Next Article in Journal
Exploring Environmental Settings to Improve the Printability of Paroxetine-Loaded Filaments by Fused Deposition Modelling
Next Article in Special Issue
Levofloxacin and Ciprofloxacin Co-Crystals with Flavonoids: Solid-State Investigation for a Multitarget Strategy against Helicobacter pylori
Previous Article in Journal
Anti-Obesity Drug Delivery Systems: Recent Progress and Challenges
Previous Article in Special Issue
The Role of Five-Membered Heterocycles in the Molecular Structure of Antibacterial Drugs Used in Therapy
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Bioactive ZnO Nanoparticles: Biosynthesis, Characterization and Potential Antimicrobial Applications

by
Md. Amdadul Huq
1,*,
Md. Aminul Islam Apu
2,
Md. Ashrafudoulla
3,
Md. Mizanur Rahman
4,
Md. Anowar Khasru Parvez
5,
Sri Renukadevi Balusamy
6,
Shahina Akter
7 and
Md. Shahedur Rahman
8,*
1
Department of Food and Nutrition, College of Biotechnology and Natural Resource, Chung-Ang University, Anseong 17546, Republic of Korea
2
Department of Nutrition and Hospitality Management, The University of Mississippi, Oxford, MS 38677, USA
3
Department of Food Science and Technology, Chung-Ang University, Anseong 17546, Republic of Korea
4
Department of Biotechnology and Genetic Engineering, Faculty of Biological Science, Islamic University, Kushtia 7003, Bangladesh
5
Department of Microbiology, Jahangirnagar University, Savar, Dhaka 1342, Bangladesh
6
Department of Food Science and Technology, Sejong University, Seoul 05006, Republic of Korea
7
Department of Food Science and Biotechnology, Gachon University, Seongnam 13120, Republic of Korea
8
Department of Genetic Engineering and Biotechnology, Jashore University of Science and Technology, Jashore 7408, Bangladesh
*
Authors to whom correspondence should be addressed.
Submission received: 7 June 2023 / Revised: 22 July 2023 / Accepted: 15 November 2023 / Published: 16 November 2023

Abstract

:
In recent years, biosynthesized zinc oxide nanoparticles (ZnONPs) have gained tremendous attention because of their safe and non-toxic nature and distinctive biomedical applications. A diverse range of microbes (bacteria, fungi and yeast) and various parts (leaf, root, fruit, flower, peel, stem, etc.) of plants have been exploited for the facile, rapid, cost-effective and non-toxic synthesis of ZnONPs. Plant extracts, microbial biomass or culture supernatant contain various biomolecules including enzymes, amino acids, proteins, vitamins, alkaloids, flavonoids, etc., which serve as reducing, capping and stabilizing agents during the biosynthesis of ZnONPs. The biosynthesized ZnONPs are generally characterized using UV-VIS spectroscopy, TEM, SEM, EDX, XRD, FTIR, etc. Antibiotic resistance is a serious problem for global public health. Due to mutation, shifting environmental circumstances and excessive drug use, the number of multidrug-resistant pathogenic microbes is continuously rising. To solve this issue, novel, safe and effective antimicrobial agents are needed urgently. Biosynthesized ZnONPs could be novel and effective antimicrobial agents because of their safe and non-toxic nature and powerful antimicrobial characteristics. It is proven that biosynthesized ZnONPs have strong antimicrobial activity against various pathogenic microorganisms including multidrug-resistant bacteria. The possible antimicrobial mechanisms of ZnONPs are the generation of reactive oxygen species, physical interactions, disruption of the cell walls and cell membranes, damage to DNA, enzyme inactivation, protein denaturation, ribosomal destabilization and mitochondrial dysfunction. In this review, the biosynthesis of ZnONPs using microbes and plants and their characterization have been reviewed comprehensively. Also, the antimicrobial applications and mechanisms of biosynthesized ZnONPs against various pathogenic microorganisms have been highlighted.

1. Introduction

Nanoparticles (NPs) have been proposed as an intervention approach for suppressing microbial growth, as well as contamination, because of their high surface-to-volume ratio. They have distinctive chemical and physical properties that may interfere with bacterial adaptation [1]. Due to their numerous uses in disciplines of research like the health sector, agriculture, textiles, food technology, electronics, and so on, nanoparticles have attracted the attention of scientists [2,3,4,5,6]. However, due to their high propensity, NPs can survive in the environment, and this persistent attribute may be a viable tactic for preventing the bacterial growth used in the manufacturing of various food products including meat products, dairy or vegetables products, sausage products, etc. [7].
Numerous nanomaterials, such as zinc oxide nanoparticles (ZnONPs), silver nanoparticles (AgNPs), gold nanoparticles (AuNPs) and titanium dioxide nanoparticles (TiO2NPs), have potent capacity to both fight bacteria and prevent microbial adhesion, as well as contamination [6,7,8,9,10,11]. Among them, ZnONPs have received a lot of attention due to their safe and non-toxic nature and powerful antibacterial characteristics, which are related to the release of reactive oxygen species (ROS) on their surface [12,13,14,15]. ZnONPs outperform their bulkier counterparts in terms of antibacterial activity because of quantum confinement and size effects [16]. Due to the multiple ways ZnONPs prevent bacterial development, they can succeed easily to protect bacterial-contamination-associated diseases in humans, whereas conventional antibiotics face difficulties to prevent the development of bacterial resistance [12]. Due to its benign properties, ZnO has “generally recognized as safe” (GRAS) classification, and the antimicrobial efficacy of ZnONPs indicates that they are a potent antimicrobial agent for preventing foodborne pathogen contamination in the food sector [17].
These nanoparticles are often created via physical and chemical processes like photochemistry, chemical reduction and microwave irradiation [18,19,20]. The main problems of these techniques are that they are costly, involve labor-intensive equipment and produce harmful consequences due to the use of poisonous substances [21,22]. Nowadays, scholars are focusing on biological strategies for affordable and simple production of nanoparticles due to the different limitations of physicochemical methods. Biological synthesis is a facile, rapid, cost effective, non-toxic and ecofriendly productive method because it is not very expensive, and it can also substitute toxic chemicals and decrease capping and stabilizing agents. A variety of biological resources such as plants and their various parts and different microbes (bacteria, algae, fungi, etc.) could be used for the facile and green synthesis of bioactive nanoparticles [23,24,25,26,27].
A serious problem for global public health is antibiotics resistance. Antibiotics resistance is mostly a result of the abuse of antibiotics. The overuse of antibiotics to treat bacterial infections in humans and aquatic animals has resulted in the spread of numerous antibiotic-resistant strains into the environment [28,29,30]. Since numerous infectious diseases that might be fatal are brought on by pathogenic bacteria, multidrug-resistant microorganisms pose a severe threat to public health globally [31,32,33]. Due to mutation, shifting environmental circumstances and high drug use, the number of multidrug-resistant bacterial strains is continuously rising. To solve this issue, researchers are working to create novel medications for the treatment of these microbial illnesses [9,34]. Biosynthesized ZnONPs could be novel and effective agents to control these multidrug-resistant pathogenic microorganisms because of their safe and non-toxic nature and powerful antibacterial characteristics. Many recent studies have shown that different pathogenic microorganisms can be successfully controlled using biosynthesized ZnONPs [15,35,36,37,38]. This review emphasizes the facile and rapid biological synthesis of ZnONPs using both microbes and plants and their characterizations, potential antimicrobial applications and antimicrobial mechanisms against pathogenic microorganisms.

2. Biosynthesis of ZnONPs

Biosynthesis of ZnONPs is a simple, facile, cost-effective and eco-friendly method compared to the physical and chemical methods that produce various toxic by-products that could be dangerous for our environment [39]. Moreover, biosynthesized nanoparticles are more biocompatible and show significantly higher antimicrobial activity than chemically or physically synthesized nanoparticles [40,41]. For these reasons, scientists are focusing more on utilizing different biological resources for the green, safe and effective synthesis of ZnONPs [3,8,23,36]. For the biosynthesis of ZnONPs, different microorganisms such as bacteria, fungi, yeast or various parts of plants such as leaf, root, fruit, flower, peel, stem, etc. could be used. Figure 1 shows the various steps of facile, cost-effective and eco-friendly biosynthesis of bioactive ZnONPs using the extracts of plants and microbes and their potential antimicrobial efficacy against pathogenic microorganisms.

2.1. Microbe-Mediated Biosynthesis of ZnONPs

Microbe-mediated nanoparticles (NPs) have recently received a lot of attention because of the availability of microorganisms, their easy reproduction and their safe utilization for the biosynthesis of nanoparticles [27,39,42]. Chemical and physical methods can be used to produce NPs, but the microbial synthesis of ZnONPs is considerably more useful than other methods due to their environmental friendliness and low cost. Because of their prevalence in living microorganisms, ZnONPs have become quite popular among other nanoparticles. ZnONPs can be produced by microbial cells, proteins and a variety of enzymes in both prokaryotes and eukaryotes [43]. There are many recent studies on the cost-effective biosynthesis of ZnONPs using various microorganisms such as bacteria, fungi, yeast, algae, etc. (Table 1). Both intracellular and extracellular methods can be used for the facile and eco-friendly synthesis of ZnONPs using microbes [8,13,44]. The culture supernatant of microorganisms and the microbial biomass contain different bioactive compounds including enzymes, proteins, amino acids and many other biomolecules that serve as reducing, capping and stabilizing agents during the synthesis process [30,39]. Previous studies have reported that the bioreduction of Zn2+ was initiated by the electron transfer from NADH by an NADH dependent reductase enzyme that acts as an electron carrier. Consequently, the ZnONPs are formed. Subsequently, various biomolecules such as proteins, amino acids, flavonoids, etc. attached with ZnO and stabilized the ZnONPs [39]. It is also reported that the amino acids present in the proteins were found to interact with the Zn2+ ions to form ZnONPs [39].
Yusof et al. [13] reported the Lactobacillus plantarum TA4 mediated the biosynthesis of ZnONPs using both intracellular and extracellular methods (Figure 2). They added the zinc nitrate solution to the cell-free supernatant (CFS) for the extracellular biosynthesis of ZnONPs; as well, the cell biomass (CB) was added to the zinc nitrate solution for the intracellular biosynthesis of the ZnONPs. The synthesis of NPs was confirmed by visual observation. The synthesized ZnONPs were collected by high-speed centrifugation and dried at 100 °C to obtain the powder form. Through the FTIR analysis, they found different biomolecules present both in the cell-free culture supernatant and in the cell biomass, as well as in the synthesized ZnONPs, and concluded that these biomolecules may be involved as reducing and capping agents during the biosynthesis process [13].
Kumar et al., 2022 [3] reported the extracellular synthesis of bioactive ZnONPs using fungal isolate (Aspergillus sp.). The authors added the culture supernatant dropwise into the zinc acetate solution and confirmed the biosynthesis of ZnONPs by visual observation of color change [3]. Abdo et al. [44] successfully synthesized ZnONPs using cell-free filtrate of P. aeruginosa. They concluded that various metabolites present in the cell-free culture supernatant of P. aeruginosa are responsible for the formation and stabilization of the synthesized ZnONPs [44]. Suba et al. [8] demonstrated the intracellular biosynthesis of ZnONPs using the cell biomass of Lactobacillus spp. within 24 h of reaction and found spherical-shaped ZnONPs with a 32 nm average size. Abdelhakim et al. [24] used endophytic fungi Alternaria tenuissima for the extracellular production of spherical-shaped ZnONPs that possess significant antimicrobial activity against different pathogenic microbes. Table 1 summarizes the microbe-mediated biosynthesis of ZnONPs and their potential antimicrobial applications.

2.2. Plant-Mediated Biosynthesis of ZnONPs

Plant-extract-mediated biosynthesis of ZnONPs has been revealed as a viable option due to its convenience, stability and ease of synthesis compared to all other organisms. Mostly during the synthesis of ZnONPs, extracted phytochemicals function as reducing and capping agents. In the synthesis of ZnONPs as a natural green medium for metallic ion reduction, active bioorganic chemicals in plant extract were crucial [60]. Plant-based NP production has various advantages including minimal cost, ease of use, fast production time, reliability and the ability to scale up production volumes [61]. Furthermore, the availability of bioorganics with many active chemicals in plant components increases demand for ZnONPs, resulting in low-cost, secure and simple syntheses [60]. Various parts of plants such as the roots, shoots, fruits, seeds, leaves, etc. were utilized for the rapid, facile and eco-friendly synthesis of ZnONPs. There are many recent reports on the biosynthesis of ZnONPs and their potential antimicrobial applications using different parts of plants (Table 2). Abomuti et al. [62] reported the plant-mediated biosynthesis of bioactive ZnONPs using the leaf extract of Salvia officinalis. They added aqueous leaf extract to the zinc nitrate solution under constant stirring at a 50 °C temperature. In the second step, they added NaOH solution dropwise under continuous stirring at 50 °C to maintain the stable pH of the reaction mixture. Finally, the biosynthesized ZnONPs were collected by centrifugation and dried to obtain the powder form. Through the FTIR analysis, they found different biomolecules including phenolic and flavonoid compounds present in both the aqueous leaf extract of Salvia officinalis and the synthesized ZnONPs and concluded that these biomolecules may be involved as reducing and capping agents during the biosynthesis process [62]. Fouda et al. [63] reported that the peel extract of Punica granatum mediated the synthesis of bioactive ZnONPs. The author added the aqueous peel extract of Punica granatum into the zinc acetate solution and confirmed the biosynthesis of ZnONPs by visual observation of color change [63]. Fruit extract of Myrica esculenta was used by Lal et al. [23] for the rapid and eco-friendly synthesis of ZnONPs. Urge et al., 2023 [2] successfully synthesized ZnONPs using the bulb extract of Allium sativum and the root extract of Zingiber officinale. They identified various functional groups associated with the formation of ZnONPs [2]. Alotaibi et al. (2022) [64] demonstrated the biosynthesis of ZnONPs using the leaf extract of Gardenia thailandica within 1 h of reaction and found spherical-shaped ZnONPs with a 37.4 nm average size. The leaf extract of Carica papaya was used for rapid and green synthesis of bioactive ZnONPs [65]. Menazea et al. (2021) [66] used the peel extract of orange for the rapid and facile synthesis of cubic-shaped ZnONPs. Suručić et al. (2020) [67] synthesized ZnONPs using flower extract of Geranium robertianum. Plant extract contains various biomolecules such as enzymes, proteins, amino acids, flavonoids, terpenoids and phenolic compounds that play significant roles during the biosynthesis of ZnONPs as reducing and capping agents. Table 2 summarizes the plant-mediated biosynthesis of ZnONPs and their potential antimicrobial applications.

3. Critical Parameters for Rapid and Stable Biosynthesis of ZnONPs

Different parameters significantly affect the rapid and stable synthesis of ZnONPs. Several critical parameters have been identified for the rapid and stable synthesis of ZnONPs, including the concentration of the plant extract and metal salt, incubation time, temperature, pH and stirring rate. The optimal conditions for each parameter may vary depending on the specific plant extract or microbial species and metal salt used. However, some general trends have been observed, such as higher concentrations of plant extracts and metal salts leading to larger yields of nanoparticles, longer incubation times leading to larger particle sizes and higher temperatures leading to faster reaction rates [86,87,88]. The pH of the reaction also significantly affects the rate of ZnONP formation [62,89].

3.1. Factors Influencing the Mass Production of ZnONPs

The mass production of ZnONPs can be affected by several parameters, including the concentration of plant extracts and metal salts, incubation time, temperature, and pH. Higher concentrations of plant extracts and metal salts generally lead to larger yields of nanoparticles, although there may be an optimal concentration beyond which further increases have little effect. Longer incubation times generally lead to larger particle sizes, which can affect the stability and biocompatibility of the nanoparticles. Higher temperatures can lead to faster reaction rates and larger yields of nanoparticles but may also promote agglomeration and reduce the stability of the particles. The pH can also affect the rate of particle formation, with more acidic or alkaline conditions generally leading to faster reaction rates [62,89,90].

3.2. Factors Influencing the Shape and Size of Synthesized ZnONPs

The shape and size of synthesized ZnONP nanoparticles can be influenced by several factors, including the concentration of plant extracts and metal salts, pH, temperature and stirring rate. Higher concentrations of plant extracts and metal salts generally lead to larger particles, while more acidic or alkaline conditions may promote the formation of rod-shaped particles. pH is an important factor for the biosynthesis of ZnONPs and could alter the shape and size of the synthesized nanoparticles [39,91]. Higher temperatures and faster stirring rates can also promote the formation of smaller particles with more uniform shapes. However, other factors, such as the type of plant extract or microbial species and metal salt used, can also play a role in determining the final size and shape of the nanoparticles [47,92,93].

4. Characterization of Biosynthesized ZnONPs

The use of various analytical techniques such as UV-visible spectrophotometry, XRD, SEM, TEM, FTIR, DLS and zeta potential analyzer analysis in the characterization of biosynthesized ZnONPs has been extensively reported in the literature. These techniques provide valuable information on the physical and chemical properties of nanoparticles, including their size, shape, surface charge, crystallinity and surface functional groups. For instance, UV-visible spectrophotometry is commonly used to determine the optical properties of ZnONPs, including their absorption spectra and bandgap energy. In ZnO, like in any other semiconductor, there is a valence band (VB) and a conduction band (CB) separated by a bandgap of a few eV. The ZnO absorption peaks at the transitions between VB and CB. Under irradiation, when enough energy is provided, an electron can be promoted from VB to CB, which will be recorded by a spectrophotometer as an absorption band/peak. The energy value of this peak is related to the value of the bandgap. Recombination of the excited electron from CB with the hole from VB will produce the fluorescent emission at about 380 nm, which is called exciton recombination. The slight variation of the absorption peak appears due to the different intermediary electronic levels generated by impurities or lattice defects [44]. XRD analysis provides information on the crystalline structure and phase purity of nanoparticles. The crystal size of the biosynthesized ZnONPs is generally calculated on the basis of XRD analysis [44]. SEM and TEM techniques are used to visualize the morphology, size and shape of nanoparticles. In both TEM and DLS, the size of nanoparticles and particle size distribution can be determined. While in TEM, the shape and crystallinity can also be determined, in DLS, the obtained size is usually larger due the presence of a solvent layer on the nanoparticle surface. DLS and a zeta potential analyzer provide information on the particle size distribution and surface charge, respectively.
FTIR analysis provides information on the functional groups present on the nanoparticle surface. The quantity of organics from plant or microbial extracts that are adsorbed on the ZnONP surface can be evaluated by thermal analysis. The chemical composition of produced ZnONP samples was also evaluated by using X-ray photoelectron spectroscopy (XPS) [36]. Several studies have reported the use of these techniques to characterize green-synthesized ZnONPs for various applications. For example, Faisal et al. [15] used UV-visible spectrophotometry, XRD, SEM, EDX and FTIR to characterize the Paraclostridium benzoelyticum-bacterium-mediated biosynthesized ZnONPs and investigate their antibacterial, antidiabetic, anti-inflammatory and antiarthritic activities. In another study, Supraja et al. [94] used FTIR, DLS and zeta potential analyzer analysis to characterize Alstonia scholaris stem-bark-extract-mediated ZnONPs and evaluate their antimicrobial efficacy. Abomuti et al. [62] used UV-visible spectrophotometry, Raman spectroscopy, SEM, TEM, XRD and FTIR to characterize the biosynthesized ZnONPs using leaf extract of Salvia officinalis and investigate their antimicrobial activity against Candida albicans isolates. TEM analysis revealed the wurtzite hexagonal shape of synthesized ZnONPs (Figure 3a), and the average size was 26.14 nm (Figure 3b). An SEM image revealed the aggregated form of synthesized ZnONPs and explored some rough, clumsy materials surrounding the ZnONPs (Figure 3c). EDX analysis confirmed the majority of ZnONPs present in the samples. Additional carbon peaks in the EDX spectrum suggested the presence of biomolecules such as vitamins, amino acids, polyphenols, flavonoids and saponins (Figure 3d).
The FTIR spectrum also showed various biomolecules such as polyphenols and other biomolecules present in both aqueous leaf extract of S. officinalis and the synthesized ZnONPs, which suggested that these biomolecules are responsible for the synthesis and stabilization of ZnONPs and their biological activities (Figure 4a–c) [62].
Other studies have reported the use of these techniques to investigate the antimicrobial and antioxidant properties of green-synthesized ZnONPs for various applications. Sonia et al. [95] used UV-visible spectrophotometry, XRD, SEM and DLS to characterize biosynthesized ZnONPs and evaluate their antimicrobial and antioxidant potential for use in a cold-cream formulation. Barsainya and Singh [96] used XRD, SEM and TEM to characterize Pseudomonas aeruginosa-mediated ZnONPs and investigate their broad-spectrum antimicrobial effects. Overall, the use of various analytical techniques in the characterization of biosynthesized ZnONPs has provided valuable insights into their physical and chemical properties, enabling researchers to optimize their synthesis and tailor their properties for specific applications. Table 3 summarizes the different characterization techniques used for biosynthesized ZnONPs.

5. Antimicrobial Applications and Mechanisms of Biosynthesized ZnONPs

In recent years, there has been a growing interest in the development and utilization of nanomaterials for various applications, particularly in the field of antimicrobial research. Among these nanomaterials, ZnONPs have emerged as a promising candidate due to their unique physicochemical properties and potent antimicrobial activity. ZnONPs have been extensively studied for their ability to inhibit the growth of a wide range of microorganisms, including bacteria, fungi and viruses [24,42,62,99,100,101,102,103,104]. ZnONPs have potential applications in various fields, including food, agriculture, pharmaceuticals and biotechnology [43]. In the food and agriculture industries, ZnONPs have been shown to have potential applications as a food preservative and to enhance the antifungal activity of endophytic Bacillus sp. Fcl1. The extracts prepared from the Bacillus sp. Fcl1 cultured in the presence of ZnONPs had an increased production of lipopeptide surfactin derivatives and iturin, which are known for their antimicrobial properties [105]. In the medical field, ZnONPs have shown promise as antimicrobial agents for the treatment of various infections, including skin and wound infections, respiratory tract infections, and urinary tract infections. They have demonstrated broad-spectrum activity against both Gram-positive and Gram-negative bacteria, including multidrug-resistant strains. Furthermore, ZnONPs have been explored for their antifungal activity against pathogenic fungi, such as the Candida species, and have shown potential as antiviral agents against a range of viruses, including herpes simplex virus and influenza virus [106,107]. In the pharmaceutical industry, ZnONPs have been investigated for their potential use as a new antimicrobial agent to combat antibiotic-resistant bacteria. ZnONPs exhibited antimicrobial activity against methicillin-resistant Staphylococcus aureus (MRSA) and vancomycin-resistant Enterococcus faecalis (VRE) [108].
ZnONPs have also been investigated for their potential use in wound healing. The incorporation of ZnONPs into chitosan hydrogels improved their antimicrobial activity against Staphylococcus aureus and Pseudomonas aeruginosa. The use of ZnONPs in wound dressings could be a promising approach to preventing infections and promoting wound healing. In agriculture, ZnONPs have been utilized as antimicrobial agents for crop protection and disease management. They have been shown to effectively inhibit the growth of plant pathogens, including bacteria and fungi, offering an eco-friendly alternative to conventional pesticides. Additionally, the use of ZnONPs in food packaging materials has gained attention due to their antimicrobial properties, which can help extend the shelf life of perishable food products by inhibiting the growth of spoilage microorganisms. Moreover, ZnONPs have been investigated for their potential in environmental remediation, particularly in water treatment, where they can effectively eliminate waterborne pathogens and provide a sustainable approach for disinfection [40,109,110,111,112]. Studies have shown that the antimicrobial activity of ZnO nanoparticles is size-dependent, with smaller particles exhibiting higher antimicrobial activity due to their increased surface area and higher reactivity [113]. In addition, the shape of ZnONPs also plays a crucial role in their antimicrobial activity, with rod-shaped particles exhibiting higher activity than spherical particles [114].
The green synthesis approach utilizes plant extracts, microbes and waste biomaterials as reducing and stabilizing agents, thus reducing the use of hazardous chemicals and energy consumption during the synthesis process. Studies have shown that green synthesis methods produce ZnONPs with superior antimicrobial activity compared to those synthesized using chemical methods. For example, ZnONPs synthesized using aqueous extracts of Heritiera fomes and Sonneratia apetala mangrove plant species showed significant antimicrobial activity against E. coli, S. aureus and B. subtilis [92]. Similarly, Alstonia scholaris stem-bark-extract-mediated ZnONPs demonstrated significant antimicrobial activity against P. aeruginosa, S. aureus and B. subtilis [94]. There are many recent reports on the biosynthesis of ZnONPs using plants and microbes and their potential utilization to control drug-resistant pathogenic microorganisms (Table 1 and Table 2). Abomuti et al. [62] reported the biosynthesis of ZnONPs using leaf extract of Salvia officinalis and evaluated their antimicrobial activity against pathogenic Candida albicans isolates. They found that the biosynthesized ZnONPs strongly suppressed the growth of C. albicans isolates and showed a strong zone of inhibition (Figure 5). Faisal et al. [15] reported on the Paraclostridium benzoelyticum-bacterium-mediated extracellular synthesis of ZnONPs and evaluated their antimicrobial activity against Helicobacter suis, H. felis, H. bizzozeronii and H. salomonis. The biosynthesized ZnONPs strongly inhibited the growth of the tested pathogenic bacteria.
ZnONPs have gained significant attention as promising antimicrobial agents due to their unique physicochemical properties and broad-spectrum activity against various microorganisms. The antimicrobial mechanisms of ZnONPs involve a combination of physical, chemical and biological processes that collectively contribute to their efficacy in inhibiting the growth and survival of microorganisms [107,111,115]. One of the primary mechanisms by which ZnONPs exert their antimicrobial activity is through the generation of reactive oxygen species (ROS). ZnONPs can undergo redox reactions and produce ROS, such as superoxide radicals (O2−), hydrogen peroxide (H2O2) and hydroxyl radicals (OH·). These ROS are highly reactive and can cause oxidative stress in microbial cells by damaging cellular components, including lipids, proteins and nucleic acids. The accumulation of ROS disrupts normal cellular functions, leading to cell membrane damage, protein denaturation and DNA/RNA degradation, ultimately resulting in microbial cell death [111,116,117]. Moreover, ZnONPs possess a high surface-area-to-volume ratio, which enhances their contact with microbial cells and facilitates physical interactions. The small size of ZnONPs allows them to penetrate microbial cell membranes and enter the cytoplasm. Once inside the cell, ZnONPs can interact with intracellular components, such as enzymes and proteins, disrupting their structure and function. This disruption further contributes to the inhibition of microbial growth and proliferation [111,115,118]. Another important antimicrobial mechanism of ZnONPs is their ability to disrupt the integrity and permeability of microbial cell membranes. ZnONPs can interact with the lipid bilayer of the cell membrane, leading to membrane destabilization and increased membrane permeability. This disruption of the cell membrane integrity compromises the structural integrity of microorganisms and leads to leakage of cellular contents, loss of vital ions and ultimately cell death [111,115,119]. Furthermore, ZnONPs have been found to interfere with microbial enzyme activity. Certain enzymes, such as ATPases and respiratory chain enzymes, are crucial for microbial metabolism and energy production. ZnONPs can inhibit the activity of these enzymes, disrupting the energy balance and metabolic processes of microorganisms. This interference with enzyme activity further contributes to the antimicrobial effects of ZnONPs [111,120]. Table 4 summarizes the modes of action of biosynthesized ZnONPs against different pathogenic microbes.
Abomuti et al. [62] applied plant-mediated biosynthesized ZnONPs to treat the pathogenic C. albicans isolates and found that the biosynthesized ZnONPs damaged the cell wall and cell membrane of C. albicans and inhibited the production of ergosterol, which lead to the death of the cell (Figure 6).
According to Ahmed et al. [42], bacterial-mediated biosynthesized ZnONPs effectively control the growth of pathogenic microorganisms B. glumae and B. gladioli. They reported that synthesized ZnONPs damaged the cell membrane, proteins, ribosome and cytoplasmic materials of B. glumae and B. gladioli, produced reactive oxygen species and were involved in the leakage of genetic materials, resulting in cell death (Figure 7).
It is important to note that the antimicrobial mechanisms of ZnONPs can vary depending on the type of microorganism and the specific conditions. While ZnONPs exhibit broad-spectrum antimicrobial activity, some microorganisms may exhibit varying degrees of susceptibility due to differences in cell wall composition, membrane structure or defense mechanisms. In conclusion, ZnONPs possess multiple antimicrobial mechanisms that collectively contribute to their effectiveness in inhibiting the growth and survival of microorganisms. These mechanisms include the generation of reactive oxygen species, physical interactions, disruption of cell walls and cell membranes, damage of DNA, interference with microbial enzyme activity, protein denaturation, ribosomal destabilization and mitochondrial dysfunction. Understanding the antimicrobial mechanisms of ZnONPs is crucial for the development of novel antimicrobial strategies and the optimization of their application in various fields, including medicine, food industry, agriculture and environmental remediation.

6. Conclusions and Future Prospects

The use of green synthesis methods for the production of ZnONPs has emerged as a promising approach to achieve enhanced antimicrobial activity with reduced environmental impact. Biosynthesis of ZnONPs using microbes and plants is a facile, non-toxic, cost-effective and eco-friendly method. In this review, the biosynthesis of ZnONPs using microbes and plants has been comprehensively reviewed. The antimicrobial applications and mechanisms of the biosynthesized ZnONPs against various pathogenic microorganisms have also been highlighted. Plant extracts, microbial biomass or culture supernatant contain various biomolecules including enzymes, amino acids, proteins, vitamins, alkaloids, flavonoids, etc., which serve as reducing, capping and stabilizing agents during the safe, facile and rapid biosynthesis of ZnONPs. The antimicrobial activity of ZnONPs is attributed to several mechanisms, including physical damage to microbial cell walls and cell membranes, production of reactive oxygen species and inhibition of microbial enzyme activity. As ZnONPs are edible and safe for utilization, ZnONPs could be potentially utilized in different food industries to control foodborne pathogens, as well as in many other sectors such as health care and agriculture to effectively control different pathogenic microorganisms. ZnONPs also exhibit antioxidant and wound-healing properties, making them suitable for use in cosmetics and dermatological formulations.
In conclusion, the antimicrobial applications of ZnONPs hold great promise for the development of new antimicrobial agents to combat the growing threat of antimicrobial resistance. ZnONPs have potential applications in various fields, including food, agriculture, pharmaceuticals and biotechnology. It is worth mentioning that the safety and potential toxicity of ZnONPs are important considerations for their practical applications. While ZnONPs have demonstrated significant antimicrobial activity, their potential adverse effects on human health and the environment should be thoroughly evaluated. Proper characterization of ZnONPs, including size, shape and surface modifications, is crucial for understanding their interactions with biological systems and optimizing their antimicrobial efficacy while minimizing potential toxic effects.

Author Contributions

Conceptualization, M.A.H.; writing—original draft preparation, M.A.H., M.A.I.A., M.A. and M.S.R.; writing—review and editing, M.A.H., M.M.R., M.A.K.P., S.R.B. and S.A. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Pezzoni, M.; Catalano, P.N.; Pizarro, R.A.; Desimone, M.F.; Soler-Illia, G.; Bellino, M.G.; Costa, C.S. Antibiofilm effect of supramolecularly templated mesoporous silica coatings. Mater. Sci. Eng. C Mater. Biol. Appl. 2017, 77, 1044–1049. [Google Scholar] [CrossRef]
  2. Urge, S.K.; Dibaba, S.T.; Gemta, A.B. Green Synthesis Method of ZnO Nanoparticles using Extracts of Zingiber officinale and Garlic Bulb (Allium sativum) and Their Synergetic Effect for Antibacterial Activities. J. Nanomater. 2023, 2023, 7036247. [Google Scholar]
  3. Kumar, R.; Vinoth, S.; Baskar, V.; Arun, M.; Gurusaravanan, P. Synthesis of zinc oxide nanoparticles mediated by Dictyota dichotoma endophytic fungi and its photocatalytic degradation of fast green dye and antibacterial applications. S. Afr. J. Bot. 2022, 151, 337–344. [Google Scholar] [CrossRef]
  4. Akter, S.; Huq, M.A. Biologically rapid synthesis of silver nanoparticles by Sphingobium sp. MAH-11T and their antibacterial activity and mechanisms investigation against drug-resistant pathogenic microbes. Artif. Cells Nanomed. Biotechnol. 2020, 48, 672–682. [Google Scholar] [CrossRef]
  5. Bachheti, R.K.; Fikadu, A.; Bachheti, A.; Husen, A. Biogenic fabrication of nanomaterials from flower-based chemical compounds, characterization and their various applications: A review. Saudi J. Biol. Sci. 2020, 27, 2551–2562. [Google Scholar] [CrossRef]
  6. Huq, M.A.; Ashrafudoulla, M.; Rahman, M.M.; Balusamy, S.R.; Akter, S. Green Synthesis and Potential Antibacterial Applications of Bioactive Silver Nanoparticles: A Review. Polymers 2022, 14, 742. [Google Scholar] [CrossRef]
  7. Motelica, L.; Ficai, D.; Oprea, O.; Ficai, A.; Trusca, R.-D.; Andronescu, E.; Holban, A.M. Biodegradable Alginate Films with ZnO Nanoparticles and Citronella Essential Oil—A Novel Antimicrobial Structure. Pharmaceutics 2021, 13, 1020. [Google Scholar] [CrossRef]
  8. Suba, S.; Vijayakumar, S.; Vidhya, E.; Punitha, V.N.; Nilavukkarasi, M. Microbial mediated synthesis of ZnO nanoparticles derived from Lactobacillus spp.: Characterizations, antimicrobial and biocompatibility efficiencies. Sens. Int. 2021, 2, 100104. [Google Scholar] [CrossRef]
  9. Ghdeeb, N.J.; Hussain, N.A. Antimicrobial Activity of ZnO Nanoparticles Prepared Using a Green Synthesis Approach. Nano Biomed. Eng. 2023, 15, 14–20. [Google Scholar] [CrossRef]
  10. Huq, M.A.; Ashrafudoulla, M.; Parvez, M.A.K.; Balusamy, S.R.; Rahman, M.M.; Kim, J.H.; Akter, S. Chitosan-Coated Polymeric Silver and Gold Nanoparticles: Biosynthesis, Characterization and Potential Antibacterial Applications: A Review. Polymers 2022, 14, 5302. [Google Scholar] [CrossRef]
  11. Huq, M.A.; Akter, S. Bacterial mediated rapid and facile synthesis of silver nanoparticles and their antimicrobial efficacy against pathogenic microorganisms. Materials 2021, 14, 2615. [Google Scholar] [CrossRef] [PubMed]
  12. Venkatasubbu, G.D.; Baskar, R.; Anusuya, T.; Seshan, C.A.; Chelliah, R. Toxicity mechanism of titanium dioxide and zinc oxide nanoparticles against food pathogens. Colloids Surf. B Biointerfaces 2016, 148, 600–606. [Google Scholar] [CrossRef] [PubMed]
  13. Yusof, H.M.; Rahman, N.A.; Mohamad, R.; Zaidan, U.H.; Samsudin, A.A. Biosynthesis of zinc oxide nanoparticles by cell-biomass and supernatant of Lactobacillus plantarum TA4 and its antibacterial and biocompatibility properties. Sci. Rep. 2020, 10, 19996. [Google Scholar] [CrossRef]
  14. Khan, M.A.; Lone, S.A.; Shahid, M.; Zeyad, M.T.; Syed, A.; Ehtram, A.; Elgorban, A.M.; Verma, M.; Danish, M. Phytogenically Synthesized Zinc Oxide Nanoparticles (ZnO-NPs) Potentially Inhibit the Bacterial Pathogens: In Vitro Studies. Toxics 2023, 11, 452. [Google Scholar] [CrossRef]
  15. Faisal, S.; Abdullah; Rizwan, M.; Ullah, R.; Alotaibi, A.; Khattak, A.; Bibi, N.; Idrees, M. Paraclostridium Benzoelyticum Bacterium-Mediated Zinc Oxide Nanoparticles and Their In Vivo Multiple Biological Applications. Oxid. Med. Cell. Longev. 2022, 2022, 5994033. [Google Scholar]
  16. Motelica, L.; Vasile, B.-S.; Ficai, A.; Surdu, A.-V.; Ficai, D.; Oprea, O.-C.; Andronescu, E.; Jinga, D.C.; Holban, A.M. Influence of the Alcohols on the ZnO Synthesis and Its Properties: The Photocatalytic and Antimicrobial Activities. Pharmaceutics 2022, 14, 2842. [Google Scholar] [CrossRef] [PubMed]
  17. Tayel, A.A.; El-Tras, W.F.; Moussa, S.; El-Baz, A.F.; Mahrous, H.; Salem, M.F.; Brimer, L. Antibacterial action of zinc oxide nanoparticles against foodborne pathogens. J. Food Saf. 2011, 31, 211–218. [Google Scholar] [CrossRef]
  18. Elemike, E.E.; Onwudiwe, D.C.; Fayemi, O.E.; Botha, T.L. Green synthesis and electrochemistry of Ag, Au, and Ag–Au bimetallic nanoparticles using golden rod (Solidago canadensis) leaf extract. Appl. Phys. A 2019, 125, 42. [Google Scholar] [CrossRef]
  19. Guzmán, M.G.; Dille, J.; Godet, S. Synthesis of silver nanoparticles by chemical reduction method and their antibacterial activity. Int. J. Chem. Biomol. Eng. 2009, 2, 104–111. [Google Scholar]
  20. Pauzi, N.; Zain, N.M.; Yusof, N.A.A. Microwave-assisted synthesis of ZnO nanoparticles stabilized with gum arabic: Effect of microwave irradiation time on ZnO nanoparticles size and morphology. Bull. Chem. React. Eng. Catal. 2019, 14, 182–188. [Google Scholar] [CrossRef]
  21. Iravani, S.; Korbekandi, H.; Mirmohammadi, S.V.; Zolfaghari, B. Synthesis of silver nanoparticles: Chemical, physical and biological methods. Res. Pharm. Sci. 2014, 9, 385. [Google Scholar]
  22. Jamkhande, P.G.; Ghule, N.W.; Bamer, A.H.; Kalaskar, M.G. Metal nanoparticles synthesis: An overview on methods of preparation, advantages and disadvantages, and applications. J. Drug Deliv. Sci. Technol. 2019, 53, 101174. [Google Scholar] [CrossRef]
  23. Lal, S.; Verma, R.; Chauhan, A.; Dhatwalia, J.; Guleria, I.; Ghotekar, S.; Thakur, S.; Mansi, K.; Kumar, R.; Kumari, A.; et al. Antioxidant, antimicrobial, and photocatalytic activity of green synthesized ZnO-NPs from Myrica esculenta fruits extract. Inorg. Chem. Comm. 2022, 141, 109518. [Google Scholar] [CrossRef]
  24. Abdelhakim, H.K.; El-Sayed, E.R.; Rashidi, F.B. Biosynthesis of Zinc Oxide Nanoparticles with Antimicrobial, Anticancer, Antioxidant and Photocatalytic Activities by the Endophytic Alternaria tenuissima. J. Appl. Microbiol. 2020, 128, 1634–1646. [Google Scholar] [CrossRef] [PubMed]
  25. Akter, S.; Lee, S.-Y.; Siddiqi, M.Z.; Balusamy, S.R.; Ashrafudoulla, M.; Rupa, E.J.; Huq, M.A. Ecofriendly synthesis of silver nanoparticles by Terrabacter humi sp. nov. and their antibacterial application against antibiotic-resistant pathogens. Int. J. Mol. Sci. 2020, 21, 9746. [Google Scholar] [CrossRef] [PubMed]
  26. Du, J.; Singh, H.; Yi, T.H. Antibacterial, anti-biofilm and anticancer potentials of green synthesized silver nanoparticles using benzoin gum (Styrax benzoin) extract. Bioprocess Biosyst. Eng. 2020, 39, 1923–1931. [Google Scholar] [CrossRef]
  27. Wang, X.; Lee, S.-Y.; Akter, S.; Huq, M.A. Probiotic-mediated biosynthesis of silver nanoparticles and their antibacterial applications against pathogenic strains of Escherichia coli O157:H7. Polymers 2022, 14, 1834. [Google Scholar] [CrossRef]
  28. Ashrafudoulla, M.; Mizan, M.F.R.; Park, S.H.; Ha, S.D. Current and future perspectives for controlling Vibrio biofilms in the seafood industry: A comprehensive review. Crit. Rev. Food Sci. Nutr. 2021, 61, 1827–1851. [Google Scholar] [CrossRef]
  29. Hamida, R.M.; Ali, M.A.; Goda, D.A.; Khalilad, M.I.; Redhwan, A. Cytotoxic effect of green silver nanoparticles against ampicillin-resistant Klebsiella pneumoniae. RSC Adv. 2020, 10, 21136–21146. [Google Scholar] [CrossRef]
  30. Huq, M.A.; Akter, S. Biosynthesis, characterization and antibacterial application of novel silver nanoparticles against drug resistant pathogenic Klebsiella pneumoniae and Salmonella Enteritidis. Molecules 2021, 26, 5996. [Google Scholar] [CrossRef]
  31. Ashrafudoulla, M.; Mizan, M.F.R.; Park, H.; Byun, K.H.; Lee, N.; Park, S.H.; Ha, S.D. Genetic Relationship, Virulence Factors, Drug Resistance Profile and Biofilm Formation Ability of Vibrio parahaemolyticus Isolated from Mussel. Front. Microbiol. 2019, 10, 513. [Google Scholar] [CrossRef]
  32. Afshari, A.; Baratpour, A.; Khanzade, S.; Jamshidi, A. Salmonella Enteritidis and Salmonella Typhimorium identification in poultry carcasses. Iran. J. Microbiol. 2018, 10, 45–50. [Google Scholar] [PubMed]
  33. Huq, M.A.; Akter, S. Characterization and Genome Analysis of Arthrobacter bangladeshi sp. nov., Applied for the Green Synthesis of Silver Nanoparticles and Their Antibacterial Efficacy against Drug-Resistant Human Pathogens. Pharmaceutics 2021, 13, 1691. [Google Scholar] [CrossRef] [PubMed]
  34. Huq, M.A. Biogenic silver nanoparticles synthesized by Lysinibacillus xylanilyticus MAHUQ-40 to control antibiotic-resistant human pathogens vibrio parahaemolyticus and salmonella typhimurium. Front. Bioeng. Biotechnol. 2020, 8, 1407. [Google Scholar] [CrossRef]
  35. Barani, M.; Masoudi, M.; Mashreghi, M.; Makhdoumi, A.; Eshghi, H. Cell-free extract assisted synthesis of ZnO nanoparticles using aquatic bacterial strains: Biological activities and toxicological evaluation. Int. J. Pharm. 2021, 606, 120878. [Google Scholar] [CrossRef] [PubMed]
  36. MuthuKathija, M.; Badhusha, M.S.M.; Rama, V. Green synthesis of zinc oxide nanoparticles using Pisonia alba leaf extract and its antibacterial activity. Appl. Surf. Sci. Adv. 2023, 15, 100400. [Google Scholar] [CrossRef]
  37. Chowdhury, M.A.H.; Ashrafudoulla, M.; Mevo, S.I.U.; Mizan, M.F.R.; Park, S.H.; Ha, S.D. Current and future interventions for improving poultry health and poultry food safety and security: A comprehensive review. Compr. Rev. Food Sci. Food Saf. 2023, 22, 1555–1596. [Google Scholar] [CrossRef]
  38. Rambabu, K.; Bharath, G.; Banat, F.; Show, P.L. Green synthesis of zinc oxide nanoparticles using Phoenix dactylifera waste as bioreductant for effective dye degradation and antibacterial performance in wastewater treatment. J. Hazard. Mater. 2021, 402, 123560. [Google Scholar] [CrossRef]
  39. Gomaa, E.Z. Microbial Mediated Synthesis of Zinc Oxide Nanoparticles, Characterization and Multifaceted Applications. J. Inorg. Organomet. Polym. 2022, 32, 4114–4132. [Google Scholar] [CrossRef]
  40. Singh, A.; Singh, N.B.; Afzal, S.; Singh, T.; Hussain, I. Zinc oxide nanoparticles: A review of their biological synthesis, antimicrobial activity, uptake, translocation and biotransformation in plants. J. Mater. Sci. 2018, 53, 185–201. [Google Scholar] [CrossRef]
  41. Raja, A.; Ashokkumar, S.; Pavithra-Marthandam, R.; Jayachandiran, J.; Khatiwada, C.P.; Kaviyarasu, K.; Ganapathi Raman, R.; Swaminathan, M. Eco-friendly preparation of zinc oxide nanoparticles using Tabernaemontana divaricata and its photocatalytic and antimicrobial activity. J. Photochem. Photobiol. B Biol. 2018, 181, 53–58. [Google Scholar] [CrossRef] [PubMed]
  42. Ahmed, T.; Wu, Z.; Jiang, H.; Luo, J.; Noman, M.; Shahid, M.; Manzoor, I.; Allemailem, K.S.; Alrumaihi, F.; Li, B. Bioinspired Green Synthesis of Zinc Oxide Nanoparticles from a Native Bacillus cereus Strain RNT6: Characterization and Antibacterial Activity against Rice Panicle Blight Pathogens Burkholderia glumae and B. gladioli. Nanomaterials 2021, 11, 884. [Google Scholar] [CrossRef] [PubMed]
  43. Yusof, H.M.; Mohamad, R.; Zaidan, U.H.; Abdul Rahman, N.A. Microbial synthesis of zinc oxide nanoparticles and their potential application as an antimicrobial agent and a feed supplement in animal industry: A review. J. Anim. Sci. Biotechnol. 2019, 10, 57. [Google Scholar]
  44. Abdo, A.M.; Fouda, A.; Eid, A.M.; Fahmy, N.M.; Elsayed, A.M.; Khalil, A.M.A.; Alzahrani, O.M.; Ahmed, A.F.; Soliman, A.M. Green Synthesis of Zinc Oxide Nanoparticles (ZnO-NPs) by Pseudomonas aeruginosa and Their Activity against Pathogenic Microbes and Common House Mosquito, Culex pipiens. Materials 2021, 14, 6983. [Google Scholar] [CrossRef] [PubMed]
  45. Jayabalan, J.; Mani, G.; Krishnan, N.; Pernabas, J.; Devadoss, J.M.; Jang, H.T. Green Biogenic Synthesis of Zinc Oxide Nanoparticles Using Pseudomonas putida Culture and Its In Vitro Antibacterial and Anti-Biofilm Activity. Biocatal. Agric. Biotechnol. 2019, 21, 101327. [Google Scholar] [CrossRef]
  46. Jayaseelan, C.; Rahuman, A.A.; Kirthi, A.V.; Marimuthu, S.; Santhoshkumar, T.; Bagavan, A.; Gaurav, K.; Karthik, L.; Rao, K.V.B. Novel microbial route to synthesize ZnO nanoparticles using Aeromonas hydrophila and their activity against pathogenic bacteria and fungi. Spectrochim. Acta Part A Mol. Biomol. Spectrosc. 2012, 90, 78–84. [Google Scholar] [CrossRef] [PubMed]
  47. Saravanan, M.; Gopinath, V.; Chaurasia, M.K.; Syed, A.; Ameen, F.; Purushothaman, N. Green synthesis of anisotropic zinc oxide nanoparticles with antibacterial and cytofriendly properties. Microb. Pathog. 2018, 115, 57–63. [Google Scholar] [CrossRef]
  48. Taran, M.; Rad, M.; Alavi, M. Biosynthesis of TiO2 and ZnO nanoparticles by Halomonas elongata IBRC-M 10214 in different conditions of medium. Bioimpacts 2018, 8, 81–89. [Google Scholar] [CrossRef]
  49. Król, A. Mechanism study of intracellular zinc oxide nanocomposites formation. Colloids Surf. 2018, 553, 349–358. [Google Scholar] [CrossRef]
  50. Mishra, M.; Paliwal, J.; Singh, S.; Ethiraj, S.; Mohanashrinivasan, V. Studies on the inhibitory activity of biologically synthesized and characterized ZnO nanoparticles using L.sporogens against Staphylococcus aureus. J. Pure Appl. Microbiol. 2013, 7, 1263–1268. [Google Scholar]
  51. Kundu, D.; Hazra, C.; Chatterjee, A.; Chaudhari, A.; Mishra, S. Extracellular biosynthesis of zinc oxide nanoparticles using Rhodococcus pyridinivorans NT2: Multifunctional textile finishing, biosafety evaluation and in vitro drug delivery in colon carcinoma. J. Photochem. Photobiol. B Biol. 2014, 140, 194–204. [Google Scholar] [CrossRef] [PubMed]
  52. Rajabairavi, N.; Raju, C.S.; Karthikeyan, C.; Varutharaju, K.; Nethaji, S.; Hameed, A.S.H.; Shajahan, A. Biosynthesis of Novel Zinc Oxide Nanoparticles (ZnO NPs) Using Endophytic Bacteria Sphingobacterium thalpophilum. In Recent Trends in Materials Science and Applications; Ebenezar, J., Ed.; Springer: Cham, Switzerland, 2017; pp. 245–254. [Google Scholar]
  53. Rauf, M.A.; Owais, M.; Rajpoot, R.; Ahmad, F.; Khan, N.; Zubair, S. Biomimetically synthesized ZnO nanoparticles attain potent antibacterial activity against less susceptible S. aureus skin infection in experimental animals. RSC Adv. 2017, 7, 36361–36373. [Google Scholar] [CrossRef]
  54. Balraj, B.; Senthilkumar, N.; Siva, C.; Krithikadevi, R.; Julie, A.; Potheher, I.V.; Arulmozhi, M. Synthesis and characterization of Zinc Oxide nanoparticles using marine Streptomyces sp. with its investigations on anticancer and antibacterial activity. Res. Chem. Intermed. 2017, 43, 2367–2376. [Google Scholar] [CrossRef]
  55. Moghaddam, A.B.; Moniri, M.; Azizi, S.; Rahim, R.A.; Ariff, A.B.; Saad, W.Z.; Namvar, F.; Navaderi, M. Biosynthesis of ZnO Nanoparticles by a New Pichia kudriavzevii Yeast Strain and Evaluation of Their Antimicrobial and Antioxidant Activities. Molecules 2017, 22, 872. [Google Scholar] [CrossRef]
  56. Chauhan, R.; Reddy, A.; Abraham, J. Biosynthesis of silver and zinc oxide nanoparticles using Pichia fermentans JA2 and their antimicrobial property. Appl. Nanosci. 2015, 5, 63–71. [Google Scholar] [CrossRef]
  57. Rajan, A.; Cherian, E.; Gurunathan, D.B. Biosynthesis of zinc oxide nanoparticles using Aspergillus fumigatus JCF and its antibacterial activity. Int. J. Mod. Sci. Technol. 2016, 1, 52–57. [Google Scholar]
  58. Kalpana, V.N.; Kataru, B.A.S.; Sravani, N.; Vigneshwari, T.; Panneerselvam, A.; Devi Rajeswari, V. Biosynthesis of zinc oxide nanoparticles using culture filtrates of Aspergillus niger: Antimicrobial textiles and dye degradation studies. OpenNano 2018, 3, 48–55. [Google Scholar] [CrossRef]
  59. Gurunathan, D.B.; Jagadeesan, C.; Fahad, K.; Praveen, A. Mycological synthesis, characterization and antifungal activity of zinc oxide nanoparticles. Asian J. Pharm. Technol. 2013, 3, 142–146. [Google Scholar]
  60. Fagier, M.A. Plant-Mediated Biosynthesis and Photocatalysis Activities of Zinc Oxide Nanoparticles: A Prospect towards Dyes Mineralization. J. Nanotechnol. 2021, 2021, 6629180. [Google Scholar] [CrossRef]
  61. Li, X.; Xu, H.; Chen, Z.-S.; Chen, G. Biosynthesis of Nanoparticles by Microorganisms and Their Applications. J. Nanomater. 2011, 2011, 270974. [Google Scholar] [CrossRef]
  62. Abomuti, M.A.; Danish, E.Y.; Firoz, A.; Hasan, N.; Malik, M.A. Green Synthesis of Zinc Oxide Nanoparticles Using Salvia officinalis Leaf Extract and Their Photocatalytic and Antifungal Activities. Biology 2021, 10, 1075. [Google Scholar] [CrossRef] [PubMed]
  63. Fouda, A.; Saied, E.; Eid, A.M.; Kouadri, F.; Alemam, A.M.; Hamza, M.F.; Alharbi, M.; Elkelish, A.; Hassan, S.E.-D. Green Synthesis of Zinc Oxide Nanoparticles Using an Aqueous Extract of Punica granatum for Antimicrobial and Catalytic Activity. J. Funct. Biomater. 2023, 14, 205. [Google Scholar] [CrossRef] [PubMed]
  64. Alotaibi, B.; Negm, W.A.; Elekhnawy, E.; El-Masry, T.A.; Elharty, M.E.; Saleh, A.; Abdelkader, D.H.; Mokhtar, F.A. Antibacterial activity of nano zinc oxide green-synthesised from Gardenia thailandica triveng. Leaves against Pseudomonas aeruginosa clinical isolates: In vitro and in vivo study. Artif. Cells Nanomed. Biotechnol. 2022, 50, 96–106. [Google Scholar] [CrossRef]
  65. Dulta, K.; Koşarsoy Ağçeli, G.; Chauhan, P.; Jasrotia, R.; Chauhan, P.K. Ecofriendly Synthesis of Zinc Oxide Nanoparticles by Carica papaya Leaf Extract and Their Applications. J. Clust. Sci. 2022, 33, 603–617. [Google Scholar] [CrossRef]
  66. Menazea, A.A.; Ismail, A.M.; Samy, A. Novel Green Synthesis of Zinc Oxide Nanoparticles Using Orange Waste and Its Thermal and Antibacterial Activity. J. Inorg. Organomet. Polym. 2021, 31, 4250–4259. [Google Scholar] [CrossRef]
  67. Suručić, R.; Šmitran, A.; Gajić, D.; Božić, L.; Antić, M.; Topić-Vučenović, V.; Umičević, N.; Antunović, V.; Jelić, D. Phytosynthesis of zinc oxide nanoparticles with acetonic extract of flowers of Geranium robertianum L. (Geraniaceae). J. Hyg. Eng. Des. 2020, 615, 28. [Google Scholar]
  68. Subramanian, H.; Krishnan, M.; Mahalingam, A. Photocatalytic dye degradation and photoexcited anti-microbial activities of green zinc oxide nanoparticles synthesized via Sargassum muticum extracts. RSC Adv. 2022, 12, 985–997. [Google Scholar] [CrossRef]
  69. Abdelmigid, H.M.; Hussien, N.A.; Alyamani, A.A.; Morsi, M.M.; AlSufyani, N.M.; Kadi, H.A. Green Synthesis of Zinc Oxide Nanoparticles Using Pomegranate Fruit Peel and Solid Coffee Grounds vs. Chemical Method of Synthesis, with Their Biocompatibility and Antibacterial Properties Investigation. Molecules 2022, 27, 1236. [Google Scholar] [CrossRef]
  70. Ramesh, R.; Parasaran, M.; Mubashira, G.T.F.; Flora, C.; Khan, F.L.A.; Almaary, K.S.; Elbadawi, Y.B.; Chen, T.-W.; Kanimozhi, K.; Bashir, A.; et al. Biogenic synthesis of ZnO and NiO nanoparticles mediated by fermented Cocos nucifera. (L) Deoiled cake extract for antimicrobial applications towards gram positive and gram negative pathogens. J. King Saud. Univ. Sci. 2022, 34, 101696–101704. [Google Scholar] [CrossRef]
  71. Alahmdi, M.I.; Khasim, S.; Vanaraj, S.; Panneerselvam, C.; Mahmoud, M.A.A.; Mukhtar, S.; Alsharif, M.A.; Zidan, N.S.; Abo-Dya, N.E.; Aldosari, O.F. Green Nanoarchitectonics of ZnO Nanoparticles from Clitoria ternatea Flower Extract for In Vitro Anticancer and Antibacterial Activity: Inhibits MCF-7 Cell Proliferation via Intrinsic Apoptotic Pathway. J. Inorg. Organomet. Polym. 2022, 32, 2146–2159. [Google Scholar] [CrossRef]
  72. Ilangovan, A.; Venkatramanan, A.; Thangarajan, P.; Saravanan, A.; Rajendran, S.; Kaveri, K. Green Synthesis of Zinc Oxide Nanoparticles (ZnO NPs) Using Aqueous Extract of Tagetes Erecta Flower and Evaluation of Its Antioxidant, Antimicrobial, and Cytotoxic Activities on HeLa Cell Line. Curr. Biotechnol. 2021, 10, 61–76. [Google Scholar] [CrossRef]
  73. Djouadi, A.; Derouiche, S. Spinach mediated synthesis of zinc oxide nanoparticles: Characterization, In vitro biological activities study and in vivo acute toxicity evaluation. Curr. Res. Green Sustain. Chem. 2021, 4, 100214. [Google Scholar] [CrossRef]
  74. Khan, M.I.; Fatima, N.; Shakil, M.; Tahir, M.B.; Riaz, K.N.; Rafique, M.; Iqbal, T.; Mahmood, K. Investigation of in-vitro antibacterial and seed germination properties of green synthesized pure and nickel doped ZnO nanoparticles. Phys. B Condens. Matter 2021, 601, 412563. [Google Scholar] [CrossRef]
  75. Vijayakumar, S.; González-Sánchez, Z.I.; Malaikozhundan, B.; Saravanakumar, K.; Divya, M.; Vaseeharan, B.; Durán-Lara, E.F.; Wang, M.-H. Biogenic Synthesis of Rod Shaped ZnO Nanoparticles Using Red Paprika (Capsicum annuum L. var. grossum (L.) Sendt) and Their in Vitro Evaluation. J. Clust. Sci. 2021, 32, 1129–1139. [Google Scholar] [CrossRef]
  76. Batool, M.; Khurshid, S.; Qureshi, Z.; Daoush, W.M. Adsorption, antimicrobial and wound healing activities of biosynthesised zinc oxide nanoparticles. Chem. Pap. 2021, 75, 893–907. [Google Scholar] [CrossRef]
  77. Kumar, N.H.K.; Mohana, N.C.; Nuthan, B.R.; Ramesha, K.P.; Rakshith, D.; Geetha, N.; Satish, S. Phyto-mediated synthesis of zinc oxide nanoparticles using aqueous plant extract of Ocimum americanum and evaluation of its bioactivity. SN Appl. Sci. 2019, 1, 651. [Google Scholar] [CrossRef]
  78. Madan, H.R.; Sharma, S.C.; Udayabhanu; Suresh, D.; Vidya, Y.S.; Nagabhushana, H.; Rajanaik, H.; Anantharaju, K.S.; Prashantha, S.C.; Sadananda-Maiya, P. Facile green fabrication of nanostructure ZnO plates, bullets, flower, prismatic tip, closed pine cone: Their antibacterial, antioxidant, photoluminescent and photocatalytic properties. Spectrochim. Acta Part A Mol. Biomol. Spectrosc. 2016, 152, 404–416. [Google Scholar] [CrossRef]
  79. Chauhan, A.; Verma, R.; Kumari, S.; Sharma, A.; Shandilya, P.; Li, X.; Batoo, K.M.; Imran, A.; Kulshrestha, S.; Kumar, R. Photocatalytic dye degradation and antimicrobial activities of Pure and Ag-doped ZnO using Cannabis sativa leaf extract. Sci. Rep. 2020, 10, 7881. [Google Scholar] [CrossRef]
  80. Sharma, S.C. ZnO nano-flowers from Carica papaya milk: Degradation of Alizarin Red-S dye and antibacterial activity against Pseudomonas aeruginosa and Staphylococcus aureus. Optik 2016, 127, 6498–6512. [Google Scholar] [CrossRef]
  81. Kahsay, M.H.; Tadesse, A.; RamaDevi, D.; Belachew, N.; Basavaiah, K. Green synthesis of zinc oxide nanostructures and investigation of their photocatalytic and bactericidal applications. RSC Adv. 2019, 9, 36967–36981. [Google Scholar] [CrossRef]
  82. Pal, S.; Mondal, S.; Maity, J.; Mukherjee, R. Synthesis and Characterization of ZnO Nanoparticles using Moringa oleifera Leaf Extract: Investigation of Photocatalytic and Antibacterial Activity. Int. J. Nanosci. Nanotechnol. 2018, 14, 111–119. [Google Scholar]
  83. Jayappa, M.D.; Ramaiah, C.K.; Kumar, M.A.P.; Suresh, D.; Prabhu, A.; Devasya, R.P.; Sheikh, S. Green synthesis of zinc oxide nanoparticles from the leaf, stem and in vitro grown callus of Mussaenda frondosa L.: Characterization and their applications. Appl. Nanosci. 2020, 10, 3057–3074. [Google Scholar] [CrossRef] [PubMed]
  84. Khalid, A.; Ahmad, P.; Khandaker, M.U.; Modafer, Y.; Almukhlifi, H.A.; Bazaid, A.S.; Aldarhami, A.; Alanazi, A.M.; Jefri, O.A.; Uddin, M.; et al. Biologically Reduced Zinc Oxide Nanosheets Using Phyllanthus emblica Plant Extract for Antibacterial and Dye Degradation Studies. J. Chem. 2023, 2023, 3971686. [Google Scholar] [CrossRef]
  85. Chandrasekar, L.P.; Sethuraman, B.D.; Subramani, M.; Mohandos, S. Green synthesised ZnO nanoparticles from Plectranthus amboinicus plant extract: Removal of Safranin-O and Malachite green dyes & anti-bacterial activity. Int. J. Environ. Anal. Chem. 2023, 1–18. [Google Scholar] [CrossRef]
  86. Ahmed, S.; Chaudhry, S.A.; Ikram, S. A review on biogenic synthesis of ZnO nanoparticles using plant extracts and microbes: A prospect towards green chemistry. J. Photochem. Photobiol. B Biol. 2017, 166, 272–284. [Google Scholar] [CrossRef]
  87. Gur, T.; Meydan, I.; Seckin, H.; Bekmezci, M.; Sen, F. Green synthesis, characterization and bioactivity of biogenic zinc oxide nanoparticles. Environ. Res. 2022, 204, 111897. [Google Scholar] [CrossRef]
  88. Iqbal, J.; Abbasi, B.A.; Yaseen, T.; Zahra, S.A.; Shahbaz, A.; Shah, S.A.; Uddin, S.; Ma, X.; Raouf, B.; Kanwal, S.; et al. Green synthesis of zinc oxide nanoparticles using Elaeagnus angustifolia L. leaf extracts and their multiple in vitro biological applications. Sci. Rep. 2021, 11, 20988. [Google Scholar] [CrossRef]
  89. Chikkanna, M.M.; Neelagund, S.E.; Rajashekarappa, K.K. Green synthesis of zinc oxide nanoparticles (ZnO NPs) and their biological activity. SN Appl. Sci. 2019, 1, 117. [Google Scholar] [CrossRef]
  90. Yazdanian, M.; Rostamzadeh, P.; Rahbar, M.; Alam, M.; Abbasi, K.; Tahmasebi, E.; Tebyaniyan, H.; Ranjbar, R.; Seifalian, A.; Yazdanian, A. The potential application of green-synthesized metal nanoparticles in dentistry: A comprehensive review. Bioinorg. Chem. Appl. 2022, 2022, 2311910. [Google Scholar] [CrossRef]
  91. Verma, A.; Mehata, M.S. Controllable synthesis of silver nanoparticles using Neem leaves and their antimicrobial activity. J. Radiat. Res. Appl. Sci. 2016, 9, 109–115. [Google Scholar] [CrossRef]
  92. Thatoi, P.; Kerry, R.G.; Gouda, S.; Das, G.; Pramanik, K.; Thatoi, H.; Patra, J.K. Photo-mediated green synthesis of silver and zinc oxide nanoparticles using aqueous extracts of two mangrove plant species, Heritiera fomes and Sonneratia apetala and investigation of their biomedical applications. J. Photochem. Photobiol. B Biol. 2016, 3, 163. [Google Scholar] [CrossRef] [PubMed]
  93. Vimala, K.; Sundarraj, S.; Paulpandi, M.; Vengatesan, S.; Kannan, S. Green synthesized doxorubicin loaded zinc oxide nanoparticles regulates the Bax and Bcl-2 expression in breast and colon carcinoma. Process Biochem. 2014, 49, 160–172. [Google Scholar] [CrossRef]
  94. Supraja, N.; Prasad, T.; Gandhi, A.D.; Anbumani, D.; Kavitha, P.; Babujanarthanam, R. Synthesis, characterization and evaluation of antimicrobial efficacy and brine shrimp lethality assay of Alstonia scholaris stem bark extract mediated ZnONPs. Biochem. Biophys. Rep. 2018, 14, 69–77. [Google Scholar] [CrossRef] [PubMed]
  95. Sonia, S.; Ruckmani, K.; Sivakumar, M. Antimicrobial and antioxidant potentials of biosynthesized colloidal zinc oxide nanoparticles for a fortified cold cream formulation: A potent nanocosmeceutical application. Mater. Sci. Eng. C 2017, 79, 581–589. [Google Scholar]
  96. Manjari, B.; Singh, D.P. Green synthesis of zinc oxide nanoparticles by Pseudomonas aeruginosa and their broad-spectrum antimicrobial effects. J. Pure Appl. Microbiol. 2018, 12, 2123–2134. [Google Scholar]
  97. Al-Kurdy, M.J.; Al-Khuzaie, M.G.A.; Abbas, G.A.; Al Dulaimi, Z.M.H. Green synthesis and characterization of zinc oxide nanoparticles using black currant extracts. AIP Conf. Proc. 2023, 1, 2776. [Google Scholar]
  98. Narayanan, K.B.; Sakthivel, N. Phytosynthesis of gold nanoparticles using leaf extract of Coleus amboinicus Lour. Mater. Charact. 2010, 61, 1232–1238. [Google Scholar] [CrossRef]
  99. Darvish, M.; Ajji, A. Effect of Polyethylene Film Thickness on the Antimicrobial Activity of Embedded Zinc Oxide Nanoparticles. ACS Omega 2021, 6, 26201–26209. [Google Scholar] [CrossRef]
  100. Halbus, A.F.; Horozov, T.S.; Paunov, V.N. Surface-Modified Zinc Oxide Nanoparticles for Antialgal and Antiyeast Applications. ACS Appl. Nano Mater. 2020, 3, 440–451. [Google Scholar] [CrossRef]
  101. Sun, Q.; Li, J.; Le, T. Zinc Oxide Nanoparticle as a Novel Class of Antifungal Agents: Current Advances and Future Perspectives. J. Agric. Food Chem. 2018, 66, 11209–11220. [Google Scholar] [CrossRef]
  102. Morowvat, M.H.; Kazemi, K.; Jaberi, M.A.; Amini, A.; Gholami, A. Biosynthesis and Antimicrobial Evaluation of Zinc Oxide Nanoparticles Using Chlorella vulgaris Biomass against Multidrug-Resistant Pathogens. Materials 2023, 16, 842. [Google Scholar] [CrossRef]
  103. Bahari, N.; Hashim, N.; Abdan, K.; Md Akim, A.; Maringgal, B.; Al-Shdifat, L. Role of Honey as a Bifunctional Reducing and Capping/Stabilizing Agent: Application for Silver and Zinc Oxide Nanoparticles. Nanomaterials 2023, 13, 1244. [Google Scholar] [CrossRef] [PubMed]
  104. Li, W.; You, Q.; Zhang, J. Green synthesis of antibacterial LFL-ZnO using L. plantarum fermentation liquid assisted by ultrasound-microwave. J. Alloys Compd. 2023, 947, 169697. [Google Scholar] [CrossRef]
  105. Ravi, A.; Nandayipurath, V.V.T.; Rajan, S.; Salim, S.A.; Khalid, N.K.; Aravindakumar, C.T.; Krishnankutty, R.E. Effect of zinc oxide nanoparticle supplementation on the enhanced production of surfactin and iturin lipopeptides of endophytic Bacillus sp. Fcl1 and its ameliorated antifungal activity. Microb. Pathog. 2020, 149, 104528. [Google Scholar] [CrossRef] [PubMed]
  106. Mishra, P.K.; Mishra, H.; Ekielski, A.; Talegaonkar, S.; Vaidya, B. Zinc oxide nanoparticles: A promising nanomaterial for biomedical applications. Drug Discov. Today 2017, 22, 1825–1834. [Google Scholar] [CrossRef]
  107. Jin, S.-E.; Jin, H.-E. Antimicrobial Activity of Zinc Oxide Nano/Microparticles and Their Combinations against Pathogenic Microorganisms for Biomedical Applications: From Physicochemical Characteristics to Pharmacological Aspects. Nanomaterials 2021, 11, 263. [Google Scholar] [CrossRef]
  108. Kim, Y.S.; Kim, J.S.; Cho, H.S.; Rha, D.S.; Kim, J.M.; Park, J.D.; Choi, B.S.; Lim, R.; Chang, H.K.; Chung, Y.H.; et al. Twenty-eight-day oral toxicity, genotoxicity, and gender-related tissue distribution of silver nanoparticles in Sprague-Dawley rats. Inhal. Toxicol. 2008, 20, 575–583. [Google Scholar] [CrossRef]
  109. Espitia, P.J.P.; Soares, N.F.F.; Coimbra, J.S.R.; De Andrade, N.J.; Cruz, R.S.; Medeiros, E.A.A. Zinc Oxide Nanoparticles: Synthesis, Antimicrobial Activity and Food Packaging Applications. Food Bioprocess Technol. 2012, 5, 1447–1464. [Google Scholar] [CrossRef]
  110. Kalpana, V.N.; Devi Rajeswari, V. A Review on Green Synthesis, Biomedical Applications, and Toxicity Studies of ZnO NPs. Bioinorg. Chem. Appl. 2018, 2018, 3569758. [Google Scholar] [CrossRef]
  111. Sirelkhatim, A.; Mahmud, S.; Seeni, A.; Kaus, N.H.M.; Ann, L.C.; Bakhori, S.K.M.; Hasan, H.; Mohamad, D. Review on Zinc Oxide Nanoparticles: Antibacterial Activity and Toxicity Mechanism. Nano-Micro Lett. 2015, 7, 219–242. [Google Scholar] [CrossRef]
  112. Kumar, R.; Umar, A.; Kumar, G.; Nalwa, H.S. Antimicrobial properties of ZnO nanomaterials: A review. Ceram. Int. 2017, 43, 3940–3961. [Google Scholar] [CrossRef]
  113. Sportelli; Chiara, M.; Picca, R.A.; Cioffi, N. Recent advances in the synthesis and characterization of nano-antimicrobials. TrAC Trends Anal. Chem. 2016, 84, 131–138. [Google Scholar] [CrossRef]
  114. Halbus, A.F.; Tommy, S. Colloid particle formulations for antimicrobial applications. Adv. Colloid Interface Sci. 2017, 249, 134–148. [Google Scholar] [CrossRef] [PubMed]
  115. Raghupathi, K.R.; Koodali, R.T.; Manna, A.C. Size-dependent bacterial growth inhibition and mechanism of antibacterial activity of zinc oxide nanoparticles. Langmuir 2011, 27, 4020–4028. [Google Scholar] [CrossRef] [PubMed]
  116. Xie, Y.; He, Y.; Irwin, P.L.; Jin, T.; Shi, X. Antibacterial activity and mechanism of action of zinc oxide nanoparticles against Campylobacter jejuni. Appl. Environ. Microbiol. 2011, 77, 2325–2331. [Google Scholar] [CrossRef]
  117. Lipovsky, A.; Nitzan, Y.; Gedanken, A.; Lubart, R. Antifungal activity of ZnO nanoparticles--the role of ROS mediated cell injury. Nanotechnology 2011, 22, 105101. [Google Scholar] [CrossRef]
  118. Arakha, M.; Saleem, M.; Mallick, B.C.; Jha, S. The effects of interfacial potential on antimicrobial propensity of ZnO nanoparticle. Sci. Rep. 2015, 5, 9578. [Google Scholar] [CrossRef]
  119. Da-Silva, B.L.; Abuçafy, M.P.; Manaia, E.B.; Oshiro, J.A., Jr.; Chiari-Andréo, B.G.; Pietro, R.C.R.; Chiavacci, L.A. Relationship between structure and antimicrobial activity of zinc oxide nanoparticles: An overview. Int. J. Nanomed. 2019, 14, 9395. [Google Scholar] [CrossRef]
  120. Jeyabharathi, S.; Chandramohan, S.; Naveenkumar, S.; Sundar, K.; Muthukumaran, A. Synergistic effects of herbal zinc oxide nanoparticles (ZnONPs) and its anti-hyperglycemic and anti-bacterial effects. Mater. Today Proc. 2021, 36, 390–396. [Google Scholar] [CrossRef]
  121. Obeizi, Z.; Benbouzid, H.; Ouchenane, S.; Yılmaz, D.; Culha, M.; Bououdina, M. Biosynthesis of Zinc Oxide Nanoparticles from Essential Oil of Eucalyptus globulus with Antimicrobial and Anti-Biofilm Activities. Mater. Today Commun. 2020, 25, 225164972. [Google Scholar] [CrossRef]
Figure 1. Schematic illustration of biosynthesis and potential antimicrobial applications of bioactive ZnONPs.
Figure 1. Schematic illustration of biosynthesis and potential antimicrobial applications of bioactive ZnONPs.
Pharmaceutics 15 02634 g001
Figure 2. (a) Reduction of Zn2+ to ZnONPs by (i) cell-free supernatant and (ii) cell biomass of L. plantarum TA4. (b) UV-Vis spectrum of (i) ZnONPs-CFS and (ii) ZnONPs-CB. This figure has been reprinted with permission from Ref. [13], copyright 2020, Nature Portfolio.
Figure 2. (a) Reduction of Zn2+ to ZnONPs by (i) cell-free supernatant and (ii) cell biomass of L. plantarum TA4. (b) UV-Vis spectrum of (i) ZnONPs-CFS and (ii) ZnONPs-CB. This figure has been reprinted with permission from Ref. [13], copyright 2020, Nature Portfolio.
Pharmaceutics 15 02634 g002
Figure 3. (a) Transmission electron microscope (TEM) scale bar: 100 nm; (b) particle size distribution histogram; (c) scanning electron microscopy (SEM) scale bar: 1 μm; (d) the EDX spectra of biosynthesized ZnONPs using leaf extract of Salvia officinalis. This figure has been reprinted with permission from Ref. [62], copyright 2021, MDPI.
Figure 3. (a) Transmission electron microscope (TEM) scale bar: 100 nm; (b) particle size distribution histogram; (c) scanning electron microscopy (SEM) scale bar: 1 μm; (d) the EDX spectra of biosynthesized ZnONPs using leaf extract of Salvia officinalis. This figure has been reprinted with permission from Ref. [62], copyright 2021, MDPI.
Pharmaceutics 15 02634 g003
Figure 4. FTIR spectra of (a) aqueous leaf extract of S. officinalis, (b) ZnONPs dried at 80 °C, and (c) ZnONPs calcinated at 400 °C. This figure has been reprinted with permission from Ref. [62], copyright 2021, MDPI.
Figure 4. FTIR spectra of (a) aqueous leaf extract of S. officinalis, (b) ZnONPs dried at 80 °C, and (c) ZnONPs calcinated at 400 °C. This figure has been reprinted with permission from Ref. [62], copyright 2021, MDPI.
Pharmaceutics 15 02634 g004
Figure 5. Zones of inhibition around discs impregnated with 1% DMSO (A), 2 µg/mL amphotericin B (B), ½ MIC of ZnONPs (C), MIC of ZnONPs (D), and MFC of ZnONPs (E) against different Candida albicans isolates. This figure has been reprinted with permission from Ref. [62], copyright 2021, MDPI.
Figure 5. Zones of inhibition around discs impregnated with 1% DMSO (A), 2 µg/mL amphotericin B (B), ½ MIC of ZnONPs (C), MIC of ZnONPs (D), and MFC of ZnONPs (E) against different Candida albicans isolates. This figure has been reprinted with permission from Ref. [62], copyright 2021, MDPI.
Pharmaceutics 15 02634 g005
Figure 6. Scanning electron micrographs (SEM) of C. albicans SC5314: (A) represents untreated control cells, whereas (B,C) represent the cells exposed to MIC and MFC of biosynthesized ZnONPs, respectively. This figure has been reprinted with permission from Ref. [62], copyright 2021, MDPI.
Figure 6. Scanning electron micrographs (SEM) of C. albicans SC5314: (A) represents untreated control cells, whereas (B,C) represent the cells exposed to MIC and MFC of biosynthesized ZnONPs, respectively. This figure has been reprinted with permission from Ref. [62], copyright 2021, MDPI.
Pharmaceutics 15 02634 g006
Figure 7. SEM (a) and TEM (b) images of rice bacterial pathogen B. glumae and B. gladioli cells after 8 h treatment with (50 µg mL−1) and without (control) biogenic ZnONPs. This figure has been reprinted with permission from Ref. [42], copyright 2021, MDPI.
Figure 7. SEM (a) and TEM (b) images of rice bacterial pathogen B. glumae and B. gladioli cells after 8 h treatment with (50 µg mL−1) and without (control) biogenic ZnONPs. This figure has been reprinted with permission from Ref. [42], copyright 2021, MDPI.
Pharmaceutics 15 02634 g007
Table 1. Microbe-mediated biosynthesis and potential antimicrobial applications of ZnONPs. NA, Not Available.
Table 1. Microbe-mediated biosynthesis and potential antimicrobial applications of ZnONPs. NA, Not Available.
Microbes Used for SynthesisSynthesis MethodOptimum Synthesis Conditions (Salt Concentration, Temperature, Incubation Time)Size (nm, Nanoparticles/Crystallite)ShapeTarget PathogensReference
Paraclostridium benzoelyticumExtra cellular0.1 M zinc nitrate, 80 °C for 24 h50 (Average)Spherical and rectangularHelicobacter suis, H. felis, H. bizzozeronii, H. salomonis[15]
Aspergillus sp.Extra cellular0.1 N zinc acetate, 40 °C for 6 h80–100Sphere shapeEscherchia coli, Pseudomonas aeruginosa, Salmonella typhi[3]
Pseudomonas aeruginosaExtra cellular2 mM zinc acetate, 35 ± 2 °C for 24 h14.9 ± 3.5SphericalStaphylococcus aureus, Escherichia coli, Bacillus subtilis, Pseudomonas aeruginosa, Candida albicans[44]
Lactobacillus spp.Intracellular500 mM zinc salt, 37 °C for 24 h32 (Average)SphericalClostridium difficile, E. coli, Clostridium perfringens, S. typhi, Aspergillus flavus, C. albicans[8]
Marinobacter sp. 2C8 and Vibrio sp. VLAExtra cellular0.1 M zinc sulfate, 30 °C for 24 h10.2–20.3SphericalE. coli, P. aeruginosa, Listeria innocua, S. aureus, Bacillus subtilis[35]
Bacillus cereus RNT6Extra cellular0.1 zinc sulfate, 80 °C for 15 min21–35SphericalBurkholderia glumae, B. gladioli[42]
Lactobacillus plantarum TA4Extra and intracellular500 mM zinc salt, 24 h at 37 °C152.8–613.5Flower patternE. coli, Salmonella sp., S. aureus, S. epidermidis[13]
Endophytic fungus Alternaria tenuissimaExtra cellular2 mM zinc sulphate, at room temperature for 20 min10–30SphericalP. aeruginosa, Klebsiella pneumoniae, E. coli, S. aureus[24]
Pseudomonas putidaCombine of intra and extracellular100 mg zinc nitrate into 100 mL culture solution, 24 h at 37 °C44.5 (Average)SphericalPseudomonas otitidis, Enterococcus faecalis, Acinetobacter baumannii, P. oleovorans, B. cereus[45]
Aeromonas hydrophilaIntracellularZinc salt, 37 °C, for 24 h57.7 (Average)SphericalP. aeruginosa, Aspergillus flavus[46]
Bacillus megateriumIntracellularZinc nitrate solution, 37 °C for 48 h45–95Rod and cubicHelicobacter pylori[47]
Halomonas elongateExtracellularZinc chloride, 37 °C for one week18.1 ± 8.9MultiformE. coli, S. aureus[48]
Lactobacillus paracasei LB3IntracellularZinc nitrate solution, 37 °C for 24 h1179 ± 137SphericalS. aureus, Acetinobacter baumannii[49]
Lactobacillus sporogensExtracellular0.1 M zinc sulfate, 37 °C for 24 h145.7 (Average)HexagonalS. aureus[50]
Rhodococcus pyridinivorans NT2bExtracellular0.1 M zinc sulfate, 30 °C for 72 h100–120Roughly sphericalS. epidermidis[51]
Sphingobacterium thalpophilumExtracellularZinc nitrate solution, 37 °C for 24 h40 (Average)TriangleP. aeruginosa, Enterobacter aerogens[52]
Staphylococcus aureusExtracellularzinc acetate solution (1 mM), 37 °C.10–50AcicularS. aureus[53]
Streptomyces sp.ExtracellularZinc chloride solution, 28 °C for 7 days20–50SphericalE. coli, B. subtilis[54]
Pichia kudriavzeviiExtracellularzinc acetate solution, 35 °C for 36 h10–61Hexagonal wurtziteB. subtilis, S. epidermidis, S. aurous, E. coli, Serratia marcescens[55]
Pichia fermentas JA2Extracellular1 mM zinc nitrate, 28 °C for 96 hNASmooth and elongatedP. aeruginosa[56]
Aspergillus fumigatus JCFExtracellular1.0 mM zinc sulfate, 32 °C for 96 h60–80SphericalK. pneumoniae, P. aeruginosa, E. coli, S. aureus, B. subtilis[57]
Aspergillus nigerExtracellular5 mM Zinc nitrate, 32 °C for 48 h61 ± 0.65SphericalE. coli, S. aureus[58]
Aspergillus terreusExtracellularZinc salt solution, 32 °C for 4 days54.8–82.6SphericalA. niger, A. fumigatus, A. aculeatus[59]
Table 2. Plant-mediated biosynthesis and potential antimicrobial applications of ZnONPs.
Table 2. Plant-mediated biosynthesis and potential antimicrobial applications of ZnONPs.
PlantUsed PartOptimum Synthesis Conditions (Salt Concentration, Temperature, Incubation Time)Size (nm, Nanoparticles/Crystallite)ShapeTarget PathogensReference
Punica granatumPeel extract5 mM Zinc acetate, room temperature for overnight10–45SphericalStaphylococcus aureus, Bacillus subtilis, Pseudomonas aeruginosa, Escherichia coli, Candida albicans[63]
Cassia siameaLeaf extract1.0 mM zinc nitrate, heated for 3 to 4 h13 (Average)Spherical, oval, spheroidalPseudomonas aeruginosa, Chromobacterium violaceum[14]
Cinnamon and bayLeavesZinc salt, room temperature for 24 h~10, 18.5 and ~30 (Average)SphericalStaphylococcus aureus, Staphylococcus epidermidis, Escherichia coli, Klebsiella pneumoniae[9]
Allium sativum, Zingiber officinaleBulb extract, root extractZinc acetate solution, 50 °C for 2 h19.8, 21.9 and 23.9 (Average)WurtziteEscherichia coli, Pseudomonas putida, Staphylococcus aureus, Streptococcus pyogenes[2]
Pisonia AlbaLeaf extract0.1 M zinc acetate, 70 °C for 2 hAggregatedNAStaphylococcus aureus, Klebsiella pneumoniae[36]
Sargassum muticumPlant extract5 mM zinc nitrate, 70 °C for 20 min and room temperature for 2 h15–50Wurtzite hexagonalBacillus flexus, Bacillus filamentosus, Acinetobacter baumannii, Pseudomonas stutzeri[68]
Punica granatum peel and coffee groundPlant extract10 mM zinc acetate, 1 h at 70 °C118.6, 115.7 and 111.2 (Average)NanorodPseudomonas aeruginosa, Staphylococcus aureus, Klebsiella pneumoniae, Enterobacter aerogenes[69]
Myrica esculentaFruits extract0.5 M zinc acetate, 40 °C for 2 h31.7 (Average)NAFusarium oxysporum, Staphylococcus aureus, Pseudomonas aeruginosa, Rosellinia necatrix, Escherichia coli[23]
Gardenia thailandica trivengLeavesZinc acetate solution, 70 °C for 30 min, room temperature for 1 h37.4 (Average)SphericalPseudomonas aeruginosa clinical isolates[64]
Cocos nuciferaExtract1 M zinc nitrate, 4 h at ambient temperature28–59Rock shapedS. aureus, E. coli, B. subtilis, K. pneumoniae[70]
Clitoria ternateaFlower extract0.1 M zinc nitrate, 4 h at 80 °C40–81RodS. aureus, E. coli[71]
Carica papayaLeaf extract0.1 M Zinc acetate, 4 h at 80 °C15–50Semi-sphericalRosellinia necatrix, Sclerotinia sclerotiorum,
Fusarium spp.
[65]
Tagetes erectaFlower extract1.5 mM zinc nitrate, 24 h at 60 °C30–50SphericalE. coli, S. aureus[72]
Spinacea oleraceaExtractAqueous zinc acetate solution, 24 h at 60 °C13.0 (Average)granularPseudomonas aeruginosa[73]
Salvia officinalisLeaf extract0.1 M zinc nitrate, 4 h at 50 °C26.1 (Average)Wurtzite hexagonalCandida albicans isolates[62]
OrangePeel extract1 M zinc nitrate, 2 h at room temperature 20–60cubicPseudomonas aeruginosa, B. subtilis[66]
Phoenix dactyliferaWaste5 g zinc nitrate in 50 mL of extract, 30 min at room temperature30 (Average)SphericalStreptococcus pyogenes, Pseudomonas aeruginosa, Staphylococcus aureus[38]
Brassica rapaLeaf extractZinc nitrate solution, 4 h at 80 °C27.5 (Average)IrregularMicrococcus luteus, Enterobacter aerogenes[74]
Red PaprikaAqueous plant extract2 M Zinc acetate, 6 h at room temperature70–80RodS. enterica. [75]
Aloe barbadenseLeaf extract10 mM zinc nitrate, 60 °C44 (Average)Quasi-hexagonalBacillus subtilis, Bacillus licheniformis, Klebsiella pneumonia, Escherichia coli, Candida albicans, Aspergillus niger[76]
Geranium robertianumFlower extract10 mM zinc acetate, 2 h at room temperature40 (Average)IrregularEscherichia coli, Pseudomonas aeruginosa, Acinetobacter baumannii, Staphylococcus
aureus isolates
[67]
Ocimum americanumPlant extract1 mM zinc nitrate, 1 h at 60 °C21 (Average)SphericalB. cereus, Staphylococcus aureus, Klebsiella pneumonia, Vibrio parahaemolyticus, Pseudomonas aeruginosa, Escherichia coli, Salmonella typhi, Candida albicans, Xanthomonas citri, Aspergillus parasiticus[77]
Azadirachta indicaLeavesZinc nitrate solution, boiled at 350 ± 10 °C for 4 min9–38HexagonalKlebsiella aerogenes and Staphylococcus aureus[78]
Cannabis sativaLeafZinc acetate solution, 80 °C for 12 h34–38SphericalEscherichia coli, Klebsiella pneumonia, MRSA, Pseudomonas aeruginosa, Salmonella typhi, Staphylococcus aureus[79]
Carica papayaLatexZinc nitrate solution, 37 °C for 36 h11–26HexagonalPseudomonas aeruginosa and Staphylococcus aureus compared to Klebsiella aerogenes and Pseudomonas desmolyticum[80]
Dolichos lablab L.LeafZinc acetate solution, incubated 70 °C for 1 h29 (Average)HexagonalBacillus pumilus and Sphingomonas paucimobilis[81]
Tabernaemontana
divaricata
Green leafZinc nitrate solution, 80 °C until precipitation.20–50SphericalSalmonella paratyphi, Escherichia coli and Staphylococcus aureus[41]
Moringa oleifera (drumstick)LeavesZinc acetate solution, 24 °C for 1 h52 (Average)Hexagonal wurtziteBacillus subtilis and Escherichia coli[82]
Mussaenda frondosaLeaf/stemZinc nitrate solution, 400 °C for 10–30 min5–20SphericalStaphylococcus aureus and Bacillus subtilis[83]
Phyllanthus emblicaPlant extractZinc chloride solution, 90 °C for 2 hAggregatedsquare-shapedS. pyogenes, S. aureus, S. typhi and E. coli[84]
Plectranthus amboinicusPlant extractZinc sulfate solution, room temperature for 2 hAggregatedIrregular aggregated nanoflakesS. aureus and E. coli[85]
Table 3. Different characterization techniques used for biosynthesized ZnONPs.
Table 3. Different characterization techniques used for biosynthesized ZnONPs.
Characterization TechniquePrincipleAdvantageReference
UV-visible spectrophotometryMeasures absorbance of lightRapid and nondestructive[97]
X-ray diffraction (XRD)Measures crystal structure and sizeProvides detailed crystallographic information[62]
Scanning electron microscope (SEM)Provides surface morphology and sizeHigh resolution imaging[98]
Transmission electron microscope (TEM)Provides detailed information on size, shape and structureHigh resolution imaging and analysis of individual particles[62]
Fourier transform infrared spectroscopy (FTIR)Measures functional groups on the nanoparticle surfaceProvides information on surface chemistry[97]
Dynamic light scattering (DLS)Measures particle size distributionRapid and nondestructive[97]
Zeta potential analyzerMeasures the surface charge of particles in solutionProvides information on particle stability[97]
Table 4. Modes of action of green synthesized ZnONPs against pathogenic microbes.
Table 4. Modes of action of green synthesized ZnONPs against pathogenic microbes.
Treated Pathogenic MicrobesMode of ActionReferences
H. suis, H. felis, H. bizzozeronii, H. salomonisLead to the damage of cell wall, cell membrane and DNA, mitochondrial dysfunction, apoptosis, generation of reactive oxygen species and, finally, cell death.[15]
S. aureus, E. coli, B. subtilis, P. aeruginosa, C. albicans.Inhibit different metabolic functions including cell metabolisms, transportation, enzyme activity, etc.; generate reactive oxygen species and lead to the death of cell.[44]
S. aureus, E. coli, E. faecalis, S. enteritidis, K. pneumoniae, P. aeruginosa, A. baumannii, S. typhimurium, C. albicansDamage of cell membrane and DNA, leakage of intracellular molecules, denaturation of enzymes and proteins, inhibition of protein synthesis, generation of reactive oxygen species.[121]
Burkholderia glumae, B. gladioliDamage cell membrane, proteins, ribosome, and cytoplasmic materials; produce reactive oxygen species and cause leakage of genetic materials, resulting cell death.[42]
E. coli, Salmonella sp., S. aureus, S. epidermidisDamage the cell membrane, cause leakage of intracellular materials and generate reactive oxygen species, which lead to the death of the cell.[13]
S. aureus, K. pneumoniaeGeneration of reactive oxygen species, DNA damage, protein denaturation and mitochondrial dysfunction.[36]
P. aeruginosa, C. violaceumAttach to cell membrane, break membrane permeability, release Zn ions, generate reactive oxygen species.[14]
E. coli, P. putida, S. aureus, S. pyogenesInteract with cell membrane, produce reactive oxygen species, damage cell wall, DNA, protein and iron. [2]
B. flexus, B. filamentosus, A. baumannii, P. stutzeriDamage of cell wall, inhibition of cellular metabolism and respiration, destruction of DNA and inactivation of protein.[68]
C. albicansDisrupt and deform the cell wall and cell membrane and inhibit the production of ergosterol, which lead to cell death.[62]
S. pyogenes, P. aeruginosa, S. aureusProduction of significant oxygen reactive species including hydroxyl radicals, superoxides and hydrogen peroxide.[38]
F. oxysporum, S. aureus, P. aeruginosa, R. necatrix, E. coliDamage cell membrane, generate reactive oxygen species, damage DNA, denature protein, cause ribosomal destabilization and mitochondrial dysfunction, which lead to the death of cell.[23]
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

Huq, M.A.; Apu, M.A.I.; Ashrafudoulla, M.; Rahman, M.M.; Parvez, M.A.K.; Balusamy, S.R.; Akter, S.; Rahman, M.S. Bioactive ZnO Nanoparticles: Biosynthesis, Characterization and Potential Antimicrobial Applications. Pharmaceutics 2023, 15, 2634. https://0-doi-org.brum.beds.ac.uk/10.3390/pharmaceutics15112634

AMA Style

Huq MA, Apu MAI, Ashrafudoulla M, Rahman MM, Parvez MAK, Balusamy SR, Akter S, Rahman MS. Bioactive ZnO Nanoparticles: Biosynthesis, Characterization and Potential Antimicrobial Applications. Pharmaceutics. 2023; 15(11):2634. https://0-doi-org.brum.beds.ac.uk/10.3390/pharmaceutics15112634

Chicago/Turabian Style

Huq, Md. Amdadul, Md. Aminul Islam Apu, Md. Ashrafudoulla, Md. Mizanur Rahman, Md. Anowar Khasru Parvez, Sri Renukadevi Balusamy, Shahina Akter, and Md. Shahedur Rahman. 2023. "Bioactive ZnO Nanoparticles: Biosynthesis, Characterization and Potential Antimicrobial Applications" Pharmaceutics 15, no. 11: 2634. https://0-doi-org.brum.beds.ac.uk/10.3390/pharmaceutics15112634

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