZnO NPs is conducted using several conventional methods, such as physical and chemical techniques; however, there are a lot of concerns regarding the toxicity of these methods, which have adverse impacts on the environment and human health [4]. Therefore, green strategies for metal nanoparticle synthesis utilising microorganisms and plants have gained considerable efforts with an emphasis on reducing toxic substances [5]. The plant-based techniques for green synthesis of nanoparticles have gained widespread use due to their vast potential applications across various sectors, including targeted and safer drug delivery, offering a promising alternative to conventional cancer treatments and environmental applications [6]. CuNPs significantly inhibited the proliferation of the lung cancer cell lines NCI-H661, NCI-H1975, NCI-H1573, and NCI-H1563. Moreover, CuNPs did not exhibit cytotoxicity against human umbilical vein endothelial cells at doses as high as 1,000 μg·mL−1 [7]. Based on the results of a recently published study [8], bio-synthetised Ag nanoparticles plant extract played a critical role as an antioxidant, anticancerous agent, and in controlling the levels of blood sugar in diabetic rats, while no adverse effects were observed or reported on the liver and pancreas tissues. Phytochemical compounds of plant extracts such as flavonoids and other polyphenols can act as reducing and capping agents in synthesising stable metal nanoparticles such as silver, zinc, gold, and copper nanoparticles. They might enhance their biological activities [9]. Borago officinalis, also known as Borage or starflower, is a medicinal plant that grows annually. Borage is cultivated in Europe, North Africa, and the Mediterranean region. Its leaves, flowers, and seeds have been used in diverse domains such as pharmaceutical and industrial, as well as traditional therapy for multiple conditions, including gastrointestinal disorders (colic, cramps, and diarrhoea), respiratory issues (asthma and bronchitis), and cardiovascular problems (as a cardiotonic, antihypertensive, and blood purifier) [10]. Therefore, B. officinalis leaf extracts have been explored for their potential in the eco-friendly synthesis of silver nanoparticles and examined the antibacterial effect and antibiofilm inhibition for various pathogenic bacteria as well as anticancer capability [11,12], while Jasem and Abas [13] used B. officinalis flower extract for Ag NPs green synthesis, and they have been tested for their larvicidal activity. However, much less is known about the potential of B. officinalis in the green synthesis of ZnO NPs. The flowers of B. officinalis are abundant in bioactive compounds that exhibit notable antioxidant, antibacterial, anti-inflammatory, and possible anticancer effects. These attributes suggest the plant’s potential for incorporation into health supplements, treatments, and medical and industrial applications [14]. Therefore, in this study, B. officinalis flowers were used to acquire the crude extract and applied to prepare and stabilise Ag NPs and ZnO NPs. Then, the antibacterial activities of Ag NPs and ZnO NPs against two Gram-positive bacteria strains, Staphylococcus aureus and Streptococcus mutans, and two Gram-negative bacteria, Klebsiella pneumoniae and Escherichia coli, were evaluated, as well as their effects on biofilm formation of bacterial strains. In addition, they investigated their role as antioxidant and anticancer agents against colon cancer cell lines (HT-29). The green synthesised nanoparticles showed promising results in biomedical applications. This study aims to develop a green synthesis approach for fabricating zinc oxide (ZnO) and silver (Ag) nanoparticles, utilising environmentally friendly methods to enhance their biocompatibility and minimise toxic effects. The synthesised nanoparticles and their antioxidant and anticancer activities will be evaluated for their antimicrobial potential, particularly against multidrug-resistant bacteria. Additionally, in silico molecular docking studies will be conducted to predict their interactions with bacterial and cancer-related target proteins, providing mechanistic insights into their biological efficacy. The study seeks to bridge the gap between computational predictions and experimental validation, offering a sustainable and effective alternative for biomedical applications.

2 Materials and methods

2.1 Materials

Silver nitrate (AgNO3, 99.9%) was purchased from Daejung, Korea, and Zinc nitrate hexahydrate (Zn (NO3)2·6H2O, 99.0%) was purchased from Himedia, India. Ethanol was purchased from Duksan, Korea. Mueller Hinton Agar was purchased from Accumix, Spain. Nutrient broth was purchased from Oxoid, United Kingdom. The dried Borago officinalis flowers were purchased from a local market in Baghdad. Deionised water was used throughout the experiment.

2.2 Preparation of Borago officinalis flower extract

The purchased dried Borago officinalis flowers were extracted with 80% aqueous ethanol at 5:1,000 (w/v) for 3 h. The extract was filtered, and the excess ethanol was dried at 70°C using the oven. The obtained extract was stored in the fridge for further investigation. The extraction process was based on the studies by Rahmah et al. and Aziz et al. [15,16].

2.3 Preparation of Ag NPs

The green synthesis approach of Ag NPs was attained by the method of [17]. A 6.5 g of AgNO3 was dissolved in 250 mL of deionised water using a magnetic stirrer (800 rpm) at room temperature (provide temperature, 25°C). Then, 100 mL of plant extract (B. officinalis) was the precursor solution. After 24 h, the colour of the mixing solvent alters to dark brown. The suspension was separated using a centrifuge, and the residue was washed with water and ethanol five times. The residue was dried using the oven at 40°C for 12 h, and Ag NPs were then obtained as the final product. Figure 1 illustrates the steps of Ag NPs preparation.

2.4 Preparation of ZnONPs

The green synthesis approach of Zn NPs was achieved by dissolving 7 g of zinc nitrate hexahydrate (Zn(NO3)2·6H2O) in 200 mL of deionised water at 50°C for 120 min using a magnetic stirrer. Then, 150 mL of B. officinalis plant extract (5.5 mg·mL−1) was mixed with the above solution and stirred for 60 min. Then, 50 mL of 1 M sodium hydroxide (NaOH) was added dropwise to the mixture solution. Then, the precipitation was washed by centrifuge with deionised water and ethanol several times. The resulting rainfall was dried at 100°C for 3 h to form powder, and impurities were removed [18,19]. Then, the calcination process was carried out at 550°C for 3 h, and ZnO NPs were formed. Figure 2 illustrates the steps of ZnO NP preparation.

Figure 1 
                  Green synthesis of Ag NPs.
Figure 1

Green synthesis of Ag NPs.

2.5 Characterisation

The optical properties of Ag NPs and ZnO NPs synthesised using the green method are determined using UV–visible spectroscopy (Shimadzu, UV-1800 spectrophotometer) at room temperature at 200–800 nm. The Fourier-transform infrared spectroscopy (FTIR) spectrum (Frontier FTIR-88091) was employed to analyse the boundary construction of resulted samples with an absorption range of 500–4,000 cm−1. The functional gropes contained in each nanoparticle were identified via FTIR analysis. Field emission scanning electron microscopy (FESEM, TESCAN, MIRA3) was employed to capture images of Ag NP and ZnO NP samples to study their morphology. Energy-dispersive X-ray spectroscopy (EDS) also provided a straightforward and effective method for analysing the samples’ elemental composition. The structural analysis of the Ag NP and ZnO NP samples was performed using an X-ray Diffraction (XRD) diffractometer (XRD diffraction 6000, Shimadzu) with CuKα radiation (λ = 1.542 Å), operating at a current of 30 mA and a voltage of 40 kV. Data were collected in the 2θ range of 10–80°. The average particle size (D) was determined using the Scherrer equation, D = /(β cosθ), where k is a constant (0.89), β represents the full width at half maximum, and θ corresponds to the Bragg angle.

2.6 Preparation of Mueller Hinton agar

Muller–Hinton (M–H) was prepared by adding 20 mL of the powdered agar into 1 L of distilled water. The mixture was heated under shaking conditions. M–H agar was sterilised at 121°C for 15 min. Then, it was cooled to 50°C and poured into a Petri dish. After solidifying the agar, the Petri dish was flipped and stored at 4°C.

2.7 Antibacterial activity

The antibacterial activity of the Ag NPs and ZnO NPs was performed using an agar well diffusion assay [20,21] against two Gram-positive bacteria, including Staphylococcus aureus and Streptococcus mutants, and two Gram-negative bacteria, including Klebsiella pneumonia and Escherichia coli. The bacterial samples were inoculated in 6-mm-diameter wells on an M–H agar petri dish. The final concentration (12.5, 25, 50, and 100 μg·mL−1) of Ag NPs and ZnO NPs are added into the wells. The plates were incubated overnight at 37°C before measuring the inhibition zone diameter.

2.8 Biofilm inhibition

The antibiofilm activity of the Ag NPs and ZnO NPs against the Gram-positive and Gram-negative bacteria was conducted using crystal violet (CV) staining obtained by Al Rugaie et al. [22]. The bacterial strains were cultured in 96-well plates at a concentration of 1 × 106, treated with Ag NPs and ZnO NPs, and then inoculated for 24 h. After incubation, the samples were washed with Phosphate-buffered saline (PBS). Adhered bacteria were stained with 0.1% CV and rinsed twice with distilled water. To assess biofilm formation, 0.2 mL of 95% ethanol was added to the CV-stained wells, followed by 2 h of incubation with shaking. The optical density was measured at 959 nm.

2.9 Bacterial reactive oxygen species (ROS) formation using DCFH-DA assay

Bacterial strains were treated with prepared nanoparticles. Then, 25 µM DCFH-DA in 1× PBS for 45 min was added. Samples were washed 2 times with 1× PBS buffer. Using the ZEISS Axioscope microscope with the Filter Set 38 HE (Oberkochen, Germany), the fluorescence was observed at a wavelength of 535 nm following the excitation of the sample at 485 nm.

2.10 DPPH assay

NPs were tested as antioxidants using (DPPH) method; 12 mg of DPPH was dissolved in 50 mL of methanol to obtain the stock solution. In the test tube, 3 mL of DPPH was combined with 100 μL of prepared NPs. Besides, 3 mL of DPPH solution in 100 μL of methanol can be represented as standard. After that, the tubes were kept in a dark place for 30 min. The absorbance was therefore determined at λ = 517 nm. The formula that used to calculate the percentage of antioxidants (%) = [(A cA s) ÷ A c] × 100% where A c is the control absorbance and A s is the testing sample absorbance.

2.11 MTT assay

This technique was used to investigate the cytotoxic effect of prepared nanoparticles against colon cancer cell line (HT-29). HT-29 cells were seeded in a 96-well plate at density 1 × 104; after 24-h incubation; cells were treated with different concentrations of NPs (0, 12.5, 25, 50, and 100 μg·mL−1). MTT solution at a concentration (2 mg·mL−1) was added and then incubated for 3 h at 37°C; 100 µL of (DMSO) was added, followed by shaking for 15 min at room temperature, and the absorbance at 492 nm was measured using a microplate reader.

2.12 In silico analyses

The molecular structure was designed by materials studio software, and a sphere of ZnNPs and AgNPs with a radius of 3 Å was generated. Four targeted proteins were selected to investigate the bioactive properties of the ZnO NP and AgNP compounds. The target proteins, Colon Cancer Antigen 10 (ID: 2HQ6), MltA from E. coli (ID: 2GAE), S. mutans (ID: 3Q6A), and Peroxiredoxin from Aeropyrum pernix K1 (ID: 6KRR) were obtained from the Protein Data Bank www.rcsb.org. Protein preparation, including removing water molecules and heteroatoms, was performed using Biovia Discovery Studio software. Energy minimisation and checks for missing regions in the selected receptors were carried out using SWISS-PDBVIEWER (version 4.1.0, SPDBV). The molecular docking process was executed using AutoDock 4.2 software. It started with the addition of polar hydrogens to the proteins, followed by the application of Kollman charges and the assignment of Gasteiger charges. The grid box was configured to encompass the entire protein structure. Docking parameters were optimised using genetic algorithms, generating 50 conformations/poses. The best conformation was selected based on the lowest binding free energy, which reflected ligand–receptor non-covalent interactions. Hydrophobic interactions between the docked drug-like molecules and the target proteins were identified using LigPlus (LigPlot + v.2.2.8), while Biovia Discovery Studio was used to visualise the best-posed conformations.

2.13 Statistical analysis

The data were analysed using GraphPad Prism software (version 6). The mean ± standard deviation (SD) is used to display the data. All experiment was carried out at least three times. Tukey’s test was used for post-hoc comparisons between groups after an independent-sample t-test or a one-way analysis of variance was suitably used for statistical analysis. Statistical significance was determined at p < 0.05 [23,24].

3 Results and discussions

3.1 Structural, optical, and morphological studies

Figure 3a reveals the UV–visible absorption spectra of Ag NPs synthesised using the green method. The UV–visible spectrum of Ag NPs was studied in a range of 200–800 nm. Figure 1a displays a broad absorption band around (406.6–538.6 nm) with a strong red shift edge at 437 nm. The broad peak of the surfers’ Plasmon resonance band around 437 nm has revealed the formation of Ag NPs. The SPR band related to Ag NPs was formed as a result of the presence of the reducing agent of compounds existing in the B. officinalis flower extract, which was employed for the incorporation of Ag NPs and the energy band gap by using the Tauc plot of the (αhυ)2 versus photon energy () curve with n = 1\2 for direct transition is about 3.76 eV, as shown in the inset of Figure 4(a). Figure 4(b) reveals the UV–visible absorption spectra of ZnO NPs synthesised by the green method. The B. officinalis flower extract–mediated ZnO NPs were characterised by a UV–visible spectrum (200–800 nm) range where ZnO NPs show a strong absorption peak at 376.7 nm, as shown in Figure 1b. Besides, the UV–visible spectroscopic reveals a sharp absorption peak and improves the stabilising agent as ZnO NPs are synthesised by the aqueous leaf of extract B. officinalis. In addition, the band gap of green fabricated Zn ONPs was calculated using the Tauc plot as 2.27 eV, as shown in the inset in Figure 4(b).

Figure 2 
                  The green synthesis of ZnONPs.
Figure 2

The green synthesis of ZnONPs.

Figure 3 
                  The UV–visible spectrum of (a) Ag NPs and (b) synthesised by green method.
Figure 3

The UV–visible spectrum of (a) Ag NPs and (b) synthesised by green method.

The XRD diffraction spectra of Ag NPs and ZnO NPs were analysed as illustrated in Figure 4(a) and (b). In Figure 4(a), the XRD patterns spectrum of Ag NPs display four prominent peaks at 2Ɵ = 38.11°, 44.30°, 64.44° and 77.40°, and these peaks‘ reflections correspond to planes of (111), (200), (220), and (311), respectively of cubic structure and matched with (JCPDS card-96-900-8460) of Ag nanoparticles [24]. These diffraction peaks indicate the preparation of silver nanoparticles using the green method by extracting plant compounds. In addition, a few peaks at 2θ = 27.8° and 32.33° were related to the green method using plant extract, revealing its role as a reducing agent of Ag ions. The average particle crystallite of Ag NPs was obtained using the Scherrer formula and found to be 18.25 nm from the peak intensity of Ag NPs (111). In state of ZnONPs prepared by green procedure using B. officinalis plant extract, the XRD patterns displays peaks at 2θ = 31.77°, 34.42°, 36.26°, 47.54°, 56.61°, 62.86°, 66.39°,67.96°, and 69.10° were related to planes (100), (002), (101), (102), (110), (103), (200), (112), and (201), respectively, as shown in Figure 4(b). The XRD data are related to the hexagonal structure of synthesised ZnO NPs by the green method [23,24]. The obtained data are matched with (JCPDS card-96-230-0451), and it can be revealed that the resulting crystallite ZnO NPs show high purity as shown in the XRD spectrum [25]. Moreover, the average particle crystallite of ZnO were obtained using the Scherrer formula and found to be 27.4, 27.6, and 47.6 nm from the peaks intensity of ZnO (100), (002), and (101), respectively. Moreover, using optimal concentrations of B. officinalis extract improves the crystallinity of AgNPs and ZnO NPs, as evidenced by sharper XRD peaks related to nanoparticle nucleation and growth rate. The Scherrer formula was employed to estimate crystallite size and confirm the structural properties relevant to the applications. Additionally, FESEM and Image J were used to validate particle morphology and size distribution, ensuring consistency in the synthesised nanoparticles.

The FTIR spectrum of biosynthesised AgNPs and ZnONPs in the range (500–4,000 cm−1) are shown in Figure 6(a) and (b) at room temperature. The FTIR spectra of Ag NPs in Figure 6(a) showed the major absorption bands at (3,425–3,351 cm−1), (2,981–2,844 cm−1), 1,697, 1,568, 1,379, 1,125 and 732 cm−1. The bands at (3,425–3,351 cm−1), (2,981–2,844 cm−1), and 1,379 cm−1 were associated with stretching vibrating of hydroxyl groups (O–H) and alkane (C–H). Besides, the peaks at 1,697, 1,568, and 732 cm−1 are according to the bending vibrating of alkene (C═O) groups. The band at 1,125 cm−1 was related to the amide band containing carbonyl groups (C═O). The peak at 877 cm−1 represented aromatic groups (C═H) for alkene. Moreover, these peaks are due to reduced sugars, flavonoids, saccharides, and proteins in the extract plant compounds. These functional groups form as reducing agents, providing stability and preventing agglomeration of Ag NPs in ZnONPs. Phytochemicals are essential in NP synthesis as they are crucial to stabilising the NPs by forming chelate complexes with metal ions. The primary mechanism explaining the formation of NPs summarised that the extract phytochemicals reduced and oxidised the ions. These compounds act as reducing agents, stabilisers, and capping agents, replacing the requirement for additional chemical stabilisers. Once the reaction is complete, the resulting product undergoes annealing to obtain NPs [26]. The FTIR spectrum of biosynthesised ZnONPs in the range of (500–4,000) cm−1 is shown in Figure 6(b) at room temperature. The FTIR shows the prominent peak positions related to synthesised ZnO NPs using plant extract at 3,377 and 3,695 cm−1, indicating broad peaks of O–H stretching vibrating. The beaks at 2,978 and 2,886 cm−1 corresponded to the CH3 stretching vibration. The peak at 1,527 cm−1 indicates C═C bond stretching vibrating. The peak at 2,162 cm−1 is related to triple bond stretching, and the peak at 1,384 cm−1 indicates C–O–H bending vibration. The peal at 1,255 cm−1 relates to C–O stretching in aromatic alcohols, while the peak at 1,079 cm−1 indicates the C–O an aliphatic C–N bond stretching. Meanwhile, the peak at 586 cm−1 corresponds to Zn-O stretching and deformation vibration (Figure 5).

Figure 4 
                  The XRD diffraction of (a) AgNPs and (b) ZnONPs synthesised by a green method.
Figure 4

The XRD diffraction of (a) AgNPs and (b) ZnONPs synthesised by a green method.

The resulting peaks reveal the formation of ZnO NPs using the extract of B. officinalis. The resulting peaks indicate the presence of functional groups like ketones, aldehydes, hydroxyl, alcohols, phenolic and carbocyclic acid with hydrogen bonds, which attributed to flavonoids, terpenoids, glycosides, and phenols and confirmed the green synthesised ZnO NPs by suing plant extracted as capping and stabilising agent reducing the aggregation of ZnONPs. The FESEM was used to study the surface morphology of Ag NPs and ZnO NPs, as shown in Figure 6(a)–(d). An expanded view in FESEM revealed that Ag NPs were relativity spherical. The average size of Ag NPs synthesised from the extract of B. officinalis flowers was measured to be in the range of (33–41 nm) as shown in Figure 6(a) and (b). in the case of ZnONPs synthesised from the extract of B. officinalis flowers, the morphology showed the spherical ZnO NPs with a size range of 50–80 nm, as shown in Figure 6(c) and (d). The resulting nanoparticles were in indirect contact, even within the aggregation, demonstrating the effect of a capping agent to stabilise the nanoparticles. Moreover, the FESEM images confirmed that using optimal extract concentration provides more nucleation sites, leading to the formation of smaller nanoparticles due to the availability of reducing and stabilising agents in the extract (flavonoids, polyphenols, and proteins). Moreover, the nanoparticles were developed by a quantitative element determined by using energy dispersive X-ray analysis (EDX) of Ag NPs and ZnO NPs, as shown in Figure 7(a) and (b), respectively. The EDX spectrum of Ag NPs displays the peaks of Ag, carbon, and oxygen elements, confirming Ag NPs formation with the highest peak intensity by using plant extracted as a capping agent without any additional peaks in EDX analysis, as shown in Figure 7(a). Furthermore, the particle size distribution obtained using image J analysis for Ag NPs and ZnONPs provides size measurements close to those observed in FESEM images, confirming its reliability. The results show that the measurements from Image J effectively match the resolution and details. In the state of ZnO NPs, the green synthesised ZnO NPs were further analysed through the EDX spectrum, as shown in Figure 7(b). The EDX analysis reveals the Zn, O, and carbon peaks, and Zn shows high peak intensity, confirming ZnO NP formation utilising plant extract as a capping agent. Besides, the appearance of Carbon elements in the EDX spectrum corresponds to biological molecules like amino, flavonoids, vitamins, and proteins in the plant extract. The average particle size for AgNPs is 44 and 69.5 nm for ZnO NPs, respectively, as shown in Figure 8(a) and (b). Image J provides an accurate representation of size variability by analysing many particles.

Figure 5 
                  The FTIR spectrum of biosynthesised (a) Ag NPs and (b) ZnO NPs.
Figure 5

The FTIR spectrum of biosynthesised (a) Ag NPs and (b) ZnO NPs.

Figure 6 
                  The FESEM images of (a) and (b) AgNPs and (c) and (d) ZnONPs.
Figure 6

The FESEM images of (a) and (b) AgNPs and (c) and (d) ZnONPs.

Figure 7 
                  EDS analysis of biosynthesised (a) Ag NPs and (b) ZnO NPs.
Figure 7

EDS analysis of biosynthesised (a) Ag NPs and (b) ZnO NPs.

3.2 Antibacterial activity of synthesised NPs

The initial screening of in vitro antibacterial activities of green synthesised Ag NPs and ZnO NPs was conducted using an agar well diffusion assay. Four various contents (12.5, 25, 50, and 100 µg·mL−1) of both Ag NPs and ZnO NPs were tested against two Gram-positive bacteria, S. aureus and S. mutans, and two Gram-negative bacteria, including K. pneumoniae and E. coli as shown in Table 1 and Figure 9. Ag NPs and ZnO NPs demonstrated vigorous antibacterial activities on S. mutans at the lowest amount of 12.5 µg·mL−1 with an inhibition zone of 22 and 24 mm, respectively; meanwhile, the largest inhibition zone was recorded at 100 µg·mL−1 of 28 and 30 mm, respectively, as shown in first-panel Figure 9. On the other hand, the second panel displays that the various concentrations of Ag NPs and ZnO NPs exhibited moderate and comparable antibacterial activities against S. aureus, with inhibition zones ranging from 12 to 22 mm. Notably, there was a direct correlation between the inhibition zone diameter and the concentrations of Ag NPs and ZnO NPs, as the inhibition zone diameter increased by 2 mm when doubling the concentrations of both Ag NPs and ZnO NPs. Regarding the results in the first panel, S. mutans was the bacteria most susceptible to Ag NPs and ZnO NPs compared to S. aureus due to its simple cell wall structure and lacking an outer membrane, which allows easy penetration by the nanoparticles and disturbs vital cellular functions [27]. Moreover, as shown in the third panel, Ag NPs and ZnO NPs produced less inhibition zone at 25 and 50 µg·mL−1 concentrations against K. pneumoniae than the control. Based on the results, 100 µg·mL−1 was the most effective against the Gram-negative bacteria K. pneumoniae with a maximum inhibition zone of 12 mm in. On the contrary, Ag NPs and ZnO NPs demonstrated strong and comparable antibacterial activity at the lowest concentration of 12.5 µg·mL−1 with inhibition zones of 25 and 24 mm, respectively. There was a gradual and slight increase in the inhibition zone, coinciding with increasing the green synthesised nanoparticle concentrations. Table 2 shows the antibacterial activity of positive control (Genatmicin and Ampicillin) against tested bacterial strains compared with prepared NPs. However, the possible explanation is Ag NPs’ ability to accumulate on the bits on the cell wall owing to the nanoscale size, causing cytoplasmic wall denaturation or changing the cell membrane structure, thereby resulting in cell lysis [28]. Another interpretation is the Ag NPs’ ability to release silver ions, which adhere to the cell wall and cytoplasmic membrane due to the electrostatic attraction and affinity to sulphur proteins, as a consequence, disrupt adenosine triphosphate and deactivation of respiratory enzymes by producing ROS, which considered the main factor in destroying DNA and eventual cell death [29]. The findings of this study documented neither Ag NPs nor ZnO NPs inhibited the growth of Gram-negative bacteria K. pneumoniae at a low amount of 12.5 µg·mL−1. As a comparison between the two Gram-negative bacteria, E. coli, as shown in the last panel, displayed more sensitivity to the green synthesised nanoparticles because of various factors, including the interactions between the cell wall and nanoparticles, the size of nanoparticles, as well as the E. coli cell wall lacking the thick layer of peptidoglycan, which make it vulnerable to the nanoparticles damaging impacts [30]. The mechanism of NPs’ antibacterial activity may be attributed to releasing toxic zinc and Ag ions and generating ROS, causing damage to the cell membrane and diffusing in the cytoplasm where they interact and affect various structures, leading to growth inhibition and cell death [31]. The Gram-positive bacteria consist of several peptidoglycan layers and teichoic acid, while the gram-negative bacteria consist of a thin layer of peptidoglycan surrounded by a secondary lipid membrane reinforced with transmembrane lipopolysaccharides and lipoproteins; the cell wall nature of the bacteria change due to the presence of NPs [32]. A semiconductor structure and surface defects characterise ZnO NPs; therefore, ZnO NPs are distinguished with magnificent photocatalytic features in the presence of light irradiation, leading to the production of ROS, including hydroxyl radicals (˙OH), singlet oxygen (1O2), superoxide anions (O2˙), and hydrogen peroxide (H2O2), these ROS considered the essential factors responsible for the antibacterial functions as they cause oxidative damage to DNA and the bacterial membrane. Furthermore, ZnO NPs released Zn ions (Zn2+), which are necessary for modulating cell functions and disturbing bacterial growth [33]. On the other hand, the antibacterial activities of Ag NPs are derived from different mechanisms, including releasing silver ions (Ag+), which stimulate the production of intracellular ROS, disturbing bacterial cell and DNA integrity. Ag ions can also bind with bacterial membrane proteins, thus inhibiting essential enzymes required for bacterial survival [34]. In addition, the direct interaction of Ag NPs with the bacteria’s cell membrane increases permeability and damages the structure, causing cell death. They also modify microbial signal transduction pathways by affecting phosphotyrosine profiles in proteins involved in growth regulation [35].

Table 1

Antibacterial activities of Ag NPs and ZnO NPs

Antibacterial inhibition zone (mm)
Bacteria samples Nanoparticles Control A B (12.5 µg·mL−1) C (25 µg·mL−1) D (50 µg·mL−1) E (100 µg·mL−1)
S. aureus Ag NPs 6 15 16 21 22
ZnO NPs 6 12 18 20 22
S. mutans Ag NPs 6 22 24 26 28
ZnO NPs 6 24 26 28 30
K. pneumoniae Ag NPs 6 6 7 8 11
ZnO NPs 6 6 7 8 12
E. coli Ag NPs 6 25 27 28 30
ZnO NPs 6 24 26 27 29
Figure 8 
                  The particle size distribution by using Image J of (a) Ag NPs and (b) ZnO NPs.
Figure 8

The particle size distribution by using Image J of (a) Ag NPs and (b) ZnO NPs.

Table 2

Antibacterial activities of positive control antibiotics

Antibacterial inhibition zone (mm)
Bacteria samples Nanoparticles (50 µg·mL−1) (100 µg·mL−1)
S. aureus Ampicillin 17.3 23.5
S. mutans Ampicillin 19.25 27.1
K. pneumoniae Gentamicin 15.36 12.6
E. coli Gentamicin 17.36 26.2

3.3 Antibiofilm activity of prepared NPs

Crystal violet assay assessed the effects of green Ag NPs and ZnO NPs in Gram-positive and Gram-negative bacteria biofilm formation. Figure 9 shows the susceptibility of tested bacteria biofilm to form AgNPs and ZnONPs using green method at different concentrations (12.5, 25, 50, and 100 μg·mL−1). Compared to the control, 12.5, 25, 50, and 100 µg·mL−1 concentrations inhibited biofilm formation. Ag NPs and ZnO NPs displayed comparable biofilm inhibition on S. aureus and S. mutans, ranging from 0.6 to 0.2 a.u., as shown in Figure 10.

Figure 9 
                  Antibacterial activities of Ag NPs and ZnO NPs against S. mutans (first panel), S. aureus, K. pneumoniae, and E. coli (last panel): (A) Control, (B) 12.5 µg·mL−1, (C) 25 µg·mL−1, (D) 50 µg·mL−1, and (E) 100 µg·mL−1.
Figure 9

Antibacterial activities of Ag NPs and ZnO NPs against S. mutans (first panel), S. aureus, K. pneumoniae, and E. coli (last panel): (A) Control, (B) 12.5 µg·mL−1, (C) 25 µg·mL−1, (D) 50 µg·mL−1, and (E) 100 µg·mL−1.

Similarly, Ag NPs and ZnO NPs exhibited comparable inhibition activity on K. pneumoniae biofilms. The lowest effect on the biofilm was displayed by Ag NPs of 0.5 a.u., at a concentration of 12.5 µg·mL−1 (well B); ZnO NPs of 0.1 a.u displayed the highest effect. at a concentration of 100 µg·mL−1 (well E), as reveal in Figure 10. Despite gradually doubling the Ag NPs and ZnO NPs concentrations, there was a slight difference in the biofilm inhibition of Gram-negative bacteria. In contrast, E. coli biofilm showed less sensitivity to both (Ag NPs and ZnO NPs). ZnO NPs exhibited the highest biofilm inhibition of 0.35 a.u. At 100 µg·mL−1 (E) concentration compared to its effect on K. pneumoniae, which exhibited biofilm inhibition of 0.1 a.u. At the same concentrations (Figure 11).

Figure 10 
                  Reduced biofilm formation by Ag NPs and ZnO NPs in S. aureus and S. mutans. (A) Control, (B) 12.5 µg·mL−1, (C) 25 µg·mL−1, (D) 50 µg·mL−1, and (E) 100 µg·mL−1.
Figure 10

Reduced biofilm formation by Ag NPs and ZnO NPs in S. aureus and S. mutans. (A) Control, (B) 12.5 µg·mL−1, (C) 25 µg·mL−1, (D) 50 µg·mL−1, and (E) 100 µg·mL−1.

The possible explanation suggested that nanoparticles can interact with and diffuse through biofilms by diffusing along the water channels, which serve as pathways for nutrient transport and can attach to the bacterial membranes, causing the reduction of exopolysaccharides secretion, which represents the primary component that forms the biofilm, eventually penetrate and accumulate inside the cells and leading to bacterial death [36]. Moreover, The enhanced antibiofilm activities of Ag NPs and ZnO NPs are primarily due to the rapid production of ROS and improved physical interactions between the nanoparticles and bacterial cells. These interactions cause significant damage to the bacterial cell membrane, hindering biofilm formation and reducing the likelihood of infection initiation and persistence. These mechanisms highlight the potential of green synthesised Ag NPs and ZnO NPs as powerful antibiofilm agents, particularly in preventing biofilm-related infections [37,38]. However, the overall results of this study can be summarised as follows: Despite the slight effect differences between Ag NPs and ZnO NPs on the bacterial biofilm, ZnO NPs are more potent as anti-biofilm.

3.4 Prepared nanoparticles induce bacterial ROS generation

In this study, the influences of NPs on bacterial ROS generation are investigated as indicated in Figure 12. The bactericidal effect of prepared NPs in bacteria is due to an increase in the cellular levels of ROS generation. The ROS generation assay findings verify that bacteria treated with prepared Ag and ZnO NPs produce ROS compared with control bacteria. This implies that the generation of ROS indirectly affects bacterial growth and contributes to bacterial death.

Figure 11 
                  Reduced biofilm formation by Ag NPs and ZnO NPs in K. pneumoniae and E. coli: (A) Control, (B) 12.5 µg·mL−1, (C) 25 µg·mL−1, (D) 50 µg·mL−1, and (E) 100 µg·mL−1.
Figure 11

Reduced biofilm formation by Ag NPs and ZnO NPs in K. pneumoniae and E. coli: (A) Control, (B) 12.5 µg·mL−1, (C) 25 µg·mL−1, (D) 50 µg·mL−1, and (E) 100 µg·mL−1.

3.5 Antioxidant activity of green fabricated NPs

Because of their inertness to chemicals and thermal stability, inorganic nanoparticles can be used to immobilise natural antioxidants. Furthermore, using naturally occurring antioxidants coupled with nanoparticles enhances these antioxidants’ chemical stability under physiological circumstances. This strategy enables the goods to be supplied in their whole molecular form at a broader range of concentrations. Additionally, it is notable that this strategy is consistent with the antioxidants’ constant release [39]. Figure 13 illustrates the DPPH radical scavenging activities of prepared NPs, which were positively correlated with increasing concentrations. Each concentration of NPs demonstrated significant DPPH radical scavenging ability. Certain oxides of nanoparticles possess inherent physicochemical characteristics that enable them to actively scavenge nitrogen and oxygen species and alleviate various conditions stemming from oxidative stress [40].

Figure 12 
                  Prepared NPs induce bacterial ROS generation.
Figure 12

Prepared NPs induce bacterial ROS generation.

3.6 Cytotoxic effect of NPs against colon cancer cell line

Green synthesis has become the most popular way to make nanoparticles. The body’s capacity to metabolise organic compounds, the relatively lower toxicity of plant-derived components, the compatibility of plant phytochemicals with therapeutic metal salts, and the environmental benefits of synthetic and organic moieties over strictly synthetic alternatives are the main factors contributing to the observed efficacy [34]. MTT assay was done to measure the antiproliferative activity of NPs against HT-29 cells. HT-29 cells were exposed to different concentrations of prepared NPs for 72 h. The results revealed the prepared NPs’ impact on the cell viability of the HT-29 cell line. However, the results indicated that the prepared NPs had no cytotoxic against normal cell line (MCF-10). These results were contrasted with control cells, reflecting a concentration-dependent manner, as shown in Figure 14. An in vitro study was prepared, and biosynthesised AgNPs were characterised. According to their findings, AgNPs were highly harmful to A549 cells. To investigate how AgNPs stop cancer cell growth, morphological alterations in A549 cells were analysed using an inverted microscope. The cell cycle distribution was measured using a flow cytometer. Furthermore, a DNA fragment test was used to investigate DNA damage caused by AgNPs in cancer cells; death in cancer cells and DNA fragmentation were observed [41,42]. ZnO nanoparticles had an IC50 value of 38.60 μg·mL−1 in their in vitro anticancer activity against the HeLa cell line, compared to the reference standard cisplatin. This discovery validates the strong cytotoxic effect of ZnO nanoparticles derived from shilajit extract on human cervical cancer cell lines [43].

Figure 13 
                  Antioxidant activity of Ag NPs and ZnO NPs.
Figure 13

Antioxidant activity of Ag NPs and ZnO NPs.

Figure 14 
                  Antiproliferative activity of Ag NPs and ZnO NPs against HT-29 cells (upper panel) and MCF-10 cells (lower panel).
Figure 14

Antiproliferative activity of Ag NPs and ZnO NPs against HT-29 cells (upper panel) and MCF-10 cells (lower panel).

3.7 Molecular docking simulation

Zn and Silver possess several bioactive properties, making them valuable for medical and industrial applications, including antimicrobial, anti-inflammatory, antioxidant, anticancer, wound healing, and anti-biofilm. In silico docking simulations of ZnO NPs and AgNPs revealed a repulsive interaction, resulting in positive binding free energy. This result was expected, as the molecule was composed entirely of Zn and Ag atoms and had a spherical structure. Spherical shapes tend to interact with the surface of proteins rather than being inserted into them, which is thermodynamically favourable. ZnNPs and AgNPs showed the lowest binding free energy with 6KRR, followed by 2GAE, 3Q6A, and 2HQ6, as detailed in Tables 3 and 4. The binding energies suggest a repulsive interaction between ZnNPs and AgNPs and the targeted receptors, leading to a lack of inhibition. Consequently, Ki values were not generated in the DLG file, the output of the docking process, which aligns with the observed non-inhibitory effect. The only interaction that may occur with Cupper atoms is metal acceptor attraction alongside metal donor repulsion, as detected. ZnO NPs interacted with chain A of 2HQ6 through SER73, GLY74, GLN112, and ILE62, while 3Q6A interacted with chain G through GLU14 and TYR48. With 6KRR, ZnO NPs interacted with chain B through THR47, ALA58, and ASP145; on the other hand, with 2GAE, it interacted with chain A through ASN55, GLY57, HIS65, and GLU61, and AgNPs interacted with chain A of 2HQ6 through ASP23, GLY6, ALA21, and PRO9, while with 3Q6A, it interacted with chain G through GLU24 and TYR48. With 6KRR, AgNPs interacted with chain B through THR122, ALA133, and ASP132; on the other hand, with 2GAE, it interacted with chain A through ASN79, GLY108, HIS65, and GLU240. Figures 15–18 detect the various poses of the best conformers from the docking complex and the 3D and 2D visualisations of the hydrophobic interactions identified through LigPlus analysis [44]. Tables 5 and 6 show that 50 conformations were detected from the blind docking simulation between the ZnO NP and AgNP compound and the four targeted proteins, 2HQ6, 2GAE, 3Q6A, and 6KRR.

Table 3

Comparison between the four targeted proteins interacting with ZnO NPs through binding energies, Ref. RMSD, and the possible interactions that occurred

Parameter ZnO NPs with 2HQ6 ZnO NPs with 3Q6A ZnO NPs with 6KRR ZnO NPs with 2GAE
Free energy of binding (kcal·mol−1) +1.23 +1.17 +0.56 +1.01
Ref. RMSD 44.71 105.06 37.37 27.29
Metal accepter interactions. LIG Zn: SER76(A) O LIG Zn: GLU27(G) O LIG Zn: THR124(B) O LIG Zn: ASN81(A) O
LIG Zn: GLY77(A) O, 2.78 A°, 3.38 A° LIG Zn: TYR45(G) O, 2.86 A°, 3.29 A° LIG Zn: ALA136(B) O, 2.29 A°, 2.66 A°, 3.25 A° LIG Zn: GLY106(A) O
LIG Zn: GLY(A) O, 2.65 A°, 3.18 A° LIG Zn: ASP132(B) O, 2.47 A°, 2.72 A°, 2.96 A°, 3.17 A° LIG Zn: HIS240(A) O, 2.85 A°, 3.39 A°
LIG Zn: ALA21(A) O LIG Zn: GLU242(A) O, 2.89 A°, 2.95 A°
LIG Zn: ASP25(A) O, 2.49 A°, 2.59 A°, 2.87 A°, 3.06 A°
Table 4

Comparison between the four targeted proteins interacting with AgNPs through binding energies, Ref. RMSD, and the possible interactions that occurred

Parameter AgNPs with 2HQ6 AgNPs with 2GAE AgNPs with 3Q6A AgNPs with 6KRR
Free energy of binding (kcal·mol−1) +0.18 +0.5 +0.36 +0.56
Ref. RMSD 41.61 29.91 98.25 37.75
Metal accepter interactions LIG Ag: SER76(A) O, 3.05 A°, 3.05 A° LIG Ag: HIS66(A) O, 2.98, 3.09 LIG Ag: GLN39(G) O, 3.28, 3.27, 3.33 LIG Ag: THR46(A) O
LIG Ag: GLY75(A) O, 2.58 A°, 2.77 A°, 2.79 A° LIG Ag: GLY58(A) O
LIG Ag: GLU62(A) O LIG Ag: GLU15(A) O.
LIG Ag: ALA104(A) O LIG Ag: ASN57(A) O
LIG Ag: ASP109(A) O, 2.57 A°, 2.58 A°, 2.58 A°, 2.95 A°, 3.14 A°, 3.26 A°
Figure 15 
                  The best pose and 3D and 2D images of the best conformers resulted from the complexation reaction between ZnO NPs (a: 2HQ6) and (b: 3Q6A).
Figure 15

The best pose and 3D and 2D images of the best conformers resulted from the complexation reaction between ZnO NPs (a: 2HQ6) and (b: 3Q6A).

Figure 16 
                  The best conformers’ best pose, 3D, and 2D images resulted from the complexation reaction between ZnO NPs and (a: 6KRR) and (b: 2GAE).
Figure 16

The best conformers’ best pose, 3D, and 2D images resulted from the complexation reaction between ZnO NPs and (a: 6KRR) and (b: 2GAE).

Figure 17 
                  The best poses, 3D, and hydrophobic 2D images of the best conformers of the complexation reaction between (a: 2HQ6), and (b: 3Q6A) targeted proteins with AgNPs.
Figure 17

The best poses, 3D, and hydrophobic 2D images of the best conformers of the complexation reaction between (a: 2HQ6), and (b: 3Q6A) targeted proteins with AgNPs.

Figure 18 
                  The best poses, 3D, and hydrophobic 2D images of the best conformers of the complexation reaction between (a: 6KRR), and (b: 2GAE) targeted proteins with AgNPs.
Figure 18

The best poses, 3D, and hydrophobic 2D images of the best conformers of the complexation reaction between (a: 6KRR), and (b: 2GAE) targeted proteins with AgNPs.

Table 5

50 conformations detected from the blind docking simulation between the ZnNP compound and the four targeted proteins, 2HQ6, 2GAE, 3Q6A, and 6KRR

Conformer’s no. Binding energy (kcal·mol−1) of ZnNPs with 2HQ6 Binding energy (kcal·mol−1) of ZnNPs with 2GAE Binding energy (kcal·mol−1) of ZnNPs with 3Q6A Binding energy (kcal·mol−1) of ZnNPs with 6KRR
1 +1.42 +1.49 +1.24 +0.77
2 +1.44 +1.46 +1.23 +0.70
3 +1.47 +1.45 +1.49 +0.65
4 +1.41 +1.57 +1.49 +0.72
5 +1.66 +1.62 +1.49 +0.70
6 +1.49 +1.46 +1.50 +0.65
7 +1.42 +1.47 +1.58 +0.67
8 +1.42 +1.43 +1.56 +0.73
9 +1.41 +1.27 +1.55 +0.65
10 +1.24 +1.28 +1.27 +0.64
11 +1.42 +1.45 +1.36 +0.70
12 +1.25 +1.46 +1.45 +0.73
13 +1.23 +1.45 +1.19 +0.65
14 +1.42 +1.47 +1.62 +0.68
15 +1.50 +1.11 +1.45 +0.67
16 +1.49 +1.47 +1.15 +0.57
17 +1.43 +1.47 +1.25 +0.67
18 +1.42 +1.37 +1.28 +0.63
19 +1.46 +1.11 +1.47 +0.65
20 +1.63 +1.41 +1.63 +0.61
21 +1.43 +1.24 +1.46 +0.68
22 +1.45 +1.38 +1.51 +0.67
23 +1.42 +1.46 +1.45 +0.69
24 +1.22 +1.13 +1.41 +0.65
25 +1.42 +1.43 +1.53 +0.61
26 +1.44 +1.36 +1.43 +0.65
27 +1.42 +1.55 +1.53 +0.65
28 +1.44 +1.09 +1.60 +0.73
29 +1.52 +1.48 +1.18 +0.87
30 +1.44 +1.36 +1.24 +0.71
31 +1.54 +1.06 +1.52 +0.70
32 +1.22 +1.27 +1.52 +0.71
33 +1.42 +1.48 +1.58 +0.74
34 +1.43 +1.36 +1.55 +0.73
35 +1.23 +1.28 +1.62 +0.74
36 +1.44 +1.07 +1.25 +0.66
37 +1.48 +1.13 +1.55 +0.64
38 +1.42 +1.42 +1.51 +0.69
39 +1.25 +1.42 +1.53 +0.63
40 +1.53 +1.08 +1.48 +0.68
41 +1.42 +1.08 +1.49 +0.61
42 +1.44 +1.08 +1.53 +0.69
43 +1.45 +1.12 +1.67 +0.66
44 +1.25 +1.08 +1.52 +0.69
45 +1.24 +1.49 +1.45 +0.72
46 +1.21 +1.43 +1.44 +0.63
47 +1.55 +1.37 +1.48 +0.69
48 +1.44 +1.47 +1.25 +0.67
49 +1.43 +1.47 +1.20 +0.68
50 +1.44 +1.43 +1.60 +0.71
Table 6

50 Conformations detected from the blind docking simulation between the AgNPs compound and the four targeted proteins, 2HQ6, 2GAE, 3Q6A, and 6KRR

Conformer’s no. Binding energy (kcal·mol−1) of AgNPs with 2HQ6 Binding energy (kcal·mol−1) of AgNPs with 2GAE Binding energy (kcal·mol−1) of AgNPs with 3Q6A Binding energy (kcal·mol−1) of AgNPs with 6KRR
1 +0.84 +0.95 +1.01 +1.11
2 +0.61 +0.51 +0.81 +0.91
3 +0.91 +1.01 +0.76 +1.03
4 +0.72 +1.02 +0.95 +1.31
5 +0.80 +0.53 +0.80 +1.12
6 +0.94 +0.97 +0.39 +0.91
7 +0.75 +0.30 +0.44 +1.07
8 +0.63 +0.57 +0.91 +0.93
9 +1.07 +0.76 +0.83 +0.97
10 +0.76 +0.86 +1.03 +0.93
11 +0.65 +0.91 +0.88 +0.92
12 +0.73 +0.32 +1.02 +1.14
13 +0.65 +0.89 +0.92 +1.14
14 +0.66 +0.30 +1.01 +0.97
15 +0.61 +0.82 +0.60 +1.18
16 +0.91 +0.32 +0.89 +0.63
17 +0.90 +0.52 +0.75 +1.34
18 +0.91 +1.16 +0.73 +0.94
19 +0.96 +0.87 +0.94 +0.92
20 +0.77 +0.92 +0.93 +0.96
21 +0.87 +1.07 +0.93 +0.99
22 +0.84 +0.78 +0.40 +0.97
23 +0.92 +1.05 +1.05 +1.17
24 +0.64 +0.95 +0.57 +0.77
25 +0.22 +0.95 +0.47 +0.66
26 +0.78 +0.78 +0.95 +0.72
27 +0.85 +0.84 +0.72 +0.93
28 +0.65 +0.78 +0.83 +1.05
29 +1.16 +0.87 +0.98 +1.17
30 +0.64 +0.79 +0.68 +1.13
31 +0.85 +1.17 +0.94 +0.77
32 +0.86 +0.85 +1.03 +0.67
33 +0.93 +1.07 +0.86 +0.46
34 +0.63 +0.75 +0.95 +0.92
35 +0.76 +0.55 +0.89 +1.04
36 +0.77 +0.87 +0.90 +1.02
37 +0.84 +0.95 +1.04 +0.96
38 +1.07 +0.53 +0.64 +1.33
39 +0.76 +0.93 +0.88 +0.77
40 +0.19 +0.83 +0.85 +0.72
41 +0.85 +0.86 +1.32 +0.56
42 +0.87 +0.86 +0.83 +1.17
43 +0.79 +0.74 +0.70 +1.19
44 +0.84 +0.82 +0.58 +0.93
45 +0.92 +0.75 +0.80 +0.89
46 +0.95 +0.89 +1.08 +0.91
47 +0.72 +1.13 +0.91 +0.92
48 +0.92 +0.31 +0.96 +0.71
49 +0.71 +1.02 +0.50 +0.93
50 +0.77 +0.91 +1.17 +1.06

4 Conclusion

In conclusion, this study highlights the potential of green synthesised Ag NPs and ZnO NPs using B. officinalis extracts as a powerful and promising frontier in nanotechnology to address global challenges in health care, environmental sustainability, and antimicrobial resistance. Employing plant extract enhances the biocompatibility and therapeutic potential of the NPs. Ag NPs and ZnO NPs exhibited antimicrobial activity and effectively disturbed and inhibited biofilm formation against the gram-positive and gram-negative bacteria. The synthesised NPs also demonstrated a significant antioxidant activity. This dual functionality of antimicrobials and antioxidants makes the NPs promising candidates for various medical applications, particularly in combating bacterial infections and biofilms. The findings especially offer a promising solution for growing concerns about biofilms-associated and antibiotic resistance. The cytotoxicity assay proved that the produced NPs were selectively harmful to cancer cells. This work significantly contributes to nanotechnology, microbiology, and sustainable materials science and is crucial to future research and advancements in antimicrobial therapies. Further research is required to optimise the synthesis protocols, understand the mechanisms of action, and evaluate the long-term effects of these nanoparticles in vivo. AgNPs exhibit stronger and more stable binding than ZnO NPs, as shown by lower free energy values. ZnO NPs interact with more residues but have higher RMSD, indicating structural flexibility. Among the ligands, 2GAE shows the most stable binding. AgNPs may be more suitable for applications requiring strong and stable interactions.

Acknowledgments

The authors extend their appreciation to the Researchers supporting Project number (RSPD2025R971), King Saud University, Riyadh, Saudi Arabia, for funding this research. The authors appreciated the University of Technology, Iraq, for the logistic support of this work.

  1. Funding information: The authors extend their appreciation to the Researchers supporting Project number (RSPD2025R971), King Saud University, Riyadh, Saudi Arabia, for funding this research.

  2. Author contributions: A. A. AL., H. M. A., Y. B. F., D. S. A., and M. S. J.; writing original draft, methodology, investigation, and formal analysis, M. M. A., A. A. AL., H. M. A., Y. B. F., D. S. A., S. G., A. S., and M. S. J.; main concept, data interpretation, and supervision. A. A. I., A. A. AL., H. M. A., D. S. A. K. K. K., B. F. L., and M. S. J.: writing-review and editing, visualisation, and data curation. All authors approved the final version of the manuscript.

  3. Conflict of interest: The authors state no conflict interests.

  4. Limitation of the study: NPs may not be the most reliable indicator of consumer satisfaction, according to studies.

  5. Data availability statement: The data sets generated during and/or analysed during the current study are available from the corresponding author on reasonable request.

References

[1] Khan I, Saeed K, Khan I. Nanoparticles: Properties, applications and toxicities. Arab J Chem. 2019;12:908–31.10.1016/j.arabjc.2017.05.011Search in Google Scholar

[2] Yin IX, Zhang J, Zhao IS, Mei ML, Li Q, Chu CH. The antibacterial mechanism of silver nanoparticles and its application in dentistry. Int J Nanomed. 2020;15:2555–62.10.2147/IJN.S246764Search in Google Scholar PubMed PubMed Central

[3] Akintelu SA, Folorunso AS. A review on green synthesis of zinc oxide nanoparticles using plant extracts and its biomedical applications. BioNanoScience. 2020;10(4):848–63.10.1007/s12668-020-00774-6Search in Google Scholar

[4] Marimuthu V, Palanisamy SK, Sesurajan S, Sellappa S. Biogenic silver nanoparticles by Gelidiella acerosa extract and their antifungal effects. Avicenna J Med Biotechnol. 2011;3:143–8.Search in Google Scholar

[5] Kisimba K, Krishnan A, Faya M, Byanga K, Kasumbwe K, Vijayakumar K, et al. Synthesis of metallic nanoparticles based on green chemistry and their medical biochemical applications: synthesis of metallic nanoparticles. J Renew Mater. 2023;11(6):2575–91.10.32604/jrm.2023.026159Search in Google Scholar

[6] Waris A, Din M, Ali A, Ali M, Afridi S, Baset A, et al. A comprehensive review of green synthesis of copper oxide nanoparticles and their diverse biomedical applications. Inorg Chem Commun. 2021;123:108369.10.1016/j.inoche.2020.108369Search in Google Scholar

[7] Dehnoee A, Javad Kalbasi R, Zangeneh MM, Delnavazi MR, Zangeneh A. Characterisation, anti-lung cancer activity, and cytotoxicity of bio-synthesised copper nanoparticles by Thymus fedtschenkoi leaf extract. J Clust Sci. 2024;35(3):863–74.10.1007/s10876-023-02512-wSearch in Google Scholar

[8] Pirabbasi E, Zangeneh MM, Zangeneh A, Moradi R, Kalantar M. Chemical characterisation and effect of Ziziphora clinopodioides green-synthesised silver nanoparticles on cytotoxicity, antioxidant, and antidiabetic activities in streptozotocin‐induced hepatotoxicity in Wistar diabetic male rats. Food Sci Nutr. 2024;12(5):3443–51.10.1002/fsn3.4008Search in Google Scholar PubMed PubMed Central

[9] Rezić I. Nanoparticles for biomedical application and their synthesis. Polymers. 2022;14(22):4961.10.3390/polym14224961Search in Google Scholar PubMed PubMed Central

[10] Mayank G, Swati S. Borago officinalis linn. An important medicinal plant of Mediterranean region. Int J Pharm Sci Rev Res. 2010;5(1):27–34.Search in Google Scholar

[11] Singh H, Du J, Yi TH. Green and rapid synthesis of silver nanoparticles using Borago officinalis leaf extract: anticancer and antibacterial activities. Artif Cells Nanomed Biotechnol. 2017;45(7):1310–6.10.1080/21691401.2016.1228663Search in Google Scholar PubMed

[12] Hazaa M, Alm-Eldin M, Ibrahim AE, Elbarky N, Salama M, Sayed R, et al. Biosynthesis of silver nanoparticles using Borago officinslis leaf extract, characterisation and larvicidal activity against cotton leaf worm, Spodoptera littoralis (Bosid). Int J Trop Insect Sci. 2021;41:145–56.10.1007/s42690-020-00187-8Search in Google Scholar

[13] Jasem SA, Abas AK. Physiological effect of green synthesized silver nanoparticles AgNPs on Albino Rat Testis. Egypt J Hosp Med. 2022;89(2):8076–82.10.21608/ejhm.2022.277594Search in Google Scholar

[14] Karimi E, Oskoueian E, Karimi A, Noura R, Ebrahimi M. Borago officinalis L. flower: A comprehensive study on bioactive compounds and its health-promoting properties. J Food Meas Charact. 2018;12:826–38.10.1007/s11694-017-9697-9Search in Google Scholar

[15] Rahmah MI, Saadoon NM, Mohasen AJ, Kamel RI, Fayad TA, Ibrahim NM. Double hydrothermal synthesis of iron oxide/silver oxide nanocomposites with antibacterial activity. J Mech Behav Mater. 2021;30:207–12.10.1515/jmbm-2021-0021Search in Google Scholar

[16] Aziz WJ, Abid MA, Hussein EH. Biosynthesis of CuO nanoparticles and synergistic antibacterial activity using mint leaf extract. Mater Technol. 2020;35:447–51.10.1080/10667857.2019.1692163Search in Google Scholar

[17] Ibrahim NH, Taha GM, Hagaggi NSA, Moghazy MA. Green synthesis of silver nanoparticles and its environmental sensor ability to some heavy metals. BMC Chem. 2024;18(1):7.10.1186/s13065-023-01105-ySearch in Google Scholar PubMed PubMed Central

[18] Nazir A, Akbar A, Baghdadi HB, ur Rehman S, Al-Abbad E, Fatima M, et al. Zinc oxide nanoparticles fabrication using Eriobotrya japonica leaves extract: Photocatalytic performance and antibacterial activity evaluation. Arab J Chem. 2021;14:103251.10.1016/j.arabjc.2021.103251Search in Google Scholar

[19] Alrubaie E, Kadhim RE. Synthesis of ZnO nanoparticles from olive plant extract. Plant Arch. 2019;19:339–44.Search in Google Scholar

[20] Alhujaily M, Jabir MS, Nayef UM, Rashid TM, Sulaiman GM, Khalil KA, et al. Au/ZnO nanocomposites prepared by laser ablation for enhancement of antibacterial activity and cytotoxic properties against cancer cells. Metals. 2023;13(4):735.10.3390/met13040735Search in Google Scholar

[21] Khashan KS, Badr BA, Sulaiman GM, Jabir MS, Hussain SA. Antibacterial activity of Zinc Oxide nanostructured materials synthesis by laser ablation method. In Journal of Physics: Conference Series. Vol. 1795, No. 1, IOP Publishing; 2021. p. 012040.10.1088/1742-6596/1795/1/012040Search in Google Scholar

[22] Al Rugaie O, Jabir MS, Mohammed MK, Abbas RH, Ahmed DS, Sulaiman GM, et al. Modification of SWCNTs with hybrid materials ZnO–Ag and ZnO–Au for enhancing bactericidal activity of phagocytic cells against Escherichia coli through NOX2 pathway. Sci Rep. 2022;12(1):17203.10.1038/s41598-022-22193-1Search in Google Scholar PubMed PubMed Central

[23] Sameen AM, Jabir MS, Al-Ani MQ. Therapeutic combination of gold nanoparticles and LPS as cytotoxic and apoptosis inducer in breast cancer cells. In AIP Conference Proceedings. Vol. 2213, No. 1, 2020.10.1063/5.0000161Search in Google Scholar

[24] Kumavat SR, Mishra S. Green synthesis of silver nanoparticles using Borago officinalis leaves extract and screening its antimicrobial and antifungal activity. Int Nano Lett. 2021;11(4):355–70.10.1007/s40089-021-00345-xSearch in Google Scholar

[25] Kumar H, Rani R. Structural and optical characterisation of ZnO nanoparticles synthesised by microemulsion route. Int Lett Chem Phys Astron. 2013;14:26–36.10.56431/p-q38442Search in Google Scholar

[26] Devi R, Gayathri R. Green synthesis of zinc oxide nanoparticles by using Hibiscus rosa-sinensis. Int J Curr Eng Technol. 2014;4(4):2444–6.Search in Google Scholar

[27] Sharma R, Garg R, Kumari A. A review on biogenic synthesis, applications and toxicity aspects of zinc oxide nanoparticles. EXCLI J. 2020;19:1325.Search in Google Scholar

[28] Abdul Kaleel SH, Bahuleyan BK, Masihullah J, Al-Harthi M. Thermal and mechanical properties of polyethylene/doped‐TiO2 nanocomposites synthesized using in situ polymerization. J Nanomaterials. 2011;2011(1):964353.10.1155/2011/964353Search in Google Scholar

[29] Hamad AM, Atiyea QM. In vitro study of the effect of zinc oxide nanoparticles on Streptococcus mutans isolated from human dental caries. In Journal of Physics: Conference Series. Vol. 1879, No. 2, IOP Publishing; 2021. p. 022041.10.1088/1742-6596/1879/2/022041Search in Google Scholar

[30] Durán N, Nakazato G, Seabra AB. Antimicrobial activity of biogenic silver nanoparticles, and silver chloride nanoparticles: an overview and comments. Appl Microbiol Biotechnol. 2016;100:6555–70.10.1007/s00253-016-7657-7Search in Google Scholar PubMed

[31] Khorrami S, Zarrabi A, Khaleghi M, Danaei M, Mozafari MR. Selective cytotoxicity of green synthesised silver nanoparticles against the MCF-7 tumor cell line and their enhanced antioxidant and antimicrobial properties. Int J Nanomed. 2018;13:8013–24.10.2147/IJN.S189295Search in Google Scholar PubMed PubMed Central

[32] Emami-Karvani Z, Chehrazi P. Antibacterial activity of ZnO nanoparticle on gram-positive and gram-negative bacteria. Afr J Microbiol Res. 2011;5(12):1368–73.10.5897/AJMR10.159Search in Google Scholar

[33] Molina-Hernandez JB, Aceto A, Bucciarelli T, Paludi D, Valbonetti L, Zilli K, et al. The membrane depolarisation and increase intracellular calcium level produced by silver nanoclusters are responsible for bacterial death. Sci Rep. 2021;11(1):21557.10.1038/s41598-021-00545-7Search in Google Scholar PubMed PubMed Central

[34] Abdal Dayem A, Hossain MK, Lee SB, Kim K, Saha SK, Yang GM, et al. The role of reactive oxygen species (ROS) in the biological activities of metallic nanoparticles. Int J Mol Sci. 2017;18(1):120.10.3390/ijms18010120Search in Google Scholar PubMed PubMed Central

[35] Nejad FS, Alizade-Harakiyan M, Haghi M, Ebrahimi R, Zangeneh MM, Farajollahi A, et al. Investigation of the impact of copper nanoparticles coated with ocimum bassilicum at chemoradiotherapy of colon carcinoma. Biochem Biophys Rep. 2024;39:101780.10.1016/j.bbrep.2024.101780Search in Google Scholar PubMed PubMed Central

[36] Bai X, Tang J, Safaralizadeh R, Feizi MA, Delirezh N, Zangeneh A, et al. Green fabrication of bioactive copper nanoparticles using Acroptilon repens extract: An enhanced anti-lung cancer activity. Inorg Chem Commun. 2024;164:112393.10.1016/j.inoche.2024.112393Search in Google Scholar

[37] Mendes CR, Dilarri G, Forsan CF, Sapata VDMR, Lopes PRM, de Moraes PB, et al. Antibacterial action and target mechanisms of zinc oxide nanoparticles against bacterial pathogens. Sci Rep. 2022;12(1):2658.10.1038/s41598-022-06657-ySearch in Google Scholar PubMed PubMed Central

[38] Husain FM, Qais FA, Ahmad I, Hakeem MJ, Baig MH, Masood Khan J, et al. Biosynthesised zinc oxide nanoparticles disrupt established biofilms of pathogenic bacteria. Appl Sci. 2022;12(2):710.10.3390/app12020710Search in Google Scholar

[39] Ansari MA, Khan HM, Khan AA, Cameotra SS, Saquib Q, Musarrat J. Gum arabic capped‐silver nanoparticles inhibit biofilm formation by multi‐drug resistant strains of Pseudomonas aeruginosa. J Basic Microbiol. 2014;54(7):688–99.10.1002/jobm.201300748Search in Google Scholar PubMed

[40] Losada-Barreiro S, Sezgin-Bayindir Z, Paiva-Martins F, Bravo-Díaz C. Biochemistry of antioxidants: Mechanisms and pharmaceutical applications. Biomedicines. 2022;10(12):3051.10.3390/biomedicines10123051Search in Google Scholar PubMed PubMed Central

[41] Djamila B, Eddine LS, Abderrhmane B, Nassiba A, Barhoum A. In vitro antioxidant activities of copper mixed oxide (CuO/Cu2O) nanoparticles produced from the leaves of Phoenix dactylifera L. Biomass Convers Biorefin. 2024;14(5):6567–80.10.1007/s13399-022-02743-3Search in Google Scholar

[42] Alharbi NS, Alsubhi NS. Green synthesis and anticancer activity of silver nanoparticles prepared using fruit extract of Azadirachta indica. J Radiat Res Appl Sci. 2022;15(3):335–45.10.1016/j.jrras.2022.08.009Search in Google Scholar

[43] Perumal P, Sathakkathulla NA, Kumaran K, Ravikumar R, Selvaraj JJ, Nagendran V, et al. Green synthesis of zinc oxide nanoparticles using aqueous extract of shilajit and their anticancer activity against HeLa cells. Sci Rep. 2024;14(1):2204.10.1038/s41598-024-52217-xSearch in Google Scholar PubMed PubMed Central

[44] Issa AA, Alkhafaji AA, Abdulwahid FS, Al-Obaidy RA, Kamel MD, Haider AJ, et al. Nanocluster-based computational creation of a potential carrier for chemotherapeutic antibacterial drugs. J Inorg Organomet Polym Mater. 2024;34:1–16.10.1007/s10904-024-03116-1Search in Google Scholar

Received: 2024-11-13
Accepted: 2025-02-25
Published Online: 2025-03-26

© 2025 the author(s), published by De Gruyter

This work is licensed under the Creative Commons Attribution 4.0 International License.

Articles in the same Issue

  1. Research Articles
  2. Optimized green synthesis of silver nanoparticles from guarana seed skin extract with antibacterial potential
  3. Green adsorbents for water remediation: Removal of Cr(vi) and Ni(ii) using Prosopis glandulosa sawdust and biochar
  4. Green approach for the synthesis of zinc oxide nanoparticles from methanolic stem extract of Andrographis paniculata and evaluation of antidiabetic activity: In silico GSK-3β analysis
  5. Development of a green and rapid ethanol-based HPLC assay for aspirin tablets and feasibility evaluation of domestically produced bioethanol in Thailand as a sustainable mobile phase
  6. A facile biodegradation of polystyrene microplastic by Bacillus subtilis
  7. Enhanced synthesis of fly ash-derived hydrated sodium silicate adsorbents via low-temperature alkaline hydrothermal treatment for advanced environmental applications
  8. Impact of metal nanoparticles biosynthesized using camel milk on bacterial growth and copper removal from wastewater
  9. Preparation of Co/Cr-MOFs for efficient removal of fleroxacin and Rhodamine B
  10. Applying nanocarbon prepared from coal as an anode in lithium-ion batteries
  11. Improved electrochemical synthesis of Cu–Fe/brass foil alloy followed by combustion for high-efficiency photoelectrodes and hydrogen production in alkaline solutions
  12. Precipitation of terephthalic acid from post-consumer polyethylene terephthalate waste fractions
  13. Biosynthesized zinc oxide nanoparticles: Multifunctional potential applications in anticancer, antibacterial, and B. subtilis DNA gyrase docking
  14. Anticancer and antimicrobial effects of green-synthesized silver nanoparticles using Teucrium polium leaves extract
  15. Green synthesis of eco-friendly bioplastics from Chlorella and Lithothamnion algae for safe and sustainable solutions for food packaging
  16. Optimizing coal water slurry concentration via synergistic coal blending and particle size distribution
  17. Green synthesis of Ag@Cu and silver nanowire using Pterospermum heterophyllum extracts for surface-enhanced Raman scattering
  18. Green synthesis of copper oxide nanoparticles from Algerian propolis: Exploring biochemical, structural, antimicrobial, and anti-diabetic properties
  19. Simultaneous quantification of mefenamic acid and paracetamol in fixed-dose combination tablet dosage forms using the green HPTLC method
  20. Green synthesis of titanium dioxide nanoparticles using green tea (Camellia sinensis) extract: Characteristics and applications
  21. Pharmaceutical properties for green fabricated ZnO and Ag nanoparticle-mediated Borago officinalis: In silico predications study
  22. Synthesis and optimization of gemcitabine-loaded nanoparticles by using Box–Behnken design for treating prostate cancer: In vitro characterization and in vivo pharmacokinetic study
  23. A comparative analysis of single-step and multi-step methods for producing magnetic activated carbon from palm kernel shells: Adsorption of methyl orange dye
  24. Sustainable green synthesis of silver nanoparticles using walnut septum waste: Characterization and antibacterial properties
  25. Efficient electrocatalytic reduction of CO2 to CO over Ni/Y diatomic catalysts
  26. Greener and magnetic Fe3O4 nanoparticles as a recyclable catalyst for Knoevenagel condensation and degradation of industrial Congo red dye
  27. Recycling of HDPE-giant reed composites: Processability and performance
  28. Fabrication of antibacterial chitosan/PVA nanofibers co-loaded with curcumin and cefadroxil for wound healing
  29. Cost-effective one-pot fabrication of iron(iii) oxychloride–iron(iii) oxide nanomaterials for supercapacitor charge storage
  30. Novel trimetallic (TiO2–MgO–Au) nanoparticles: Biosynthesis, characterization, antimicrobial, and anticancer activities
  31. Green-synthesized chromium oxide nanoparticles using pomegranate husk extract: Multifunctional bioactivity in antioxidant potential, lipase and amylase inhibition, and cytotoxicity
  32. Therapeutic potential of sustainable zinc oxide nanoparticles biosynthesized using Tradescantia spathacea aqueous leaf extract
  33. Chitosan-coated superparamagnetic iron oxide nanoparticles synthesized using Carica papaya bark extract: Evaluation of antioxidant, antibacterial, and anticancer activity of HeLa cervical cancer cells
  34. Antioxidant potential of peptide fractions from tuna dark muscle protein isolate: A green enzymatic approach
  35. Clerodendron phlomoides leaf extract-mediated synthesis of selenium nanoparticles for multi-applications
  36. Optimization of cellulose yield from oil palm trunks with deep eutectic solvents using response surface methodology
  37. Nitrogen-doped carbon dots from Brahmi (Bacopa monnieri): Metal-free probe for efficient detection of metal pollutants and methylene blue dye degradation
  38. High energy density pseudocapacitor based on a nanoporous tungsten(VI) oxide iodide/poly(2-amino-1-mercaptobenzene) composite
  39. Green synthesized Ag–Cu nanocomposites as an improved strategy to fight multidrug-resistant bacteria by inhibition of biofilm formation: In vitro and in silico assessment study
  40. In vitro evaluation of antibacterial activity and associated cytotoxicity of biogenic silver nanoparticles using various extracts of Tabernaemontana ventricosa
  41. Fabrication of novel composite materials by impregnating ZnO particles into bacterial cellulose nanofibers for antimicrobial applications
  42. Solidification floating organic drop for dispersive liquid–liquid microextraction estimation of copper in different water samples
  43. Kinetics and synthesis of formation of phosphate composites from low-grade phosphorites in the presence of phosphate–siliceous shales and oil sludge
  44. Removal of minocycline and terramycin by graphene oxide and Cr/Mn base metal–organic framework composites
  45. Microfluidic preparation of ceramide E liposomes and properties
  46. Therapeutic potential of Anamirta cocculus (L.) Wight & Arn. leaf aqueous extract-mediated biogenic gold nanoparticles
  47. Review Article
  48. Sustainable innovations in garlic extraction: A comprehensive review and bibliometric analysis of green extraction methods
  49. Rapid Communication
  50. In situ supported rhodium catalyst on mesoporous silica for chemoselective hydrogenation of nitriles to primary amines
  51. Special Issue: Valorisation of Biowaste to Nanomaterials for Environmental Applications
  52. Valorization of coconut husk into biochar for lead (Pb2+) adsorption
  53. Corrigendum
  54. Corrigendum to “An updated review on carbon nanomaterials: Types, synthesis, functionalization and applications, degradation and toxicity”
Downloaded on 13.9.2025 from https://www.degruyterbrill.com/document/doi/10.1515/gps-2024-0237/html
Scroll to top button