Next Article in Journal
Two-Step Preparation of Protein-Decorated Biohybrid Quantum Dot Nanoparticles for Cellular Uptake
Previous Article in Journal
Targeting Peptides: The New Generation of Targeted Drug Delivery Systems
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Updates on Biogenic Metallic and Metal Oxide Nanoparticles: Therapy, Drug Delivery and Cytotoxicity

by
Maria P. Nikolova
1,*,
Payal B. Joshi
2 and
Murthy S. Chavali
3
1
Department of Material Science and Technology, University of Ruse “A. Kanchev”, 8 Studentska Str., 7017 Ruse, Bulgaria
2
Shefali Research Laboratories, 203/454, Sai Section, Ambernath (East), Mumbai 421501, Maharashtra, India
3
Office of the Dean (Research), Dr. Vishwanath Karad MIT World Peace University (MIT-WPU), Kothrud, Pune 411038, Maharashtra, India
*
Author to whom correspondence should be addressed.
Pharmaceutics 2023, 15(6), 1650; https://doi.org/10.3390/pharmaceutics15061650
Submission received: 21 April 2023 / Revised: 20 May 2023 / Accepted: 30 May 2023 / Published: 3 June 2023

Abstract

:
The ambition to combat the issues affecting the environment and human health triggers the development of biosynthesis that incorporates the production of natural compounds by living organisms via eco-friendly nano assembly. Biosynthesized nanoparticles (NPs) have various pharmaceutical applications, such as tumoricidal, anti-inflammatory, antimicrobials, antiviral, etc. When combined, bio-nanotechnology and drug delivery give rise to the development of various pharmaceutics with site-specific biomedical applications. In this review, we have attempted to summarize in brief the types of renewable biological systems used for the biosynthesis of metallic and metal oxide NPs and the vital contribution of biogenic NPs as pharmaceutics and drug carriers simultaneously. The biosystem used for nano assembly further affects the morphology, size, shape, and structure of the produced nanomaterial. The toxicity of the biogenic NPs, because of their pharmacokinetic behavior in vitro and in vivo, is also discussed, together with some recent achievements towards enhanced biocompatibility, bioavailability, and reduced side effects. Because of the large biodiversity, the potential biomedical application of metal NPs produced via natural extracts in biogenic nanomedicine is yet to be explored.

1. Introduction

Different prokaryotic and eukaryotic organisms, their primary or secondary metabolites, or biomolecules contain components capable of reducing metal salts into metallic or metallic oxide nanoparticles (NPs). For example, in plants and especially in their leaf extracts, there are plenty of phytochemicals, such as terpenoids, ketones, aldehydes, carboxylic acids, phenols, flavones, etc., which are thought to be involved in the synthesis of metal or metal oxide [1]. Renewable natural resources (such as plant extracts, microorganisms, algae, etc.) are used as biological precursors to synthesize NPs, avoiding the production of toxic and harmful by-products. Eco-friendly biosynthesis is useful for minimizing waste and reduction of pollution by utilizing sustainable and renewable feedstock. In contrast to chemical or physical methods that require highly toxic reductants or high radiation, as well as high temperature and pressure [2], biosynthesis utilizes low energy for the initiation the bio-reduction. Moreover, the implementation and use of toxic chemicals and solvents prevent the use of NPs in a range of biomedical and clinical applications [3].
The biogenic synthesis of NPs includes the adoption of unicellular and multicellular entities that operate as a pattern for the formation, assembly, and organization of nano-scaled materials or the assembly follows a bottom-up production route. Various metallic and metal oxide NPs have been synthesized from biogenic precursors by using different process parameters such as pH, pressure, temperature, solvents, etc. The term “biosynthesis” is usually used to denote the production of substances or compounds from simple precursors by living organisms, their metabolites or biomolecules. Ionic liquids are sometimes used instead of other solvents because they easily dissolve organic compounds, gases, or catalysts even though they have a polarity comparable to alcohol. They also operate in large temperature ranges and their solubility can be modified by ions associated with them [1]. Thus, scientists have utilized tissues/organs or whole organisms as the main bioreactors to execute the biosynthesis of metallic and metal oxide NPs with different properties. The production of NPs arising from the resistance mechanism of a certain living organism against a specific metal may alter the chemical nature of the toxic materials making them less- or non-toxic [4]. The natural synthesis can be a result of the bioreduction of the metal ions by enzymes that are oxidized or linking peptides that assemble into a more stable nanostructure [5]. This means that the well-known instability of NPs in solutions due to agglomeration and dissolution can also be overcome by the natural capping mechanism that usually occurs during biosynthesis. Simultaneously, by using different organisms or extracts, biomolecules, or bioactive compounds in the presence of metal salts, researchers can produce NPs of various shapes, sizes, compositions, and biological activities, such as antimicrobial, anticancer, larvicidal, and antioxidant ones [6]. The strong bounding between biological structures by their functional groups and nanoparticles can produce benign nanoscale materials with demanding sustainable advantages.
Nanotechnology plays a major role in today’s biomedicine. Nanoparticles display distinction over classical therapy methods in terms of effectiveness and competence. Drugs or other active compounds can be loaded in bioengineered NPs for effective transfer to a certain site in a living organism. On the one hand, noble metal NPs such as gold (Au), silver (Ag), palladium (Pd), platinum (Pt), other metals like copper (Cu and its oxide CuO), and some other oxides (zinc oxide, titanium oxide, iron oxides, etc.) are characterized by their outstanding physical-chemical, optical, magnetic, and biological properties [7]. On the other, except in allopathic medicines, worthy pharmaceutical compounds as key sources of antioxidants, antimicrobial, and cytotoxic moieties have been produced by using biomaterials from plants, herbs, different species, and macromolecules (Figure 1) with fewer side effects [1]. Natural materials from renewable energy sources also use safe solvents and reactants and minimize waste products that have health, social, and environmental benefits. Thus, biogenic NPs for biomedical applications have been engineered so that they have improved bioactive performance by complimenting the efficacy of NPs with that of the capping biological agent or are specifically directed towards the diseased cells while decreasing the side effects. A considerable advance has been achieved in many biogenic nanoengineering processes for biomaterial production. The commercialization of successful cost-effective and eco-friendly technologies is of utmost importance for humans but a critical view is necessary for the potential side effects and toxicity of these materials. For that reason, in this review, we have attempted to summarize the biogenic principles of bio-assembly and the construction of nanosystems with biomedical applications. Since many nano drugs rely on metal and metal oxide NPs, we have focused on the biosynthesis of these nanomaterials for therapeutics by giving a background on bio-nanotechnology prospective evolution. Furthermore, some significant applications of biogenic NPs such as cancer therapy and the delivery of antitumor drugs, antimicrobial/antifungal therapy, anti-inflammatory, wound healing, osteoinduction, anti-viral, and antiparasitic functions, as well as their nanotoxicity, have also been discussed.

2. Characteristics of Some Biosynthesized NPs

Nanoparticles have been known to be utilized for a great number of biomedical and pharmaceutical applications. When NPs are used, there is a huge increase in surface area available for reactions and hence an increased effect is observed. Nanomaterials have not only a large specific surface area but also a high surface area to volume ratio which increase with the decrease in size, morphology, and distribution of NPs [7]. Also, the electronic configurations of metal NPs are such that they allow the acceptation or donation of an electron to quench free radicals [8]. Metallic NPs can adopt various physical geometries and electronic assemblies that allow them to exhibit metallic, semiconducting, or insulating characteristics.
NPs have been used in various biological and pharmaceutical applications because of their small size (1–100 nm) and similarity to cellular components that allow them to enter living cells by using numerous cellular mechanisms. Nanoparticles provide great advantages in the pharmaceutical industry as drug carriers due to their protection of drugs in vivo, increased and sustained drug activity, improved delivery efficiency and selectivity, and extended release profile [9]. The nano size of the particles also enlarges the penetration potential thus aiding in better utilization of NP properties. Thus, because of the higher site-specific delivery of drugs, the required dose of drugs and side effects will be reduced substantially. In nanosized form, NPs can penetrate the circulatory system and translocate the blood-brain barrier. Such NPs interact in different ways with a large range of biomolecules thus directing various cellular, physicochemical, and biochemical properties [10]. Most biological entities assist stabilization and act as capping agents protecting from coalescence and aggregation. Agglomeration is a phenomenon where the NPs lower the surface energy resulting in a reduction in surface area by increasing the particle size. The biopolymers enhance the biocompatibility of these NPs and prevent agglomeration in clusters that affect their bio-dispersity [11]. The stability of a NP can modulate the biological response by changing the bioavailability, cellular uptake, and toxicity in vitro and in vivo [11]. Factors that manipulate the stability of NPs are based on composition, surface properties, and aggregation state. The combination of molecules from a biological origin that participate in the reduction and capping of NPs promotes their stability, inhibits agglomeration, and impacts their therapeutic effect [7].
Historically, the process of bio-nano-material fabrication dates back to 1980, when Beveridge et al. synthesized gold NPs by using Bacillus subtilis [12]. Today, the most commonly examined metallic NPs are Au, Ag, Pt, Pd, Cu, ZnO, and TiO2 [13].

2.1. Gold NPs

Au NPs are unique with enhanced active properties, unlike their bulk noble counterparts. Au NPs were extensively used in cancer therapy, drug delivery, imaging, and other biomedical applications such as antibacterial [14,15]. They demonstrate outstanding surface plasmon resonance and can be combined with different biological assemblies, such as oligosaccharides, and proteins to enhance their functions. In the typical chemical synthesis of gold nanoparticles, chloroauric acid is reduced using NaBH4 or sodium citrate as reducing agents, and, at times, seeded growth control is performed to maintain their uniform nanoparticle morphologies. With the use of Au NPs in medicine, these chemical protocols seem detrimental and toxic to humans. Numerous efforts are taken in preparing hybrid AuNPs through chemical processes that involve drug–gold complexation and require surfactants and reducing agents. Various organisms like plants (Lonicara japonica) [16], bacteria (Delftia acidovorance) [17], fungi (Aspergillum sp.) [18], algae [19], etc. are used for the biosynthesis of Au NPs. The production is usually initiated by tetrachloroaurate salt (HAuCl4) while the most common shape is spherical, hexagons, and triangles. The preparation of hybrid AuNPs utilizing biomolecules such as proteins [20] and antibodies [21] is experimented with and has shown efficacy in cancer therapy. Choosing the right materials for biogenesis with a high abundance of reactive compounds, the reduction and capping properties of biocomponents assure stability for different biomedical applications.

2.2. Silver NPs

Silver NPs are massively used in many biomedical applications. Ag NPs have been widely utilized in pharmaceutics, medical implant coatings, and wound dressing because of their antimicrobial, anti-inflammatory, and antioxidant properties [22,23]. Silver was found to be the most frequently used nanomaterial (435 products) [24]. Ag NPs are characterized by high sensitivity, conductivity, and chemical stability. Ag NPs exhibit antitumor activity by reducing cell proliferation and promoting intracellular ROS, DNA damage, and apoptosis [25]. It is also well known that silver is highly toxic to microorganisms, including 16 major species of bacteria [26]. Nonetheless, some microorganisms may survive and grow under certain metal ion concentrations due to their resistance to silver. It is thought that the main mechanisms triggering resistance involve the presence of nitrate reductase enzyme [27]. The synthesis of Ag NPs through plant extracts (Madhuca longifolia) [28], bacteria (Bacillus subtilis) [29], fungi (Beauveria bassiana) [30], and algae (Botryococcus braunii) [31] with different sizes and shapes have been reported. AgNO3 is usually used as precursor salt for biomimetic synthesis while the obtained NPs commonly have spherical, triangular, or hexagonal shapes. The biosynthesis of Ag NPs enhances their stability and may reduce their toxicity. However, despite many therapeutical and medical benefits, depending on the size, shape, and capping agents there may be a problem with the nanotoxicity of silver NPs in humans during long exposition time which will be discussed later in this study.

2.3. Platinum and Palladium NPs

Platinum (Pt) and palladium (Pd) NPs have exhibited potential therapeutic effects apart from their excellent catalytic prowess in chemical reactions. Both NPs were exploited in antibacterial and biomedical applications [32]. Few studies revealed that Pt NPs can be biosynthesized from cyanobacteria [33], seaweeds [34], tea extracts [35], honey [36], and eggs [37]. Platinum salts such as H2PtCl6, K2PtCl6, K2PtCl4, PtCl2, Pt(AcAc)2, Pt(NH3)4(OH)2, Pt(NH3)4(NO3)2, and Pt(NH3)4Cl2 are applied for biosynthesis [38]. Although the biomimetic synthesis of Pt NPs is limited, a concise review to understand its mechanism and factors that influence NP morphologies has been conducted [39].
Among nanoparticles for biomedical use, Pd is among the least examined. Pd is a high-density metal. Pd NPs have significant thermal and chemical stability and can be biofunctionalized to become suitable for biomedical applications. It also possesses antimicrobial, antioxidant and cytotoxic activity [40]. The biosynthesis of palladium (Pd) NPs is majorly reported from plant-based extracts such as leaf extracts [41], peel extracts [42], bark extracts [43], fruit extracts [44], root extracts [45] and plant gums [46]. For biosynthesis, K2PdCl4 solution is usually used. Natural antioxidants, such as monosaccharides, vitamin C, and gallic acid, have been used as reducing agents in Pd NP biosynthesis [47].

2.4. Copper and Copper Oxide NPs

Cu NPs are widely utilized as cheap and effective bactericidal agents largely applied in medical treatments [48]. Nowadays, as a promising contender at lower cost, Cu NPs are taking the place of Au and Ag NPs but Cu NPs are highly oxidant in air and water. Another challenging job is to stabilize the NPs after synthesis. The main advantage of using a biogenic route for the production of Cu NPs is stabilization [49]. For example, capping agents in the phytosynthesis of Cu NPs help them to stabilize for more than 30 days in contrast to chemically produced ones that settle down after 24 h [50]. Biosynthesis involves the use of algae [51], sea cucumber [52], microorganisms [53], and plants [54]. According to Ying et al., since copper is a transition metal, usually Cu NPs cannot be directly obtained from simple copper salt [55], while in biosynthesis cupric acetate (monohydrate (CH3COO)2Cu.H2O) [56] or CuSO4 [49] are often used for the direct bio-production.
After oxidation, copper oxide (CuO) can be formed. It is a p-type semiconductor compound with a monoclinic structure. CuO NPs are characterized by their antimicrobial properties and ability to easily cross biological barriers to reach target organs according to their size and surface properties [57]. Biosynthesized CuO NPs have been prepared from leaf extract of Eucaliptus globulos [58], mint leaves and orange peels [59], algae (Anabaena cylidrica) [60], and bacteria (Serratia sp.) [61], usually by using copper sulfate salt precursor. The biosynthesized CuO NPs are thought to have increased antimicrobial activity compared to commercial NPs due to the participation of natural extracts used for reduction [3]. However, CuO NPs were found to be highly toxic compared to other metal oxide nanomaterials [62].

2.5. Zinc Oxide NPs

ZnO NPs are multifunctional n-type semiconductors with high selectivity and low toxicity, making them suitable for drug delivery and targeted therapy [63]. ZnO NPs are considered one of the most promising antibacterial agents because of their low cytotoxicity, compatibility, and good heat resistance [63]. Chemically, the surface of ZnO NPs is rich in -OH groups allowing ZnO to dissolve slowly in both acidic (as in tumor cell microenvironment) and strong basic conditions. For that reason, in medicine, it is used as an antiplatelet agent, anti-inflammatory, anti-angiogenesis, gene- and drug delivery, dental material, and anti-cancer agent [64]. Due to its good antimicrobial and disinfectant properties, ZnO has found widespread utilization in various dermatological substances [65]. However, the application of ZnO NPs can be limited because of the wide energy bind gap (3.37 eV) and the high complexation of photogenerated electron-hole pairs [66]. Fortunately, doping with other metal ions such as Au, Ag, Pd, etc., or compounding/capping with other non-metal materials such as chitosan [67] can reduce the wide bind gap and strengthen the antibacterial effect. For its biosynthesis, plants (Mentha pulegium) [68], bacteria (S. aureus) [69], fungi (Xylaria acuta) [70], and algae (Sargassum multicum) [71] have been used. The precursor salt solution can be zinc nitrate (Zn(NO3)2·2H2O) [67] or zinc acetate dihydrate (Zn(C2H3O2)2·2H2O) [72].

2.6. Other Metal Oxide NPs

Iron oxide NPs are commonly used in target drug delivery, MRI, diagnosis of cancer, and tissue engineering [73]. As a result of Neelian and Browning relaxations, magnetite (Fe2O3) and maghemite (γ-Fe2O3) NPs respond to an external magnetic field and can be naturally directed toward magnetic targeting to remotely control the distribution of drug molecules. As magnetic material, iron oxide NPs exposed to alternating magnetic fields can convert electromagnetic energy into heat, which makes them suitable agents for hyperthermia therapy of cancer. The magnetic properties of iron NPs can be tuned by changing the size and shape of the nanocrystals, inducing core-shell coupling, doping with other metals, or forming nanoclusters by crosslinking or encapsulation [74]. The biosynthesis of Fe3O4 NPs allows the production of particles with a size of 2–80 nm, which is much smaller than the 87–400 nm size of the same wet chemically synthesized Fe3O4 particles [75].
The three main polymorphs of titanium dioxide (TiO2) NPs are anatase, rutile, and brookite. Recognition of TiO2 NPs for their unique thermal, electric, optic, catalytic, magnetic, and antimicrobial properties for various applications was found in the literature. Titanium dioxide NPs have the potential to induce cell death since they form reactive oxygen species when their aqueous solution is illuminated with visible or UV light [76]. This property utilized for treating cancer is called photodynamic therapy. The biosynthesized TiO2 NPs are ecologically friendly and demonstrate high oxidizing potential [77]. When comparing the anticancer activity of plain and biomodified TiO2 NPs, the superior performance of the biogenic NPs was also demonstrated [78]. Moreover, plant-mediated production of TiO2 NPs resulted in a synthesis of nanomaterial, having functional groups such as tannins and phenols, which help in the stabilization of TiO2 NPs and their increased antioxidant potential [79].
Cerium oxide (CeO2) NPs have also received much attention because of their improved redox properties as opposed to their bulk counterpart [80]. In the biomedical field, CeO2 NPs are mostly used as therapeutic agents, drug-delivery carriers, and antioxidant, antimicrobial, and antiparasitic ointments because of their unique surface properties and their chemistry, biocompatibility, and high stability [81]. The presence of a mixed valence state (both Ce3+ and Ce4+) and quick transition of the oxidation states play an important role in the scavenging reaction of oxygen while the surface Ce3+:Ce4+ ratio influences biological interactions [80]. Biogenic synthesis of CeO2 NPs has been reported by using plant extracts [82], bacteria [83], fungi [84], algae [85], or biological products [86]. Biogenic CeO2 NPs have shown promising results in treating drug-resistant pathogens and fungi, as well as cancers such as colon, cervical, breast, and osteosarcoma ones [81].

3. Biogenic Metallic NPs Production

To biosynthesize metallic NPs, the living matter uses dissimilatory metal reduction at the expense of the oxidation of enzymes. The different compounds such as terpenoids, flavonoids, alkaloids, glycosides, proteins, carbohydrates, vitamins, polymers, and antioxidants act as reducing and capping/stabilizing agents in the synthesis of sustainable NPs [1]. Figure 2 illustrates the mechanisms, participating components and the affecting factors of the biogenic fabrication of NPs using plants and microorganisms. Despite being sustainable and environmentally friendly, the biosynthesis could be sluggish and time-consuming while the synthesized NPs may be not monodispersed [87]. For that reason, optimization of the control factors and careful selection of the organisms may allow the implementation of such methods in large-scale production. After optimization, the time for biogenic production could be much shorter than physicochemical synthesis. For example, by using brown alga Padina pavonia, Abdel-Raouf et al. were able to synthesize silver NPs within 2 min [88]. Adding 200 mg of P. pavonia ethanolic or chloroform extract the color changed from colorless to brownish black within 2 min suggesting the formation of silver NPs whose sizes ranged from 49.6 to 86.4 nm. Moreover, to activate the process, a fusion of biogenic methods with alternative routes such as ultrasound and microwave was also applied [89]. Microwave heating, for example, provides fast primary heating and enhanced reaction kinetic thus boosting the reaction rate and increasing the yields [6].
The biosynthesis process offers simple and eco-friendly production, and by controlling the process, particles with desired geometries and composition can be formed. Various NPs with different morphology, such as square, rectangular, triangular, polygonal, spherical, cylindrical, flower-like, etc., can be obtained (Figure 3). Simultaneously, the biogenic synthesis could face challenges in producing monodisperse NPs with fine tune particle size distribution [90]. Additionally, some nanoparticles produced by microorganisms can be less stable as opposed to those synthesized by chemical processes [91]. However, with the proper selection of the best biological candidates based on their inherited properties of growth rate and biochemical activities as well as parameters such as pH, temperature, process time, and reagent concentration, the biogenic processes can be improved with an attempt to obtain high quality and yield of biosynthesized NPs at the expanse of lower cost because of lack of organic solvents, thermal stabilizers, and expensive production techniques.

3.1. Biosynthesized NPs by Using Plants and Their Extracts

At a large scale, the synthesis methods utilizing plant extracts are comparatively easy, simple, and cost-effective as opposed to fungi or bacteria-mediated production and they present feasible methods and alternatives to conventional NP production methods. Because of the broad availability of biologically active plant extracts and their biodegradability, the biosynthesized NPs from plant extracts are receiving great interest. Plant materials like leaves, flowers, fruits, roots, seeds, etc. are available at all times and seasons at large volumes for large-scale synthesis while using them will not affect crop productivity.
Plants are known to possess the potential to accumulate some quantity of heavy metals at various parts, which makes them suitable for in vivo synthesis of metal NPs by absorbing soluble salts. The steps involved in NPs production include bio-reduction and nucleation of reduced ions, growth, and termination including the formation of the final shape of the nanomaterial [5]. Although the mechanisms of plant-mediated synthesis of NPs are still under continual research, it is known that organic matter plays a vital role in crystal growth and forming certain particle sizes [92]. Negatively charged groups concentrate positive ions from the solution, thus contributing to ion saturation to initiate the nucleation. The main functional groups involved in the reduction of metal ions are hydroxyl, amino, carbonyl, and methoxide, which electrostatically interact with the ions and lead to their reduction [93]. The ionic forms of metal can be easily detached from anionic parts because of the reduction process that renders them stable in the presence of plant extracts [94,95].
The major part of the research focused on the ex vivo production of NPs from plant extracts. The extracellular synthesis is advantageous because it excludes the presence of intracellular proteins and additional treatments with chemicals or ultrasonic cleaning. The synthesis parameters such as pH, temperature, salt concentration, etc. control the yield, size, rate of formation, and NPs’ stability [96]. For instance, the many metal ions absorbed on the surface of preformed nuclei will lead to secondary reduction and enlargement of NPs [97]. As the reaction temperature increases, the reduction rate rises and many metal ions are consumed for NPs formation, thus blocking the secondary reduction process on the surface of the preformed nuclei, which leads to small and highly dispersed NPs formations with increased yield [98]. Simultaneously, with the increase in reaction time the size of the metal NPs also increases due to aggregation. Under different pH conditions, desired size and shape uniformity can be obtained [99]. The growth rate of NPs increases with the increase of the reducing agents, but too many reducing agents may trigger a bridging effect among the produced NPs and their aggregation [100]. Additionally, the composition of the plant extract that contains a different concentration of phytochemicals is also an important factor in the bio-assembly. The phytochemicals responsible for the reduction and cupping of NPs are sugars, ketones, carboxylic acids, terpenoids, flavones, aldehydes, amides, alkaloids, etc. These constituents effectively activate the reduction mechanisms and reduce and stabilize the produced NPs. Some plants, such as cruciferous vegetable extracts, are also capable of the synthesis of bimetallic NPs in the form of core-shell structures of Au core and Ag shell [101].
Plants seem to be superior candidates for large-scale biomimetic synthesis of NPs because of the fast rate of formation and the variety in shape and size of NPs. Table 1 demonstrates some recent achievements of plant-mediated NP synthesis together with the particle characteristics and their application. It can be summarized that biosynthesis is simple, relatively rapid, and cost-effective. When used in the nanobiotechnology field, many plants containing pharmacologically active substances can support the action of treatment. However, when using plants, it is hard to maintain monodisperse in the NP population as well as reproducibility of the process [102]. Because of the presence of many organic compounds in the plant extracts, it may also be difficult to identify the exact reactive components. Another issue in photosynthesis is the use of plants of commercial value as reducing and stabilizing agents, which affects the efficacy of the synthetic procedure [103]. Challenges associated with unoptimized reaction process parameters and unexplored qualitative growth kinetics of NPs are also obstacles for the successful commercialization of phytogenic processes [103].

3.2. Biosynthesized NPs by Using Microorganisms

3.2.1. Bacterial Synthesis

Bacteria are prokaryotic microorganisms. Metal-reducing procaryotic bacteria and actinomyces have been broadly used for the production of metal and metal oxide NPs because bacteria are easily genetically manipulated [127]. Bacteria are extremely adaptable, fast-growing, and have environmental abundance. Bacteria are also capable of surviving various stressful conditions, including the presence of higher concentrations of toxic metals [128]. When exposed to high concentrations of heavy metal ions, they use different defense mechanisms to overcome the stress conditions. Depending on the location of NP formation, the biogenesis of NPs can be either extracellular or intracellular by encapsulation in the cytoplasm. The mechanism behind the extracellular synthesis of NPs includes the aggregation of metal ions on the cell surface and the involvement of enzymes. The negatively charged polysaccharides containing phosphoric and carboxylic groups can bind to the positively charged metal/metal oxide ions via electrostatic interaction [129]. Moreover, because of the presence of negatively-charged amino acids, such as aspartic acid and glutamic acid, various enzymes and peptides are essential for the reduction of metal ions to NPs. The main enzymes are oxidoreductases, such as nitrate reductase, sulfate reductase, and cellular transporters [130,131,132]. For example, after the initial electrostatic interaction of the metallic ions with the cell wall of the bacteria, reduction and the bio-reduced metallic nuclei growth can be assisted by NADH (nicotinamide adenine dinucleotide) and NADH-dependent nitrate reductase that transferred electrons into NPs [130]. During the enzyme-assisted reduction, various amino-acid residuals can link to the metal or metal oxide ions and then reduce them to metal or metal oxide NPs [131]. Metal-reducing strains such as Shewanella have evolved a mechanism for electron transfer across the cell membrane via periplasm to the surface that involves c-type cytochromes from quinol at the inner membrane to metal oxide surface (Fe(III) oxide) at the outer membrane [132]. Other authors found that hemiacetal and aldehyde groups of exopolysaccharides in E. coli were involved as bio-reducing agents in Ag NPs production [133]. After contact with exopolysaccharides, metal ions are chelated and after that reduced and capped by the various groups of reducing sugars. Although actinomycetes were found to synthesize NPs via both intracellular and extracellular mechanisms, extracellular biomineralization was commonly used for commercial applications [13]. Among them, Streptomyces sp. is most widely utilized in pharmaceutical applications because 55% of the known antibiotics are produced by them.
Some magnetotactic bacteria have been used for the production of magnetosomes with magnetic radicals with controlled geometry and composition of the NPs [134]. The magnetosomes are organic-coated nanocrystals usually containing iron oxide NPs with an organic layer of bacterial phospholipid membrane that can be applied in cancer therapy, molecular imaging [135], or carrying drugs or other loads in cancer treatment [136].
The advantages of the bacterial-synthesized NPs are their production from impure materials and other feedstocks at ambient conditions and mild temperatures [137]. However, compared with plant synthesis, the bioproduction of NPs using microorganisms can be in some cases pathogenic to humans, and it requires precise control over cultivating and maintaining the microbial cells. Other main issues are the need for a mechanical breakdown of the cell wall, centrifugation, purification of NPs, and difficulty in controlling their geometry [138]. Compared with fungi, bacteria produce a much smaller volume of NPs because of their lower protein production [87]. The bacteria culturing is also tedious, the process conditions for bacteria growth are inflexible, and the reduction process is slow, lasting from hours to days [139].

3.2.2. Fungal Synthesis

Fungi are easier to culture in laboratories and industries than microorganisms. Fungi are found to be efficient in the biogenesis of monodisperse NPs with well-defined morphology. Because of the existence of a large number of intracellular enzymes, they are thought to be more effective microorganisms in the production of metallic NPs compared to bacteria [140]. Fungi can secrete a large amount of red-ox active extracellular proteins, thus increasing the synthesis of NPs forming insoluble metal-protein conjugates. They were also found to have higher productivity and higher tolerance to metals because of the higher cell wall binding capacity of these ions with biomass and higher bioaccumulation ability [141]. Additionally, fungi possess reducing components, proteins, and enzymes (such as reductase) on their cell surface [142]. The polysaccharide chitin, a key ingredient in the fungal cell wall, was also found to be involved in heavy metal complexation and biogenic NP production [143].
Fungi can produce NPs intra- and extra-cellularly. For example, Fusarium oxysporum was found to produce Ag NPs by extracellular mechanisms with the action of NADPH nitrate reductase and anthraquinones [144]. The reduction of the metal ions to NPs is accompanied by the oxidation of NADPH to NADP+, whereas naphthoquinones and anthraquinones facilitate the reduction by a quinone-based shuttle [145]. Alternatively, NPs or metal ions may diffuse through the cell membrane and be reduced by red-ox systems in the cytoplasm. The change in pH in the medium can trigger the formation of NPs with different sizes since the changes in pH influence the acidic or basic nature of amino acids that participate in biogenic synthesis [146]. The incubation temperature, salt concentration, and incubation time also affected the size and rate of the extracellular bio-fabrication of Au NPs [147].
Extracellular synthesis is cheaper and simpler, while intracellular production provides NPs that should be additionally purified. Many fungi form mycelia, which provide higher surface area than bacteria for supporting fungal biomass-ions interactions. Filamentous fungi are preferred to bacteria because they are highly tolerant to metals, capable of extracellular NP formation, and easy to handle [148]. Fungal micelles can withstand pressure, agitation, flow, or other conditions compared to plant materials or bacteria. However, similarly to bacteria, the major drawback of fungi-produced NPs relates to biosafety and low rate for synthesis in comparison with plant extract. Some species, such as F. oxysporum, are pathogenic, which makes them a health concern, while others, such as Trichoderma fungus, are vastly utilized in food and medical application [149]. Attention should also be paid to the production of mycelia-free filtrates.

3.2.3. Yeast Synthesis

These single-cell eukaryotes have been reported to be used in the successful biosynthesis of various metallic NPs because they possess various detoxication mechanisms. Yeasts are classified in the kingdom of fungi and, similarly to them, have an envelope composed of chitin, glycoproteins, and β-glucans that may participate in extracellular NP biogenesis. Intracellular reduction of metal ions occurs after the passive diffusion of metal salts followed by reduction mediated by the transport of reductive agents into the cell [150]. More focus is put on extracellular synthesis, since additional operations like ultrasound processing and chemical treatment are eliminated.
Silver-tolerant species such as Saccharomyces cerevisiae were usually used for the synthesis of Au [151] and Ag [152] NPs, but, recently, some other species have been adopted as biogenic NP producers. For example, Zhang et al. demonstrated a biosynthesis of Au NPs with hexagonal, spherical, and triangular morphology through the yeast Magnusiomyces ingens LH-F1 [153]. The genetically engineered yeast strain Pichia pastoris has also been used for the biogenic fabrication of Ag NPs [154]. The yeast strain was transformed with an upregulated metal-resistant gene from Mucor racemosus produced cytochrome b5 reductase enzyme. Thus, the yeasts were capable to produce stable Ag NPs with a size of 70–180 nm. Not only commonly produced Ag NPs but also Pd NPs with hexagonal form and a size of 32 nm were synthesized by Saccharomyces cerevisiae [155]. The yeast Saccharomyces cerevisiae and its three mutants with deleted genes were also used for the production of ZnO NPs with an average size of 20 nm [156]. Likewise, anatase TiO2 NPs with an average size of less than 12 nm were biosynthesized through incubation with Baker’s yeast [157]. They showed a significant antibacterial effect on selected pathogens especially against Gram-positive bacteria in the presence or absence of sunlight exposure. Easy control of mass production and rapid growth of yeasts make them preferred microorganisms for NP biosynthesis.
A schematic presentation of the synthesis process by biogenic routes that use microorganisms is shown in Figure 4.
The biogenetic synthesis from microbes including bacteria, yeast, and fungi may suffer from complications related to extraction, purification, and cell separation from the surface of NPs, which may cause serious adverse effects on living cells in the biomedical application [158]. Extra steps such as detergent treatment or ultrasonication, which further increase the price of production, are mandatory for the separation of NPs in the intracellular biogenesis route [159]. Additional limitations can be operational impediments such as (1) longer time (up to 120 h) for the reduction reactions; (2) hard retrieval of NPs from biomass; (3) selection of the best organism for the biosynthesis and biocatalyst state; and (4) picking the optimal conditions for cell growth and enzyme activity [103]. Additionally, problems related to the commercialization of the biosynthesis protocol at an industrial level are the main expected obstacles [103].

3.3. Virus Synthesis

Viruses are unicellular organisms that use the replication machinery of the host. Their structure consists of nucleic acid (DNA or RNA) surrounded by a protein shell, and/or lipid envelope. Viruses present in host cells can also be used to generate biogenicNPs. They are considered valuable biogenic resources for the synthesis of NPs in organized assemblies including several morphologies. The outer capsid proteins of the viruses offer a very reactive surface for interaction with metallic ions [13]. Hydroxyl groups in tyrosine and carboxyl groups in asparagine and glutamine or indole groups in tryptophan may participate in the reduction of the metallic ions. The sequence configuration has the potential to attract metal ions and controls the size and morphology of NPs [160]. Pathogens such as animal and plant viruses and bacteriophages have been used for nanobiotechnological purposes because of their structural and chemical stability, lack of toxicity and pathogenicity in humans, and ease of production [158]. Plant viruses are easy to cultivate and non-pathogenic to humus and, therefore, can be used to produce NPs with the nanomedical application. For example, a high yield of Au and Ag NPs with small sizes was reported to be biosynthesized when a low concentration of Tobacco mosaic virus (TMV) was added to Hordeum vulgare (barley) or Nicotiana benthamiana plant extract as opposed to those without viral particles [161]. Recently, Thangavelu et al. produced Au-Ag composite NPs through plant pathogenic squash leaf curl China virus [162]. The hybrid NPs biosynthesized through virus nanobiotemplate (32 nm) showed good biocompatibility. Bacteriophage-inspired biogenesis of Au NPs through a rare bacteriophage (Podoviridae family, C3 morphotype), also used as a reducing agent, revealed the presence of NPs in the form of spheres, hexagons, triangles, rhomboids, and rectangles, with a size ranging from 20 to 100 nm [163]. Viruses can be used for the production of nanocomposites with metal NPs useful for target drug delivery and cancer therapy [164]. For example, retargeted adenoviral vectors covalently bound to AuNPs have been also used as selective agents for the delivery of NPs to tumor cells [165]. However, the involvement of host organisms for protein synthesis and a small amount of research on large-scale applications still limit NPs’ production by viruses. Moreover, a few viruses displaying nucleation properties have been recently used as bio-templates for NPs synthesis [162]. Selecting the optimal process conditions for virus template synthesis and identifying the types of amino acids involved in the surface mineralization process are also crucial for the optimization of nucleation and growth of NPs [162].

3.4. Algae Synthesis

Algae are aquatic eukaryotic photosynthetic organisms that lack the multicellular structure of the plant. They can vary in size from microalgae to giant macroalgae. Microalgae are known to transform metal ions into NPs in a biogenic way, including active compounds in the cell wall such as laminarin that contain reactive groups [166]. Cyanobacteria strains grow faster than plants and are suitable for the eco-friendly and time-saving synthesis of NPs at ambient temperature. The interaction between Ag-containing solution and biomass allows the formation of Ag NPs without degradation of the biomass [167]. Microalgae-like diatoms (D. gallica and N. atomus) were also capable of producing Au NPs and Au-silica NPs [168]. It was also discovered that the formulation of Ag NPs includes extracellular compounds (polysaccharides) participating in the cell-free culture media [169]. Al-Katib et al. [170] found that the presence of specific functional groups in proteins is important in cupping and stabilizing during the extracellular synthesis of Ag NPs in Gloeocapsa sp. In addition to microalgae, biosynthesis of ZnO NPs by marine brown seaweeds (Turbinaria conoidea and Padina tetrastomatica) [171] was also reported. Tetraselmus kochinensis has also been shown to produce intracellular Au NPs [172]. The marine algae are found to be suitable for many biomedical applications, such as anticancer, antiviral, antioxidant, etc. agents, because the biosynthesized NPs contain different bioactive compounds and secondary metabolites [149]. Algae are characterized by easy culturing, high growth rate and biomass productivity compared to other microorganisms, which makes the biogenic algae-mediated process cost-effective. However, similar to microorganisms, downstream processing for the extraction of intracellularly produced NPs is still challenging while extracellular biogenesis is greatly influenced by cultivation conditions [173]. Various challenges, such as strain selection, low reaction rate and yield, size control and monodispersity, and cytotoxicity to non-target cells, still limit the scaling up of the bio-process to a commercial scale [174]. Limited information is also available on NPs biogenesis using microalgae, which hinders their applicability [175].

3.5. Biosynthesized NPs by Using Enzymes and Biomolecules

The synthesis of NPs using fungi or microbial systems is a slow process with a low yield that may result in the production of polydisperse NPs. For that reason, different enzymes and their metabolites such as polysaccharides, peptide chains, carbohydrates, nucleic acids, etc., are utilized as reducing and capping agents for the production of metal/metal oxide ions into NPs. Due to the excellent biocompatibility and a great number of functional groups, biopolymers (both native and modified) are vastly explored for the biogenesis of metallic NPs. There are three main approaches for the synthesis of biopolymer-supported NPs: (1) impregnation of the polymer with ion salt followed by its in situ reduction by hydroxyl groups of biopolymers; (2) impregnation with ion salt and addition of external reducing agent; and (3) synthesis of colloidal NPs and adoption into biopolymer branches [176]. The first one is the most convenient and frequently used. Several natural compounds such as cellulose [177], chitosan [178], and silk fibers [179] have been reported as reducing agents of Au3+ utilized for the biosynthesis and stabilization of Au NPs. The biogenesis of TiO2 NPs using starch [180], cellulose fibers [181], albumen [182], and lysozymes [183] has been reported. Although different polysaccharides, such as cellulose, chitosan, dextran, and starch, have been used for the production of metallic NPs, they require additional stabilizing. In contrast, pullulan was found to be effective in the reduction and stabilization of Au NPs produced from HAuCl4 solution [184]. An interesting study by Safat et al. demonstrated cerium oxide NP synthesis obtained from marine oyster extract used as a bio-reducing and capping/stabilizing agent. The marine biosynthesized NPs with a size of 15 ± 1 nm had no cytotoxic effect on normal cells [185]. Therefore, a great variety of polymers with several functional groups can produce NPs. However, all these polymers should be precisely examined for their reducing activity, efficacy, and bio-toxicity. The difference in concentration of active molecules because of external factor changes can affect the biosynthesis procedure [186].
With the help of commercially available enzymes in pure form, the production and purification of biogenic NPs can be eased. For example, Maddinedi et al. proposed the use of a natural enzyme, diastase, to synthesize Au NPs with a spherical shape and size of 9.7 nm [187]. With the increasing volume of enzymes, smaller NPs were produced. Similarly, an enzyme, Ia bacteriocin, belonging to nisin peptides, was also used for the biosynthesis of Au NP with a small size (25 nm) [188]. The enzyme consisting of 34 amino acids was obtained from Lactococcus lactis susbp. Arib et al. also reported a one-step enzyme-based synthesis of hybrid Au NPs using manganese superoxide dismutase protein in the presence of a zwitterionic sulphonic acid-based buffer as a reducing agent [189]. It follows that the enzymatic approach of NP synthesis proposes superior characteristics and easy purification of the obtained NPs but it is more expensive and time-consuming.
Table 2 sums up the biosynthesized NPs of various sources together with their characteristics and application.

4. Therapy and Drug Delivery of Biosynthesized NPs

Green chemistry focuses on the development of biogenic NPs that provide a reduction of harmful and toxic moieties of physiochemical processes and enables lower dose prescription of drugs during treatment. On the one hand, biogenic NPs themselves showed remarkable properties as antimicrobial and anticancer compounds [190,202,211,214]. However, when loaded with drugs, their efficiency can be additionally enhanced. On the other hand, to achieve target delivery and seize the rapid degradation of drugs, the concept of controlled drug-delivery systems (DDS) that overcome both physicochemical and biological barriers has been developed [223]. Because of the large surface area, NPs have drawn significant attention as potential DDS that can bind, adsorb, or entrap different drug molecules. The drug intended for release is usually directly bound to the nanocarrier and the time for release is of prime importance because it should not dissociate before reaching the target. The small NPs can pass through the extracellular matrix and assess layers that normal drugs are unable to reach [223]. Because of their unique chemical and physical properties that completely differ from their bulk counterparts, using NPs as potent DDS can improve the loading capacity and resides in sustained drug release, higher bioavailability, and enhanced target intracellular penetration [224]. It was also found that nanocarriers containing natural compounds may be useful for delaying the development of drug resistance and stimulating the low response to treatment with conventional medicine [225].
The targeting can be active (when peptides or antibodies are coupled with drug-delivery vectors to link to the receptor structure) or passive (when circulating in the bloodstream the nanosystem is passively attracted by changing factors such as temperature, pH, molecular folding, etc.). Therefore, the drug release can follow the stimuli-controlled release, chemical or physical reactions, etc. [226]. The selection of a particular metallic nano-vector is based on the physicochemical features of the drugs. Moreover, natural materials can provide different solutions for improving drug loading challenges. By presenting different active groups, chemicals such as covalent or hydrogen bonds, or physical interactions, such as electrostatic, may take place between drug polarities and capping or nanomaterials. Some drugs themselves, such as vancomycin and doxorubicin, also come from natural substances.

4.1. Antitumor Activity

Recently, traditional cancer drugs not only inhibit the growth and kill tumor cells but also demonstrate obvious cytotoxicity to healthy cells and trigger tolerance of the body during long-term use. Developing tolerance to commonly used therapeutic drugs, such as docetaxel, paclitaxel, epiamicyn, adriamycin, etc., requires increasing the dosage that induces more pronounced adverse effects [227]. Therefore, the necessity of new drug development is an obvious and urgent task. Because of their large surface area, high drug-loading capacity, and tumor targeting, nano-drugs have attracted much attention with superiority in tumor treatment. Many studies, some of which are reviewed below, have shown that biogenic NPs themselves indicate significant chemotherapeutic effects on various tumor cells. The main mechanisms accounting for antitumor activity of biogenic NPs are (1) apoptotic pathways depending on increased levels of ROS that led to oxidative stress and DNA fragmentation; (2) interference with macromolecules such as proteins and DNA resulting in disturbance of cell functions; and (3) interaction to cell membranes thus changing its permeability and causing mitochondrial dysfunction [100]. The scientists mainly focus on biogenic NPs synthesized by plants because phytoconstituents could exert activity against cancer cells through the production of ROS participating in phagocytosis, regulation in cell proliferation, and intercellular signaling [228]. Furthermore, phenols and polyphenolic compounds were discovered to be associated with atherosclerosis and inhibition of cancer [229].

4.1.1. Biosynthesized Metal NPs in Cancer Therapy and Delivery of Antitumor Drugs

Some nanomaterials have inherited antitumor activity and can be directly used as drugs. They can work as cytotoxic agents due to their physical-chemical and surface properties. NPs act as nanocarriers to passively target the tumor via enhanced permeability and retention effect (EPR) or actively target solid tumors by ligand-receptor interactions [230]. Active targeting includes the conjugation of different molecules, such as DNA, peptides, and antibodies, to target specific cells [4]. At present, Au NPs are known to be efficient as pharmaceutical drug carriers that improve antitumor efficiency. Au NPs themselves, with a size of 15–35 nm, extracellularly developed by using Paracoccus haeundaensis (marine bacterium), have demonstrated non-toxicity of human cells while preventing the growth of AGS and A549 cancerous cells at different concentrations [231]. Au NPs synthesized from Solanum xanthocarpum leaf extract also demonstrated induced apoptosis of the nasopharyngeal cancer C666-1 cell line with a decline in both cell viability and colony formation upon treatment [232]. The cell death was proved to be caused by autophagy and mitochondrial-dependent apoptotic pathway. Similarly, the amalgamation of Au NPs with Scutellaria barbata plant extract possessed effective anticancer activity against the pancreatic cell line PANC-1 [233]. The authors found upregulation of the expression levels of Caspase-3, Caspase-9, and Bax genes while the gene expression of Bid and BCl-2 genes was downregulated. Likewise, Zhang Y et al. reported that verboascoside (a major bioactive constituent of the Tsoong herb)-loaded Au NPs exhibited apoptosis in both in vitro (K562 tumor cells) and in vivo studies with mouse models [234]. The intravenous injection of the NPs effectively inhibited the growth and induced apoptosis of tumor cells. Additionally, the well-adsorbed tumor cells Au NPs were susceptible to radiation in the near-infrared region which is frequently used in heat therapy for selective cancer elimination [235].
Specific targeting of cytotoxic drugs to malignant cells appeared to be a promising strategy for reducing the adverse side effects and efficacy of anticancer treatment. Nanocarrier delivery systems can propose some advantages in tumor treatment for better efficacy, such as increased solubility of the drug, reduced interaction with non-target tissue, and avoiding drug degradation thus reducing adverse reactions [236]. By using trisodium citrates capping and reducing agent, Au NPs with an average size of about 22 nm were biosynthesized and functionalized with PEG and folic acid and loaded with the anticancer drug Docetaxel [237]. As compared with Docetaxel, the functionalized Au NPs demonstrated good potential for target drug delivery when tested by using the A549 cell line (lung cancer cell line). By using chitosan as a reducing and capping agent, Malati et al. demonstrated that Au NPs conjugated with rifampicin ensured the biocompatible controlled release of the drug in the body [238]. Modified by activated folic acid ligands, biosynthesized Au NPs loaded with chlorambucil showed higher toxicity towards HeLa, RKO, and A549 cell lines as compared to Au NPs and chlorambucil alone [239]. The release rate of the drug was faster at lower pH (5.4) than at pH 6.7 and pH 7.2. Therefore, the synthesized biogenic NPs demonstrated potential for pH-sensitive drug delivery in a tumor microenvironment where the pH value varied between 4.5 and 6.5 and slowed drug release in the bloodstream (pH 7.2).
Alqahtani et al. reported the biosynthesis of Ag NPs with a size of 1 to 40 nm using lichens, namely, Xanthoria parietina and Flavopunctelia flaventior. They demonstrated the antitumor activity of the biogenic Ag NPs against human colorectal (HCT 116), breast (MDA-MB-231), and pharynx (FaDu) cancer cells using a routine MTT assay. The biogenic Ag NPs exhibited higher cytotoxicity towards human colorectal and pharynx cancer cells than breast cancer cells [240]. Doxorubicin is a commonly used chemotherapeutic agent and several attempts have been made to tether them with metal nanoparticles. These efforts are continuing since most cancer cells have developed resistance to doxorubicin. By using curry leaves and neem extracts, Thirumurugan et al. [241] formulated Pt NPs and Ag NPs with encapsulated Doxorubicin (Dox). In vitro tests with MCF-7 cell lines revealed that a better inhibition value (IC50) against MCF-7 demonstrated Dox-coated Ag NPs than Dox-coated Pt NPs. Simultaneously, the toxicity of both drug-loaded nanomaterials was found to be concentration-dependent. It follows that the increased pharmacokinetic characteristics of the engineered nanocomposite result in better therapeutic effects. Phytogenic monodisperse (below 10 nm) Pt NPs and Pd NPs synthesized by using the medical plant Gloriosa superba tuber extract showed high anticancer activity against the human breast adenocarcinoma MCF-7 cell line [242]. Pt NPs showed higher (49.6%) anticancer activity than Pd NPs (36.3%). It was confirmed that the predominant mechanism in anticancer activity was apoptosis-induced due to the externalization of phosphatidyl serine and membrane blebbing. However, biosynthesized Pd NPs containing phenols and flavonoids acquired from white tea (Camellia sinensis) extract with a size of 6–18 nm were found to be more antiproliferative towards human leukemia (MOLT-4) cells than pure white tee extract, doxorubicin or cisplatin [243]. The anticancer activity of the biosynthesized NPs was mediated through the induction of apoptosis and G2/M cell cycle arrest. Recently, Prakashkumar et al. reported the preparation of neem gum-coated Pd NPs that exhibited anticancer and antimicrobial activities. The authors biosynthesized Pd NPs using an aqueous leaf extract of Orthosiphon stamineus. The neem gum-coated Pd NPs also exhibited an anticancer effect in A549 cells via ROS-induced MMP loss activating apoptosis. The authors also demonstrated the biocompatibility of neem gum-coated Pd NPs as they did not exhibit any cytotoxicity or hemolysis, thus making them an ideal candidate for antimicrobial and cancer therapy [244].

4.1.2. Biogenic Metal Oxide NPs in Cancer Therapy and Delivery of Antitumor Drugs

ZnO is considered to be one of the five metal oxides that are recognized as safe and approved by FDA material. Biosynthesized ZnO NPs using Albizia lebbeck stem bark extracts and an average size of 66.3 nm and spherical morphology revealed a strong cytotoxic effect on MDA-MB 231 and MCF-7 breast cancer cell lines in a concentration-dependent manner [122]. Similarly, biogenic porous rod ZnO NPs synthesized by using Borassus flabellifer fruit extract and an average size of 55 nm were utilized for drug delivery of Doxorubicin [245]. The conjugate had low systemic toxicity in a murine model and high loading capacity and therapy efficacy. Drug-loaded ZnO NPs indicated a pH-responsive release of drug preferentially to malignant HT-29 and MCF-7 cells. Not only plant-synthesized but also fungus-mediated produced ZnO NPs also demonstrated anticancer activity. Through Aspergillus niger-mediated synthesis, ZnO NPs with a size of 39–115 nm showed significant cytotoxic effects against rapidly dividing HEP-2 cell lines [246].
Fadeel et al. reported a novel biosynthesis of TiO2 NPs that was prepared from Aloe vera leaf extracts and was utilized as a drug delivery agent for doxorubicin. The authors demonstrated through in vivo studies that the presence of biogenic TiO2 NPs enhanced the anticancer activity of doxorubicin in Ehrlich tumor-bearing mice in comparison to anticancer activities of doxorubicin-loaded liposomes and pure doxorubicin aqueous solutions [247]. Nageshwara et al. reported biogenesis of Ag-doped TiO2 spherical nanoparticles with Acacia nilotica extraction. The authors demonstrated that Ag-doped TiO2 NPs induced cytotoxicity in human breast adenocarcinoma cell lines, thereby proving to be anticancer agents with their utilization as drug delivery agents [248].
Biosynthesized spherical copper oxide NPs (577 nm in size) produced by using Ficus religiosa leaf extract showed stable anti-cancer effects against A549 tumor cells [249]. The authors attributed the apoptotic effect of copper oxide NPs to the generation of ROS triggering disruption of mitochondrial membrane potential.
Because superparamagnetic iron NPs (SPIONs) loaded with a drug can be guided with the help of an external magnetic field to the desired site, they were found to be a promising agent in cancer therapy and magnetic resonance imaging (MRI). To reduce their toxicity and agglomeration, these NPs are generally coated with non-toxic and biologically active compounds [250]. For that purpose, Tyagi et al. [251] have biosynthesized SPIONs via biogenic techniques and further coated them with tamoxifen-conjugated bovine albumin. The NPs were efficiently internalized into breast cancer cell lines (MCF-7 and T47D) and effectively reduced cell proliferation with IC50 values of 5 ± 0.4 μM and 6.3 ± 0.2 μM in MCF-7 and T47D, respectively. The SPIONs were confirmed to be safe for use as DDS by acute toxicity studies in the rat. Likewise, cellulose nanofiber composites with a size of 62.5 nm with Fe3O4-Ag2O NPs loaded with etoposide and methotrexate demonstrated a slow and steady release of methotrexate (63.8%) and etoposide (78.9%) together with non-toxicity which made it suitable for anticancer application [252].
Table 3 shows some examples of biosynthesized metallic and metal oxide NPs, their characteristic features, tested objects, and the main outcomes of their antitumor activities. The cytotoxicity triggered by biosynthesized NPs is based on their type, shape, size, and surface chemistry. The mechanisms of action of biogenic NPs are still not fully studied but based on the results reviewed in this study, it can be concluded that they include both ROS generation and interaction with macromolecules. The formation of ROS by the NPs is an important mechanism in toxicity that interferes with the equilibrium between oxidant and antioxidant species thus causing changes in intercellular activities. The latter mainly includes DNA damage, cell cycle arrest, induction of oxidative stress, and membrane disruption (Figure 5). Both mechanisms activate apoptotic pathways or trigger necrosis. Most of the biogenic NPs achieve both passing targeting and obvious antitumor effects while drug loading for cancer therapy can additionally improve the efficacy and reduce the death rate.

4.2. Biogenic NPs for Antimicrobial/Antifungal Therapy

The development of resistant bacteria and fungi against antibiotics is a globally increasing problem whose solution lies in finding novel materials to alleviate resistant strains. Metals such as Ag, Cu, Au, Fe, and their nano-metallic forms exhibit numerous biocidal activities against resistant Gram-positive and Gram-negative bacteria and eukaryotes. Biosynthesized NPs are known to present higher antimicrobial activity as opposed to commercial or chemically produced NPs because of the medical properties of some plants [276]. Therefore, the activity of biogenic metal NPs depends on the type, size, shape, charge, and capping, as well as genus and species [277]. For example, Hamed et al. demonstrated antimicrobial and antibiofilm activities of Ag NPs synthesized from actinomycetes extracted from the marine sponge Crella cyathophora. These biogenic Ag NPs showed antimicrobial activity towards various Gram-positive and negative bacterial strains particularly P. aeruginosa and E. cloacae. Further, Ag NPs derived from marine actinomycetes exhibited significant biofilm inhibition against B. subtilis, S. aureus, and P. aeruginosa [278]. Shabaan et al. reported the biogenesis of silver, selenium, and zinc oxide NPs from Streptomyces enissocaesilis, which also exhibited antimicrobial activities against the various standard and resistant bacterial isolates [279]. Small-sized (2–7 nm) Pt NPs biosynthesized from a plant extract of Taraxacum leavigatum showed strong antibacterial activity against P. aeruginosa and B. subtilis, which have strong defensive mechanisms against various antibiotics [280]. Phytosynthesized NiO NPs reduced with garlic and ginger increased the bactericidal activity against multi-drug resistant Staphylococcus aureus [281]. ZnO NPs with an average size of 66.25 nm also biosynthesized with Albizia lebbeck stem bark extracts and revealed strong antibacterial potential against Bacillus cereus, S. aureus, E. coli, Klebsiella pneumonia, Salmonella typhi pathogens [122]. Likewise, by using bark extract of Acacia ceasia (L.) Wild, ZnO NPs with an average size of 32.32 nm and hexagonal morphology have been synthesized by Ashwini et al. [282]. The biomimetically synthesized ZnO NPs exhibited strong antibacterial activity against both gram-positive (S. aureus) and, especially, against gram-negative (E. coli) bacteria strains at all tested (250, 500, and 1000 μg) concentrations.
Several biosynthesized NPs have shown noteworthy growth inhibitory results against both bacteria and fungi. For example, biosynthesized iron NPs from Acacia nilotica seedless pods with an average size of 230 nm inhibited the growth of E. coli, S. aureus, Salmonella, Marsa, and Candida [283]. High inhibitory activity, especially against E. coli bacteria and T. harzianum fungal strain, was demonstrated by biosynthesized spherical CuO NPs (5–13 nm in size) by using Syzygium alternifolium stem bark [284]. However, the zone of inhibition was lesser in fungi as opposed to bacteria due to the difference in the structure of the cell walls of both organisms.
The cell membranes of bacteria are made up of peptidoglycans, which are less firm, and the passage of NPs is comparatively easy, especially in the case of gram-negative bacteria. In prokaryotes, multiple mechanisms are efficient in the antimicrobial activity of biogenic NPs. The biogenic NPs exert their antimicrobial activity due to releasing metal ions, interaction, and damage of the cell membrane, formation of pits, and fragmentation of the cell membrane [285]. Thus, the membrane permeability and intercellular communication are interrupted. NPs can interact with thiol groups and phosphorus of proteins and DNA thus disturbing the metabolic processes like DNA replication, protein synthesis, respiratory chains, etc., and causing cell death [286]. Some NPs can trigger the production of ROS thus causing phospholipid oxidation and the collapse of internal RNA, DNA, and proteins [287]. In thick cell wall microorganisms, such as gram-positive bacteria, moderate antibacterial activity is usually observed, since the cell dislocates the NPs by a transport system.
The fungal cell walls are made of chitin having N-acetylglucosamine and a nitrogen group which is firmer to allow the passage of NPs. Antifungal activity of metallic NPs is triggered by interaction with the cell wall and membrane diffusion of metallic ions which can inhibit β-glucan synthase or N-acetylglucosamine which are important components of the cell wall [288]. The induction of ROS and the initiation of oxidative stress cause interaction with different macromolecules such as DNA, RNA, and proteins that can lead to cell death [289].
A schematic representation of the mechanisms of antimicrobial activity of biosynthesized NPs is presented in Figure 6. They include direct contact of the NPs with the cell wall that damages the surface structures and penetration inside the cell. Additionally, the biogenic NPs can indirectly interact with the inside and outside cell environment generating ROS and metal ions that further damage the cell wall, change the permeability of the membrane, and affects protein, ribosome, and DNA functions. The biosynthesized NPs can also directly interfere with macromolecules, ribosomes, and mitochondria, thus ultimately triggering cell death by both mechanisms.
When biogenic NPs were loaded with biocidal drugs, synergetic effects or enrichment of biocidal properties were observed. For instance, Ag NPs with an average size of 9.8 nm that were bio-assembled by cell-free supernatant of Delfitia sp. strain showed increased efficiency against various pathogenic candida strains when conjugated with antifungal drug Miconazole [290]. The authors explained the fungicidal activity of the nanocomposite with inhibition of ergosterol synthesis and biofilm formation by increasing ROS levels. Furthermore, bio-synthesized by Chlorella vulgaris composites of biogenic CuFe2O4@Ag NPs loaded with Ciprofloxacin demonstrated high antibacterial activity against multi-resistant Staphylococcus aureus [291]. The synergetic effect of biosynthesized Ag NPs with Azithromycin and Clarithromycin against microorganisms causing dental caries and periodontal disease (such as Lactobacillus acidophilus, Staphylococcus aureus, Micrococcus luteus, Streptococcus mutans, Bacillus subtilis, E. coli, Pseudomonas aeruginosa, Candida albicans) was also reported by Emmanuel et al. [292]. All of these findings might be useful for determining if the potential efficacy of biosynthesized NPs of various types, sizes, and shapes might be boosted if different species are used for biogenesis. The effect of biogenic NPs on many bacteria is even better than that of bactericides or antibiotics. Most of the outcomes underline that these alternative antibacterial biogenic agents can prevent the growth and spread of microorganisms and alleviate microbial disease illnesses.

4.3. Anti-Inflammatory Activity

Inflammation is a response mechanism of the body to various external harmful factors. Mechanisms such as a gathering of macrophages and killer cells and the synthesis of cytokines, such as IL-1, IL-1β, IL-6, TNF-α, etc. in the desired site, develop the onset of inflammation [293]. Mild inflammation has a positive effect on living tissue for preventing the spread of pathogens and limiting the lesions. On the contrary, severe inflammation processes can cause denaturation or necrosis of organs, and can lead to fluid accumulation and mutual adhesion between organs, thus harming their normal functionality. Therefore, anti-inflammatory drug application is important in some disease treatments. The usually used steroidal and non-steroidal anti-inflammatory drugs have adverse effects and cause harmful reactions, which limit their application. Additionally, during acute inflammation, the drug efficacy is delayed because of slow absorption [294].
Biosynthesized nano-based formulations are proven in developing antitumor drugs by blocking pro-inflammatory cytokines and ROS scavenging mechanisms and inhibiting different signal pathways. For example, silver NPs generated from Selaginella myosurus aqueous extract showed in vitro and in vivo anti-inflammatory potential by inhibiting protein denaturation. In the Carrageenen-induced rat hind paw edema model, the Ag NPs interfered with inflammatory mediators like histamines, serotonins, prostaglandins, kinins, etc., and blocked their action [295]. Likewise, gold NPs synthesized from Prunus serrulate fruit extract suppressed the production of pro-inflammatory cytokines and reduced the expression of inflammatory mediators such as nitric oxide (NO), prostaglandin E2, inducible nitric oxide synthase (iNOS) and cyclooxygenase-2 (COX-2) in RAW264.7 cells [296]. Phospholipase A2 enzyme is known to participate in the primary way of producing prostaglandins. Biogenic ZnO NPs produced from Lantana camara flower extract with an average size of 21.4–27.2 nm were found to inhibit phospholipase A2 activity via binding Zn2+ ions to its active site [297], thus performing their anti-inflammatory activity. Similarly, biosynthesized ZnO NPs employing leaf extract of Pelargonium odoratissimum (L.) with an average size of 34 nm showed a concentration-dependent anti-inflammatory effect via stabilizing the biological membranes, thus preventing the release of active mediators of inflammation and lytic enzymes [298]. A scheme summarizing the anti-inflammatory action of biogenic NPs is presented in Figure 7. Consequently, the development of biogenic nano candidates as an alternative clinical approach for anti-inflammatory treatment requires an understanding of the regulatory pathways and molecular mechanisms triggered by these nanomaterials.

4.4. Wound Healing Properties

Wounds are classified into acute and chronic based on healing time and other complications [299]. During the healing of various cells, growth factors, cytokines, coagulation, and vascularization growth factors intricate each other. Current therapies involve applying steroidal, nonsteroidal, and chemotherapeutic drugs with various side effects or therapies such as dressing materials, vascular surgery, hyperbaric oxygen therapy, etc. With the advance of the nano concept and knowledge of the significant antibacterial and anti-inflammatory properties of biogenic NPs, these nano-substances were found to be appropriate to facilitate skin wound healing. Predominate bacteria found during the early stages of infections are Gram-positive (like S. aureus, S. pyogenes), whereas Gram-negative (such as E. coli, P. aeruginosa) are found later when a chronic wound develops [300]. When the immune system is unable to remove the pathogens, the wound-healing process is disrupted [300].
When using biogenic NPs, wound therapy could be based on nanomaterials acting in the manner of drugs or biogenic nano-vehicles for delivery of agents to repair wounds. For example, Lindera strychnifolia extract produced Ag NPs with a 15.7 nm average size and showed the highest % (64%) of wound closure among seven biosynthesized Ag NPs determined by the scratch method on NH3T3 cells [301]. Magnetic (Fe2O3) NPs with 15–30 nm size biosynthesized by using Aloe vera extract in isolated bacterial nano-cellulose (BNC) were able to promote wound healing of human dermal fibroblast cells after 48 h [302]. The Fe2O3/BNC nanocomposite reduced the expression of microRNA (miR-29b and miR-29c) thus causing an increase in gene expression of other genes (TGF-β1, CTNNB1, MMP2, MMp9, WNT4), which resulted in faster wound healing. Similarly, Barleria gibsoni leaf extract mediated synthesis of ZnO NPs with sizes between 30 and 80 nm improved substantially the wound healing efficacy in male albino Wistar rats [303]. The wound contraction percentage of the nano-ZnO gel formulation reached above 98% on the 20th day as compared with the control group, thus avoiding wound chronicity. Sankar et al. confirmed that the Ficus religiosa leaf extract-tailored copper oxide NPs demonstrated both substantial inhibition activity against human pathogenies strains and enhanced wound healing activity in Wistar Albino rats [304].
To prevent would infections, various dressing materials with bacteria barrier properties, the ability to promote wound healing, and antibacterial potential have been designed [305]. In this regard, promising candidates are NP-loaded wound dressings that can reduce wound contraction time and wound infections without substantial side effects [306]. Such an approach includes the development of biodegradable and porous polymer membranes loaded with biogenic antimicrobial agents to promote healing and inhibit pathogens [307]. For example, Augustine et al. used photosynthesized Ag NPs through black pepper (P. nigrum) to incorporate them in electrospun polycaprolactone (PCL) membranes, and provided antimicrobial properties of the dressing composite material [308]. The membranes exhibited higher mechanical properties and antibacterial activities against both E. coli and S. aureus [309]. Moreover, the same research group utilized Ag NPs photosynthesized by Mimosa pudica to obtain electrospun biodegradable polyvinyl alcohol (PVA) membranes. The latter demonstrated good mechanical strength, wound fluid uptake, blood compatibility, cytocompatibility, and antibacterial potential. At optimum Ag NP composition, the composite membranes booted the wound healing in an in vitro wound contraction model.
Although the exact mechanism of wound healing is not yet well understood, its initial reaction is supported by inhibiting the proliferation of the microbial population. Therefore, the use of biogenic metal or metal oxide NPs with intrinsic antibacterial activity is a promising approach. Further, the biogenic NPs can induce ROS formation and activate angiogenesis factors (VEGF, fibroblast growth factor (FGF)) to accelerate wound healing. To clarify the exact mechanism of NP-mediated wound healing, extensive research on different phases of healing, toxicity, and biocompatibility is needed to enhance NPs’ therapeutic potential.

4.5. Osteoinductive and Angiogenetic Activities

Recently, scientists have been exploring various methods to enhance bone tissue regeneration. Important examples are metallic and metal oxide NPs. Materials for bone regeneration and remodeling should not only be biocompatible but also ought to exhibit anti-infective activity and promote angiogenesis. NPs have likewise appeared to have important antibacterial and antifungal properties. Moreover, it was found that they also have osteoinductive properties or the ability to induce bone formation after implantation. For example, Jadhav et al. [310] phyto-synthesized Au NPs by using Salacia chinensis to evaluate their osteoinductive activities. The NP exposure to GNPs and MG63 cell lines increased the percent of cell viability from 96 ± 3.7 (control) up to 138 ± 27.4, which confirmed their osteogenic potential. Consequently, it appears that such NPs may be used as active bone inductive material during implant placement. Plant (Anogenissus latifolia) polyphenol biosynthesized Au NPs that were found to be stable in a vast range of blood components were also utilized as an effective pain reliever and osteoinductive adjuvant of dental tissue implantation [311]. Similarly, ZnO NPs with an average size of 26 nm phyto-synthesized by Scutellaria baicalensis showed enhanced differentiation, proliferation, and mineralization during a 14-day period of cultivation with MG63 cells [312]. The authors proposed that the contact between the cells and biomaterial induced signal transduction pathway-regulated inductions, which influenced cell behavior. The biogenic ZnO NPs were non-toxic and biocompatible. Increasing the biocompatibility of the implant will ultimately lead to a longer life span of the graft and higher effectiveness of the implantation.
Angiogenesis is a vital process of the development of new blood vessels as a sequence of several cellular and molecular mechanisms [313]. This includes activation of endothelial cells in pro-angiogenic growth, degradation of capillary walls of the existing vessels, branch point formation inside the vessel, and migration of endothelial cells into the extracellular matrix and then to the starting point of angiogenetic stimulation [314]. The activation of endothelial cells can happen in response to specific conditions (such as hypoxia and inflammation) or during would healing to enhance tissue regeneration [315]. Angiogenesis is controlled by pro-angiogenic and anti-angiogenic factors such as vascular endothelial growth factor (VEGF), fibroblast growth factor (FGF), transforming growth factor- beta (TGF-β), tumor necrosis growth factor-alpha (TNF-α), IL-8, etc. [316]. The role of nanostructures in promoting angiogenesis is widely recognized in tissue engineering [317]. It was found that the application of magnetic stimulation can synergistically boost bone formation and pro-angiogenic potential more than just magnetic NPs alone [318]. For that reason, Moise et al. [319] bacterially (Geobacter sulfurreducens) synthesized Zn- and Co-doped magnetite NPs to tailor their magnetic properties. Negligible cytotoxicity of both moderately doped NPs on primary human bone marrow-derived mesenchymal stem cells (hMSCs) and human osteosarcoma-derived cells (MG63) was observed. A relatively strong AC susceptibility signal was detected from Zn-doped NPs compared to Co-doped NPs. With time the NPs were internalized by cells and stored within lysosomes. The osteogenic potential of hMSCs was unaffected by the uptake of Zn- and Co-doped NPs. It follows that utilizing biogenic NPs can yield superior bone regeneration accompanied by blood vessel formation in complex bone tissue engineering. However, more studies are required.

4.6. Anti-Viral Activity

Viruses are one of the contagious agents causing various diseases in humans. Some viral infections triggered by hepatitis, herpes-simples virus (HSV), influenza, COVID, human immunodeficiency virus (HIV), etc. cause pathologically complicated diseases and could be life-threatening. Available antiviral drugs, such as acyclovir, inhibit viral multiplication but have many side effects. Numerous drugs generated from plant and microbial sources have been recently formulated and commercialized as antiviral agents [320]. Moreover, metallic NPs are found to act as viricidal agents capable of blocking viral spread, interacting with viruses, and suppressing free virions [321]. The antiviral potential of NPs can be attributed to adsorption onto viral surfaces thus blocking the virus penetration and entry into the host [322], direct interaction of NPs with the viral proteins and viral genome [323], and inhibition of generic expression and production of viral constituents [324]. When supported by the viricidal potential of capping biocomponents, their anti-viral activity can be accelerated.
For instance, biosynthesized Ag NPs through Andrographis paniculate, Phyllanthus niruri, and Tinospora cordifolia demonstrated antiviral activity against Chikungunya virus [325]. The cytotoxicity assays in Vero cells revealed that Andrographis paniculate Ag NPs were the most toxic with a maximum non-toxic dose (MNTD) value of 31.25 μg/mL, followed by Tinospora cordifolia and Phyllanthus niruri Ag NPs. The cell viability tests also confirmed that Andrographis paniculate Ag NPs inhibited the virus to a maximum extent. Al-Sanea et al. synthesized Ag NPs by a biogenic route that were highly potent against SARS-CoV-2 [326]. The authors used strawberry (Fragaria ananassa Duch.) and ginger (Zingiber officinale) methanolic extracts for the biosynthesis of these Ag NPs with an average size of 5.9 and 5.8 nm, respectively, both with spherical shapes. The strawberry-synthesized NPs showed the highest antiviral activity against SARS-CoV-2 with IC50 value of 0.099 μg/mL. Molecular docking studies have shown that neohesperidin, a dereplicated compound, can bind to Aak1 (adoptor-associated kinase 1—a host kinase that regulates the intracellular viral trafficking during the entry) through hydrogen bonds and thus can act as a dual inhibitor for both Aak1 and viral NSP 16 (SAM-dependent methyltransferase that methylates the RNA cap). In another study, Ag NPs have been synthesized by a biogenic route through Cinnnamomum cassia bark extract and tested against H7N3 influenza A virus in Vero cells [327]. The biosynthesized NPs were found to be effective in both treatments when incubated with the virus before infection and when introduced after infection. The biomimetic NPs were also proved to be non-toxic to Cero cells up to 500 μg/mL. Biogenic Ag NPs (size of 12–28 nm) biosynthesized through an extract of mangrove Rhizophora lalarckii also showed anti-HIV activity by inhibiting HIV 1 type reverse transcriptase even at low doses [328]. In vitro, the mangrove fabricated Ag NPs exhibited an IC50 of 0.4 μg/mL on the HIV reverse transcriptase. Simultaneously, Dev et al. showed that Ag NPs conjugated with the anti-retroviral drug lamivudine were found to be highly prospective drug-delivery agents acting as a potent and selective inhibitors of type 1 and type 2 HIV [329].
Except for silver, gold NPs were also proven to be effective in antiviral therapy. Seaweed Sargassium wightii was used for the synthesis of biogenic Au and Ag NPs that were evaluated against herpes simplex virus (HSV-1 and HSV-2) [330]. The biogenic Au NPs showed a reduced by 70% cytopathic effect of HSV-1 and HSV-2 at 10 μL and 25 μL, respectively. Ag NPs showed a decrease in cytopathic effect by 70% of HSV-1 and HSV-2 at 2.5 μL. At the same time, the Au NPs were significantly non-toxic at all tested concentrations, while Ag NPs were toxic at higher concentrations.
Metal oxide NPs also demonstrated potential in viricidal therapy. For example, Yugandhar et al. synthesized biogenic CuO NPs by using fruit extract of Syzygium alternifolium [331]. The average size of the biogenic Cu NPs was 17.5 nm varying from 2 to 69 nm. The authors showed a significant growth inhibitory effect against Newcastle Disease Virus (NDV). The effective concentration of CuO NPs against NDV was 100 μg/mL.
It follows that by coupling the anti-viral activities of metallic NPs, biochemicals, and/or drugs, more accessible and low-cost anti-viral nanotherapeutics with minimal cytotoxicity can be produced.

4.7. Antiparasitic Activity

Parasites are organisms that live within other living species while utilizing their food and shelter. They are considered more pathogenic than bacteria to humans, since they can cause chronic diseases [332]. Parasitic infections are usually treated by chemotherapeutic agents and phytoextracts. However, some of the treatments are no longer effective because some protozoa have developed resistance [333]. Additionally, the bioavailability of antiparasitic drugs is low because of their short half-life and insolubility [334]. The synthesis of biogenic NPs is an approach to combine the efficacy of plant-based material with the achievements of biogenic nanotechnology to obtain effective antiparasitic therapeutics. For example, the parasiticidal effect of Karthik et al. reported a biosynthesis of Ag NPs from Streptomyces sp. and the obtained nanoparticles exhibited antiparasitic (acaricidal) activity against Rhipicephalus microplus and Haemaphysalis bispinosa [335]. Alajmi et al. observed antiparasitic activity of biosynthesized Ag NPs prepared by a combination of plant extracts of Phoenix dactylifera and Ziziphus spina-chrisri (traditionally used against toxoplasmosis) against oblige intracellular apicomplexan protozoan parasite Toxoplasma gondii [336]. As an alternative to standard sulfadiazine drug therapy, nanoparticle pretreatment prevented liver damage as determined by significant inhibition of hepatic NO levels and elevation in liver superoxide dismutase (SOD) and catalase (CAT) activities. NP treatment decreased proinflammatory cytokines and boosted the antioxidant enzyme activity of liver homogenate compared with an untreated control group. NP treatment induced a reduction in immunoreactivity to TGF-β and NF-kB in hepatic tissue, too. Improvements in the histological features of liver tissue and fewer degenerations were also observed.
Similarly, Leishmaniasis is a life-threatening disease caused by the parasite Leishmania. Some of the commercially applied anti-Leishmania drugs suffer from resistance mechanisms. For that reason, Ullah et al. synthesized Ag NPs through both chemical and biogenic methods from Teucrium stocksianum aqueous plant extract and evaluated their antileishmanial activity [337]. The study indicated strong antiparasitic activity against Leishmania infantum promastigotes for both NPs with higher efficacy for the biosynthesized (IC50 30.71 ± 1.91 μg/mL) as opposed to the chemically obtained (IC50 51.23 ± 2.2 μg/mL). At the same time, the MTT assay showed that the chemical Ag NPs exert higher toxicity than the biogenic Ag NPs. An efficient method of drug delivery to enhance the antileishmanial activity of miltefosine with Ag NPs was proposed by Kalangi et al. [338]. The biosynthesized Ag NP through Anethum graveolens leaf extract had an average size of 35 nm and demonstrated biocompatibility pertaining to >80% viability of macrophages. In combination with miltefosine (12.5 μM and 25 μM), the Ag NPs magnified the leishmanicidal effect of miltefosine by 2-folds. The enhanced effect of miltefosine (12.5 μM) in combination with AgNPs (50 μM) was confirmed by morphological aberrations and DNA fragmentations in promastigotes. Similarly, ZnO NPs synthesized by using Aquilegia pubiflora showed a dose-dependent cytotoxicity against Leishmania tropica (KWH23) with significant IC50 for both promastigote (48 μg/mL) and amastigote form (51 μg/mL) of the parasite [339]. In another study, antileishmanial properties of ZnO were enhanced by formulating bimetallic ZnO/Ag NPs that were synthesized using Mirabilis jalapa [340]. The activity of the bimetallic NPs was greater than monometallic ones while increasing with the increase of concentration. A possible mechanism of the antiparasitic activity of biosynthesized NPs can be their interaction with the cell membrane and cellular enzymes, causing intracellular damage and a high level of production of ROS that trigger apoptotic mechanisms in the cell.

5. Cytotoxicity of Biosynthesized NPs

Although nanomaterials and nanotechnology have developed considerably during the last 20 years, their potential toxicological effects on humans, animals, plants, and the environment have only currently received adequate attention. Previous studies discovered that NPs induce toxicity by a generation of ROS or increased oxidative stress in cells that can trigger cell death by inducing apoptosis or necrosis process [341]. When biogenic NPs are subject to drug-delivery application, they can suffer from instability in the hostile environment and bioaccumulation that can trigger biotoxicity. The biogenic techniques can impart substantial stability by capping with various agents [342]. According to Docea et al. the cellular uptake and colloidal stability of “coated NPs” are among the factors that control the toxicological features of NPs against cells [343]. However, the shape, size, and composition of NPs are characteristics that also determine biogenic nanomaterial toxicity [344]. Therefore, modulating these parameters will enable the designing of metallic NPs with improved interactions, biological activity at a target site, and low cytotoxicity. Sharma et al. [345] stated that the rate of NP dissolution can be controlled by the size of the material, capping, or surface functionalization. It was found that the cell viability (MTT analysis) of bone stem cells after incubation for 24, 73, and 96 h with biogenic and chemically synthesized Ag NPs was decreased but the toxicity of the biogenic NPs was substantially (11 folds) lower [223]. It is thought that the major toxicity of Ag NPs is associated with free silver ions that may negatively affect the eyes, skin, kidney, and liver cause respiratory problems and change blood cells [346]. Therefore, the non-toxic coatings (capping agents) provide a higher degree of biocompatibility and lower toxicity.
Rheder et al. observed that Ag NPs synthesized from an infusion of roots (AgNPR) and leaf extract (AgNPE) of the same plant (Althea officinalis) display different toxicity against A549 and V79 cell lines and none of them presented cytotoxicity towards HaCat cells [347]. The viability of the A549 and V79 cells presented cell death with low IC50 values when treated with AgNPE compared to AgNPR, but both materials exhibited physicochemical changes during exposure assays. In vitro, the cytotoxic effect of Ag NPs obtained by Butea monosperma bark extract against human PBMCs and leukemic KG-1A cell lines was tested by Pattanayak et al. [348]. In KG-1A the IC50 value of Ag NPs was 11.5 μg/mL while in human PBMCs it was 43.2 μg/mL. In leukemic cells, the ROS generation level increased over 4 folds at the effective dose and typical characteristics of apoptosis like plasma membrane blebbing and formation of apoptotic bodies were observed. Chromatin condensation and fragmentation also suggested that the Ag NPs deserved potential genotoxic effects. However, no in vivo studies have been performed.
Murugesan et al. [349] synthesized coated Ag NPs by curcumin derivate (ST06) and assessed their in vivo cytotoxicity in the Ehrlich Ascites carcinoma tumor-induced mouse model. ST06-Ag NPs with sizes ranging between 50 and 100 nm exhibited higher ani-tumor efficacy because of impaired angiogenesis than free ST06 or Ag NPs. In a dose of 5 mg/kg, the biogenic NPs showed no significant toxic effects in the animals. Using an acute inflammation model in Wistar rats, Moldovan et al. [350] demonstrated that the presence of some functional groups with biological origin on the NP surface can reduce cytotoxic effects. Spherical Ag NPs (10–50 nm) photosynthesized from aqueous extract of European cranberry bush (Viburnum opulus L.) decreased the level of cytokines in the soft hind foot pad tissue early after the induction of inflammation (2h) and reduced the increase in paw volume by inhibiting the inflammatory processes with edema formation.
It is thought that due to the higher concentration in blood circulation, Ag NPs are located in almost all vital organs, especially the liver [351]. The levels of antioxidant enzymes, such as CAT (catalase) and SOD (superoxide dismutase), decrease in the presence of tumors since cancer cells generate a large amount of hydrogen peroxide [352]. When treated with F. religiosa derived Ag NPs (5–35 nm) the levels of these liver antioxidant enzymes increased near to normal in Dalton’s ascite lymphoma mice models [353]. Histopathological studies also revealed recovery of liver architecture upon treatment with Ag NPs whereas hematological parameters revived to normal values after treatment. Similarly, Morus alba leaf extract formulated Ag NPs were examined for their hepatotoxicity in Wistar rats by intraperitoneal injection [223]. The authors observed that the therapy with Ag NPs at all doses (25, 50, and 100 μg/kg) significantly restored the metabolic enzyme activities, which indicated improved physiological functioning of hepatic tissue. Histopathological observation confirmed the healing of hepatic parenchyma and regeneration of hepatocytes.
However, in vivo, studies of biogenic Ag NPs synthesis using Desmodium gangeticum aqueous extract revealed obvious nephrotoxicity. This nephrotoxicity of both chemically and biosynthesized Ag NPs in a proximal epithelial cell line, renal mitochondria, and rats were evaluated by Vasanth and Kurian [354]. After 15 days of oral administration of Ag NPs (100 mg/kg) to the Wister rats, significant changes in renal architecture were seen in both receiving rats. The urine and blood chemistry data, as well as the renal epithelial cells and renal mitochondria, confirmed the toxic similarities between the Ag NPs produced by two different routes. Similarly, Ag NPs synthesized using sulfonated polysaccharide extract from Sargassium siliquosum, a brown alga, with polydisperse size (20–480 nm) [355], did not cause mortality to rats up to 2000 mg/kg BW but triggered elevation of serum creatinine and blood urea nitrogen. A low dose was effective in a revival of liver enzyme parameters to normal in intoxicated groups with paracetamol-induced liver injury. However, an apparent toxic effect on the kidney was also observed.
Many researchers have reported lower cytotoxicity of biomimetic Au NPs than Ag NPs irrespective of the cell type used [356]. For example, biosynthesized Au NPs from Curcuma manga were found to be cytocompatible with human lung fibroblast cells (MRC-5) and human colon fibroblast cells (CCD-18Co) [357]. These Au NPs also demonstrated hemocompatibility with less than 10% of hemolysis without any aggregation of erythrocytes. At the same time, Nandhini et al. reported the biogenesis of gold NPs from Enterococcus sp. RMAA and studied its cytotoxic potential against human hepatocellular cancer (HCC) cell lines. It was revealed that Au NPs resulted in about 20% cytotoxicity of cancer cell lines due to oxidative stress and reduced mitochondrial membrane potential that eventually induces cytochrome c release into cytosol leading to apoptosis of cancer cells [358]. Moreover, biosynthesized Au NPs using plant extract were shown to be non-toxic to normal L-cells at various concentrations (1 to 100 μL) [359]. The results underlined the potential of an extract-based method for the synthesis of Au NPs for therapy and anticancer treatment.
In another study, in vitro toxicity of pullulan-supported Au NPs loaded with 5-fluorouracil and folic acid against HepG2 cells was studied [360]. It was concluded that the amount of 5-fluorouracil required to achieve 50% of growth inhibition was much lower for the conjugate than in free 5-fluorouracil. Additionally, the in vivo distribution of gold was increased mostly in the liver than in other organs. Lee et al. reported the synthesis of chitosan-capped gold nanospheres and nanostars, derived from green tea extracts. At the same time, gold nanorods were prepared by a conventional method. The authors demonstrated cytotoxicity of the biogenic Au NPs towards four cancer cell lines, of which the highest toxicity towards hepatocyte carcinoma was exhibited in nanorods, followed by nanostars and finally nanospheres. Further, the cellular uptake of Au NPs by human hepatocyte carcinoma cells (HepG2) followed the order nanospheres > nanorods > nanostars [361]. Moreover, the study of Mironova et al. demonstrated that after 3 days of incubation, 142 μg/mL of 13 nm AuNPs cause nearly 40% apoptosis in human fibroblast cell lines while only 13 μg/mL of 45 nm AuNPs triggered the same effect [362]. However, after 14 days full recovery occurred for the fibroblast cells after the removal of both particles. Hence, it is demonstrated that the shape, size, and origin of metal NPs play a crucial role in therapy and as drug delivery agents.
Eisa et al. reported the biosynthesis of TiO2 NPs from titanium isopropoxide and lupin bean extracts and demonstrated its antimicrobial and cytotoxic activities. These biogenic TiO2 NPs exhibited cytotoxicity against breast cancer cell (MCF-7) lines with an IC50 of about 41.1 µg. The authors compared the cytotoxicity of biogenic TiO2 NPs with Vinblastine against breast cancer cells, where the former showed greater efficacy in cancer therapy [363]. Selim et al. reported cytotoxic effects of a plant (Deverra tortuosa aqueous extract)-derived ZnO NPs with an average size of 15.2 nm on cancer cells. The authors demonstrated anticancer activity in vitro model against human colon (Caco-2) and lung (A549) cancer cell lines and compared them to human lung fibroblast (WI38) cell lines using MTT assay [364]. The biogenic NPs showed remarkable selective cytotoxicity against both examined cancer cell lines. As of now, one of the most challenging fields in anticancer treatment is the development of new anticancer drugs with minimal side effects and enhanced efficacy and selectivity, and the biosynthesis of NPs could pave the way toward more precise and effective anti-cancer treatment. However, it follows that the method of biosynthesis, reducing and capping biological material, type of NP, and its shape are all determinative for the toxicity of that NP.

6. Future Perspectives and Outlook

Nanobiotechnology, by combining chemistry, engineering, biology, and medicine, offers various benefits in treating severe or chronic human diseases via target-oriented delivery. The biosynthesis of NPs is now an important branch in nanobiotechnology. By combining the unique properties of metallic NPs with the benefits of conjugated synthesis and surface functionalization, a wide range of opportunities give room for discoveries and various biomedical applications. From the relevant studies, it appears that biogenesis is not only an eco-friendly and inexpensive alternative for conventional NPs production but also an option to design multifunctional nanomaterial that can effectively treat diseases or deliver drugs without disturbing the organism or causing less harmful effects. Biogenic synthesis usually allows synergy between the nanoparticle and the substance from the organism employed for production. The use of genetically modified microorganisms for the biogenic production of NPs with a desired shape and size is a novel and promising approach in this area. The potential activity and cytotoxicity of these biosynthesized NPs will be affected by the presence and type of coating on their surface.
However, most of the biosynthesis methods are still under development, and issues such as purification and separation of NPs are to be further overcome. Moreover, there are certain gaps in knowledge in a thorough understanding of mechanisms and biochemical pathways of bio-mediated synthesis of NPs. When dedicated to biomedical purposes, it is vital to know the active moieties that bind and participate in synthesis and provide surface stability and biocompatibility. Large-scale production of biocompatible nanostructures with narrow size distribution is another challenge in front of recent biotechnology methods. The heterogeneity may hamper their use in different biomedical applications where precise shape and size are of vital importance. Exact determination of the reducing components should be performed to establish precise and repeatable methods for biosynthesis. Moreover, the structure and composition of biogenic NPs together with their coatings must be carefully analyzed and the potential interactions with the biological system must be evaluated. The possibility of aggregation and formation of agglomerates that can affect the properties and bioavailability of the individual NPs should also be assessed to avoid unexpected toxic effects. Actinomycetes and plant-mediated metal nanoparticle synthesis have shown promise as therapeutic agents whose mechanism of action is yet to be explored.
Concerning the biomedical application of these NPs, another issue is related to their distribution profile, release kinetics, and clearance in the organism. However, most of the developed conjugates by biogenic processes are biomaterials with reduced toxicity when compared to their physiochemically synthesized counterparts. As such systems, biogenic NPs are advantageous to avoid the main issue faced by modern drug-delivery systems: nano-cytotoxicity. The fabrication of biogenic nanocomposites for drug delivery in a more intelligent and focused approach, it seems, will have a bright future in biomedicine.

Author Contributions

Conceptualization, M.P.N.; methodology M.P.N.; software M.P.N.; validation, M.P.N., P.B.J. and M.S.C.; writing—original draft preparation, M.P.N.; writing—review and editing, M.P.N., P.B.J. and M.S.C.; visualization—M.P.N.; funding acquisition—M.P.N. All authors have read and agreed to the published version of the manuscript.

Funding

This publication was developed with the support of Project BG05M2OP001-1.001-0004 UNITe, funded by the Operational Programme “Science and Education for Smart Growth”, co-funded by the European Union through the European Structural and Investment Funds.

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. Singh, J.; Dutta, T.; Kim, K.-H.; Rawat, M.; Samddar, P.; Kumar, P. ‘Green’ synthesis of metals and their oxide nanoparticles: Applications for environmental remediation. J. Nanobiotechnol. 2018, 16, 84. [Google Scholar] [CrossRef] [PubMed]
  2. Ahmed, S.; Saifullah, A.M.; Swami, B.L.; Ikram, S. Green synthesis of silver nanoparticles using Azadirachta indica aqueous leaf extract. J. Radiat. Res. Appl. Sci. 2019, 9, 1–7. [Google Scholar] [CrossRef] [Green Version]
  3. Hua, S.; de Matos, M.B.C.; Metselaar, J.M.; Storm, G. Current trends and challenges in the clinical translation of nanoparticulate nanomedicines: Pathways for translational development and commercialization. Front. Pharmacol. 2018, 9, 790. [Google Scholar] [CrossRef] [Green Version]
  4. Zhang, D.; Ma, X.-L.; Gu, Y.; Huang, H.; Zhang, G.-W. Green Synthesis of Metallic Nanoparticles and Their Potential Applications to Treat Cancer. Front. Chem. 2020, 8, 799. [Google Scholar] [CrossRef] [PubMed]
  5. Shah, M.; Fawcett, D.; Sharma, S.; Tripathy, S.K.; Jai Poinern, G.E. Green synthesis of metallic nanoparticles via biological entities. Materials 2015, 8, 7278–7308. [Google Scholar] [CrossRef] [Green Version]
  6. El Shafey, A.M. Green synthesis of metal and metal oxide nanoparticles from plant leaf extracts and their applications: A review. Green Process. Synth. 2020, 9, 304–339. [Google Scholar] [CrossRef]
  7. Phan, H.; Haes, A.J. What Does Nanoparticle Stability Mean? J. Phys. Chem. C 2019, 123, 16495–16507. [Google Scholar] [CrossRef]
  8. Chung, I.M.; Rahuman, A.A.; Marimuthu, S.; Kirthi, A.V.; Anbarasan, K.; Padmini, P.; Rajakumar, G. Green synthesis of copper nanoparticles using Eclipta prostrata leaves extract and their antioxidant and cytotoxic activities. Nanoscale Res. Lett. 2017, 11, 18–24. [Google Scholar] [CrossRef] [Green Version]
  9. Bonifacio, B.; da Silva, P.; Ramos, M.; Negri, K.; Maria Bauab, T.; Chorilli, M. Nanotechnology-based drug delivery systems and herbal medicines: A review. Int. J. Nanomed. 2014, 9, 1–5. [Google Scholar] [CrossRef] [Green Version]
  10. Mody, V.V.; Nounou, M.I.; Bikram, M. Novel nanomedicine-based MRI contrast agents for gynecological malignancies. Adv. Drug Deliv. Rev. 2009, 61, 795–807. [Google Scholar] [CrossRef]
  11. Tinajero-Díaz, E.; Salado-Leza, D.; Gonzalez, C.; Velázquez, M.M.; López, Z.; Bravo-Madriga, J.; Knauth, P.; Flores-Hernández, F.Y.; Herrera-Rodríguez, S.E.; Navarro, R.E.; et al. Green Metallic Nanoparticles for Cancer Therapy: Evaluation Models and Cancer Applications. Pharmaceutics 2021, 13, 1719. [Google Scholar] [CrossRef] [PubMed]
  12. Beveridge, T.; Murray, R. Sites of metal deposition in the cell wall of Bacillus subtilis. J. Bacteriol. 1980, 141, 876–887. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  13. Gahlawat, G.; Choudhury, R. A review on the biosynthesis of metal and metal salt nanoparticles by microbes. RSC Adv. 2019, 9, 12944. [Google Scholar] [CrossRef] [Green Version]
  14. Deshmukh, S.P.; Patil, S.M.; Mullani, S.B.; Delekar, S.D. Silver nanoparticles as an effective disinfectant: A review. Mater. Sci. Eng. C 2019, 97, 954–965. [Google Scholar] [CrossRef] [PubMed]
  15. Siddiqi, K.S.; Husen, A. Recent advances in plant-mediated engineered gold nanoparticles and their application in biological system. J. Trace Elem. Med. Biol. 2017, 40, 10–23. [Google Scholar] [CrossRef]
  16. Patil, M.P.; Bayaraa, E.; Subedi, P.; Piad, L.L.A.; Tarte, N.H.; Kim, G.D. Biogenic synthesis, characterization of gold nanoparticles using Lonicera japonica and their anticancer activity on HeLa cells. J. Drug Deliv. Sci. Technol. 2019, 51, 83–90. [Google Scholar] [CrossRef]
  17. Mathivanan, K.; Selva, R.; Chandirika, J.U.; Govindarajan, R.K.; Srinivasan, R.; Annadurai, G.; Duc, P.A. Biologically synthesized silver nanoparticles against pathogenic bacteria: Synthesis, calcination and characterization. Biocatal. Agric. Biotechnol. 2019, 22, 101373. [Google Scholar] [CrossRef]
  18. Qu, Y.; Pei, X.; Shen, W.; Shen, W.; Zhang, X.; Wang, J.; Zhang, Z.; Li, S.; You, S.; Ma, F.; et al. Biosynthesis of gold nanoparticles by Aspergillum sp. WL-Au for degradation of aromatic pollutants. Physica E Low-Dimens. Syst. Nanostruct. 2017, 88, 133–141. [Google Scholar] [CrossRef]
  19. Senapati, S.; Syed, A.; Moeez, S.; Kumar, A.; Ahmad, A. Intracellular synthesis of gold nanoparticles using alga Tetraselmis kochinensis. Mater. Lett. 2012, 79, 116–118. [Google Scholar] [CrossRef]
  20. Liua, Q.; Aouidat, F.; Sacco, P.; Marsich, E.; Djaker, N.; Spadavecchia, J. Galectin-1 protein modified gold (III)-PEGylated complex-nanoparticles: Proof of concept of an alternative probe in colorimetric glucose detection. Colliods Surf. B 2020, 185, 110588. [Google Scholar] [CrossRef]
  21. Busch, R.T.; Karim, F.; Weis, J.; Sun, Y.; Zhao, C.; Vasquez, E.S. Optimization and Structural Stability of Gold Nanoparticle–Antibody Bioconjugates. ACS Omega 2019, 4, 15269–15279. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  22. Ameen, F.; AlYahya, S.; Govarthanan, M.; Aljahdali, N.; Al-Enazi, N.; Alsamhary, K.; Alshehri, W.A.; Alwakeel, S.S.; Alharbi, S.A. Soil bacteria Cupriavidus sp. mediates the extracellular synthesis of antibacterial silver nanoparticles. J. Mol. Struct. 2020, 1202, 127233. [Google Scholar] [CrossRef]
  23. Hooda, H.; Singh, P.; Bajpai, S. Effect of quercitin impregnated silver nanoparticle on growth of some clinical pathogens. Mater. Today Proc. 2020, 31, 625–630. [Google Scholar] [CrossRef]
  24. Vance, M.E.; Kuiken, T.; Vejerano, E.P.; McGinnis, S.P.; Hochella Jr, M.F.; Rejeski, D.; Hull, M.S. Nanotechnology in the real world: Redeveloping the nanomaterial consumer products inventory. J. Nanotechnol. 2015, 6, 1769–1780. [Google Scholar] [CrossRef] [Green Version]
  25. Husain, S.; Verma, S.K.; Yasin, D.; Hemlata, F.; Rizvi, M.M.A.; Fatma, T. Facile green Bio-Fabricated Silver Nanoparticles from Microchaete Infer Dose-dependent Antioxidant and Anti-proliferative Activity to Mediate Cellular Apoptosis. Bioorg. Chem. 2021, 107, 104535. [Google Scholar] [CrossRef] [PubMed]
  26. Zhao, G.J.; Stevens, S.E. Multiple parameters for the comprehensive evaluation of the susceptibility of Escherichia coli to the silver ion. Biometals 1998, 11, 27–32. [Google Scholar] [CrossRef]
  27. Vaidyanathan, R.; Gopalram, S.; Kalishwaralal, K.; Deepak, V.; Pandian, S.R.; Gurunathan, S. Enhanced silver nanoparticle synthesis by optimization of nitrate reductase activity. Colloids Surf. B 2010, 75, 335–341. [Google Scholar] [CrossRef]
  28. Patil, M.P.; Singh, R.D.; Koli, P.B.; Patil, K.T.; Jagadale, B.S.; Tipare, A.R.; Kim, G.D. Antibacterial potential of silver nanoparticles synthesized using Madhuca longifolia flower extract as a green resource. Microb. Pathog. 2018, 121, 184–189. [Google Scholar] [CrossRef]
  29. Alsamhary, K.I. Eco-friendly synthesis of silver nanoparticles by Bacillus subtilis and their antibacterial activity. Saudi J. Biol. Sci. 2020, 27, 2185–2191. [Google Scholar] [CrossRef]
  30. Tyagi, S.; Tyagi, P.K.; Gola, D.; Chauhan, N.; Bharti, R.K. Extracellular synthesis of silver nanoparticles using entomopathogenic fungus: Characterization and antibacterial potential. SN Appl. Sci. 2019, 1, 1545. [Google Scholar] [CrossRef] [Green Version]
  31. Arya, A.; Mishra, V.; Chundawat, T.S. Green synthesis of silver nanoparticles from green algae (Botryococcus braunii) and its catalytic behavior for the synthesis of benzimidazoles. Chem. Data Collect. 2019, 20, 100190. [Google Scholar] [CrossRef]
  32. West, P.R.; Ishii, S.; Naik, G.V.; Emani, N.K.; Shalaev, V.M.; Boltasseva, A. Searching for Better Plasmonic Materials Paul. Laser Photonics Rev. 2010, 4, 795–808. [Google Scholar] [CrossRef] [Green Version]
  33. Brayner, R.; Hélène, B.; Miryana, H.; Chakib, D.; Claude, Y.; Thibaud, C.; Jacques, L.; Fernand, F.; Alain, C. Cyanobacteria as bioreactors for the synthesis of Au, Ag, Pd, and Pt nanoparticles via an enzyme-mediated route. J. Nanosci. Nanotechnol. 2007, 7, 2696–2708. [Google Scholar] [CrossRef]
  34. Shiny, P.; Mukherjee, A.; Chandrasekaran, N. DNA damage and mitochondria-mediated apoptosis of A549 lung carcinoma cells induced by rnatechzati silver and platinum nanoparticles. RSC Adv. 2016, 6, 27775–27787. [Google Scholar] [CrossRef]
  35. Alshatwi, A.A.; Athinarayanan, J.; Subbarayan, P.V. Green synthesis of platinum nanoparticles that induce cell death and G2/M-phase cell cycle arrest in human cervical cancer cells. J. Mater. Sci. Mater. Med. 2015, 26, 7. [Google Scholar] [CrossRef] [PubMed]
  36. Leo, A.J.; Oluwafemi, O.S. Plant-mediated synthesis of platinum nanoparticles using water hyacinth as an efficient biomatrix source—An eco-friendly development. Mater. Lett. 2017, 196, 141–144. [Google Scholar] [CrossRef]
  37. Nadaroglu, H.; Gungor, A.A.; Ince, S.; Babagil, A. Green synthesis and rnatechzation of platinum nanoparticles using quail egg yolk. Spectrochim. Acta Part A Mol. Biomol. Spectrosc. 2017, 172, 43–47. [Google Scholar] [CrossRef] [PubMed]
  38. Mikhailova, E.O. Green Synthesis of Platinum Nanoparticles for Biomedical Applications. J. Funct. Biomater. 2022, 13, 260. [Google Scholar] [CrossRef] [PubMed]
  39. Jameel, M.S.; Aziz, A.A.; Dheyab, M.A. Green synthesis: Proposed mechanism and factors influencing the synthesis of platinum nanoparticles. Green Process. Synth. 2020, 9, 386–398. [Google Scholar] [CrossRef]
  40. Fahmy, S.A.; Preis, E.; Bakowsky, U.; El-Said Azzazy, H.M. Palladium Nanoparticles Fabricated by Green Chemistry: Promising Chemotherapeutic, Antioxidant and Antimicrobial Agents. Materials 2020, 13, 3661. [Google Scholar] [CrossRef] [PubMed]
  41. Nasrollahzadeh, M.; Sajadi, S.M.; Maham, M. Green synthesis of palladium nanoparticles using Hippophae rhamnoides Linn leaf extract and their catalytic activity for the Suzuki-Miyaura coupling in water. J. Mol. Catal. A Chem. 2015, 396, 297–303. [Google Scholar] [CrossRef]
  42. Bankar, A.; Joshi, B.; Kumar, A.R.; Zinjarde, S. Banana peel extract mediated novel route for the synthesis of palladium nanoparticles. Mater. Lett. 2010, 64, 1951–1953. [Google Scholar] [CrossRef]
  43. Khazaei, A.; Rahmati, S.; Hekmatian, Z.; Saeednia, S. A green approach for the synthesis of palladium nanoparticles supported on pectin: Application as a catalyst for solvent-free Mizoroki-Heck reaction. J. Mol. Catal. A Chem. 2013, 372, 160–166. [Google Scholar] [CrossRef]
  44. Hemmati, S.; Mehrazin, L.; Ghorban, H.; Garakani, S.H.; Mobaraki, T.H.; Mohammadia, P.; Veisi, H. Green synthesis of Pd nanoparticles supported on reduced graphene oxide, using the extract of Rosa canina fruit, and their use as recyclable and heterogeneous nanocatalysts for the degradation of dye pollutants in water. RSC Adv. 2018, 8, 21020. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  45. Khan, M.; Albalawi, G.H.; Shaik, M.R.; Khan, M.; Adil, S.F.; Kuniyil, M.; Alkhathlan, H.Z.; Al-Warthan, A.; Siddiqui, M.R.H. Miswak mediated green synthesized palladium nanoparticles as effective catalysts for the Suzuki coupling reactions in aqueous media. J. Saudi Chem. Soc. 2017, 21, 450–457. [Google Scholar] [CrossRef] [Green Version]
  46. Kora, A.J.; Rastogi, L. Green synthesis of palladium nanoparticles using gum ghatti (Anogeissus latifolia) and its application as an antioxidant and catalyst. Arab. J. Chem. 2018, 11, 1097–1106. [Google Scholar] [CrossRef] [Green Version]
  47. Ko, Y.L.; Krishnamurthy, S.; Yun, Y.S. Facile synthesis of monodisperse Pt and Pd nanoparticles using antioxidants. J. Nanosci. Nanotechnol. 2015, 15, 412–417. [Google Scholar] [CrossRef]
  48. Lee, H.-J.; Lee, G.; Jang, N.R.; Yun, J.H.; Song, J.Y.; Kim, B.S. Biological Synthesis of Copper Nanoparticles Using Plant Extract. Nanotechnology 2011, 1, 371–374. [Google Scholar]
  49. Amjad, R.; Mubeen, B.; Ali, S.S.; Imam, S.S.; Alshehri, S.; Ghoneim, M.M.; Alzarea, S.I.; Rasool, R.; Ullah, I.; Nadeem, M.S.; et al. Green Synthesis and Characterization of Copper Nanoparticles Using Fortunella margarita Leaves. Polymers 2021, 13, 4364. [Google Scholar] [CrossRef]
  50. Jayandran, M.; Haneefa, M.M.; Balasubramanian, V. Green synthesis of copper nanoparticles using natural reducer and stabilizer and an evaluation of antimicrobial activity. J. Chem. Pharm. Res. 2015, 7, 251–259. [Google Scholar]
  51. Koopi, H.; Buazar, F. A novel one-pot biosynthesis of pure alpha aluminum oxide nanoparticles using the macroalgae Sargassum ilicifolium: A green marine approach. Ceram. Int. 2018, 44, 8940–8945. [Google Scholar] [CrossRef]
  52. Sepahvand, M.; Buazar, F.; Sayahi, M.H. Novel marine-based gold nanocatalyst in solvent-free synthesis of polyhydroquinoline derivatives: Green and sustainable protocol. Appl. Organomet. Chem. 2020, 34, e6000. [Google Scholar] [CrossRef]
  53. Rezazadeh, N.H.; Buazar, F.; Matroodi, S. Synergistic effects of combinatorial chitosan and polyphenol biomolecules on enhanced antibacterial activity of biofunctionalized silver nanoparticles. Sci. Rep. 2020, 10, 19615. [Google Scholar] [CrossRef] [PubMed]
  54. Buazar, F.; Sweidi, S.; Badri, M.; Kroushawi, F. Biofabrication of highly pure copper oxide nanoparticles using wheat seed extract and their catalytic activity: A mechanistic approach. Green Process. Synth. 2019, 8, 691–702. [Google Scholar] [CrossRef]
  55. Ying, S.; Guan, Z.; Ofoegbu, P.C.; Clubb, P.; Rico, C.; He, F.; Hong, J. Green synthesis of nanoparticles: Current developments and limitations. Environ. Technol. Innov. 2022, 26, 102336. [Google Scholar] [CrossRef]
  56. Rajesh, K.M.; Ajitha, B.; Reddy, Y.A.K.; Suneetha, Y.; Reddy, P.S. Assisted green synthesis of copper nanoparticles using Syzygium aromaticum bud extract: Physical, optical and antimicrobial properties. Optik 2018, 154, 593–600. [Google Scholar] [CrossRef]
  57. Perreault, F.; Melegari, S.P.; da Costa, C.H.; de Oliveira Franco Rossetto, A.L.; Popovic, R.; Matias, W.G. Genotoxic effects of copper oxide nanoparticles in Neuro 2A cell cultures. Sci. Total Environ. 2012, 441, 117–124. [Google Scholar] [CrossRef]
  58. Alhalili, Z. Green synthesis of copper oxide nanoparticles CuO NPs from Eucalyptus globoulus leaf extract: Adsorption and design of experiments. Arab. J. Chem. 2022, 15, 103739. [Google Scholar] [CrossRef]
  59. El Din Mahmoud, A.; Al-Qahtani, K.M.; Alflaij, S.O.; Al-Qahtani, S.F.; Alsamhan, F.A. Green copper oxide nanoparticles for lead, nickel, and cadmium removal from contaminated water. Sci. Rep. 2021, 11, 12547. [Google Scholar] [CrossRef]
  60. Bhattacharya, P.; Swarnakar, S.; Ghosh, S.; Majumdar, S.; Banerjee, S. Disinfection of drinking water via algae mediated green synthesized copper oxide nanoparticles and its toxicity evaluation. J. Environ. Chem. Eng. 2019, 7, 102867. [Google Scholar] [CrossRef]
  61. Singh, D.; Jain, D.; Rajpurohit, D.; Jat, G.; Kushwaha, H.S.; Singh, A.; Mohanty, S.R.; Al-Sadoon, M.K.; Zaman, W.; Upadhyay, S.K. Bacteria assisted green synthesis of copper oxide nanoparticles and their potential applications as antimicrobial agents and plant growth stimulants. Front. Chem. 2023, 11, 1154128. [Google Scholar] [CrossRef] [PubMed]
  62. Wang, Y.; Aker, W.G.; Hwang, H.-M.; Yedjou, C.G.; Yu, H.; Tchounwou, P.B. A study of the mechanism of in vitro cytotoxicity of metal oxide nanoparticles using catfish primary hepatocytes and human HepG2 cells. Sci. Total Environ. 2011, 409, 4753–4762. [Google Scholar] [CrossRef] [Green Version]
  63. Kim, I.; Viswanathan, K.; Kasi, G.; Thanakkasaranee, S.; Sadeghi, K.; Seo, J. ZnO Nanostructures in Active Antibacterial Food Packaging: Preparation Methods, Antimicrobial Mechanisms, Safety Issues, Future Prospects, and Challenges. Food Rev. Int. 2020, 38, 537–565. [Google Scholar] [CrossRef] [Green Version]
  64. Sonia, S.; Linda Jeeva Kumari, H.; Ruckmanib, K.; Sivakumara, 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 Mater. Biol. Appl. 2017, 79, 581–589. [Google Scholar] [CrossRef]
  65. Basnet, P.; Inakhunbi Chanu, T.; Samanta, D.; Chatterjee, S. A Review on Bio-Synthesized Zinc Oxide Nanoparticles Using Plant Extracts as Reductants and Stabilizing Agents. J. Photochem. Photobiol. B 2018, 183, 201–221. [Google Scholar] [CrossRef] [PubMed]
  66. Nguyen, V.T.; Vu, V.T.; Nguyen, T.H.; Nguyen, T.A.; Tran, V.K.; Nguyen-Tri, P. Antibacterial Activity of TiO2- and ZnO-Decorated with Silver Nanoparticles. J. Compos. Sci. 2019, 3, 61. [Google Scholar] [CrossRef] [Green Version]
  67. Yusof, N.A.A.; Zain, N.M.; Pauzi, N. Synthesis of ZnO nanoparticles with chitosan as stabilizing agent and their antibacterial properties against Gram-positive and Gram-negative bacteria. Int. J. Biol. Macromol. 2019, 124, 1132–1136. [Google Scholar] [CrossRef]
  68. Rad, S.S.; Sani, A.M.; Mohseni, S. Biosynthesis, characterization and antimicrobial activities of zinc oxide nanoparticles from leaf extract of Mentha pulegium (L.). Microb. Pathog. 2019, 131, 239–245. [Google Scholar] [CrossRef]
  69. 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] [Green Version]
  70. Sumanth, B.; Lakshmeesha, T.R.; Ansari, M.A.; Alzohairy, M.A.; Udayashankar, A.C.; Shobha, B.; Niranjana, S.R.; Srinivas, C.; Almatroudi, A. Mycogenic synthesis of extracellular zinc oxide nanoparticles from Xylaria acuta and its nanoantibiotic potential. Int. J. Nanomed. 2020, 15, 8519. [Google Scholar] [CrossRef]
  71. Sanaeimehr, Z.; Javadi, I.; Namvar, F. Antiangiogenic and antiapoptotic effects of green-synthesized zinc oxide nanoparticles using Sargassum muticum algae extraction. Cancer Nanotechnol. 2018, 9, 3. [Google Scholar] [CrossRef] [PubMed]
  72. Naseer, M.; Aslam, U.; Khalid, B.; Chen, B. Green route to synthesize Zinc Oxide Nanoparticles using leaf extracts of Cassia fistula and Melia rnatech and their antibacterial potential. Sci. Rep. 2020, 10, 9055. [Google Scholar] [CrossRef] [PubMed]
  73. Moreira, A.F.; Rodrigues, C.F.; Reis, C.A.; Costa, E.C.; Correia, I.J. Gold-core silica shell nanoparticles application in imaging and therapy: A review. Microporous Mesoporous Mater. 2018, 270, 168–179. [Google Scholar] [CrossRef]
  74. Tong, S.; Zhu, H.; Bao, G. Magnetic iron oxide nanoparticles for disease detection and therapy. Mater. Today 2019, 31, 86–99. [Google Scholar] [CrossRef]
  75. Gokila, V.; Perarasu, V.; Rufina, R.D.J. Qualitative comparison of chemical and green synthesized Fe3O4 nanoparticles. Adv. Nano Res. 2021, 10, 71–76. [Google Scholar] [CrossRef]
  76. Senarathna, U.; Fernando, S.S.; Gunasekara, T.D.; Weerasekera, M.M.; Hewageegana, H.G.; Arachchi, N.D.; Siriwardena, H.D.; Jayaweera, P.M. Enhanced antibacterial activity of TiO2 nanoparticle surface modified with Garcinia zeylanica extract. Chem. Cent. J. 2017, 11, 7. [Google Scholar] [CrossRef] [Green Version]
  77. Verma, V.; Al-Dossari, M.; Singh, J.; Rawat, M.; Kordy, M.G.M.; Shaban, M. A Review on Green Synthesis of TiO2 NPs: Photocatalysis and Antimicrobial Applications. Polymers 2022, 14, 1444. [Google Scholar] [CrossRef]
  78. Maheswari, P. Bio-modified TiO2 nanoparticles with Withania somnifera, Eclipta prostrata and Glycyrrhiza glabra for anticancer and antibacterial applications. Mater. Sci. Eng. C Mater. Biol. Appl. 2020, 108, 110457. [Google Scholar] [CrossRef]
  79. Ikram, M.; Javed, B.; Hassan, S.W.U.; Satti, S.H.; Sarwer, A.; Raja, N.I.; Mashwani, Z.-R. Therapeutic potential of biogenic titanium dioxide nanoparticles: A review on mechanistic approaches. Nanomedicine 2021, 16, 1429–1446. [Google Scholar] [CrossRef]
  80. Charbgoo, F.; Ahmad, M.B.; Darroudi, M. Cerium oxide nanoparticles: Green synthesis and biological applications. Int. J. Nanomed. 2017, 12, 1401–1413. [Google Scholar] [CrossRef] [Green Version]
  81. Nadeem, M.; Khan, R.; Afridi, K.; Nadhman, A.; Ullah, S.; Faisal, S.; Ul Mabood, Z.; Hano, C.; Abbasi, B.H. Green Synthesis of Cerium Oxide Nanoparticles (CeO2 NPs) and Their Antimicrobial Applications: A Review. Int. J. Nanomed. 2020, 15, 5951–5961. [Google Scholar] [CrossRef] [PubMed]
  82. Nezhad, S.; Haghi, A.; Homayouni, M. Green synthesis of cerium oxide nanoparticle using Origanum majorana L. leaf extract, its characterization and biological activities: Green synthesis of nanoparticle. Appl. Organomet. Chem. 2019, 34, e5314. [Google Scholar] [CrossRef]
  83. Krishnaveni, M.P.; Priya, M.L.; Annadurai, G. Biosynthesis of nanoceria from bacillus subtilis: Characterization and antioxidant potential. Res. J. Life Sci. Bioinform. Pharmac. Chem. Sci. 2019, 5, 632–644. [Google Scholar] [CrossRef]
  84. Sugumaran, V.K.; Gopinath, K.; Palani, N.S.; Arumugam, A.; Jose, S.P.; Bahadura, S.A.; Ilangovan, R. Plant pathogenic fungus F. solani mediated biosynthesis of Nanoceria: Antibacterial and antibiofilm activity. RSC Adv. 2016, 6, 42720–42729. [Google Scholar] [CrossRef]
  85. Rosi, H.; Ethrajavalli, R.; Janci, M.I. Synthesis of Cerium Oxide Nanoparticles Using Marine Algae Sargassum wightii Greville Extract: Implications for Antioxidant Applications. In Proceedings of the 2020 International Conference on System, Computation, Automation and Networking (ICSCAN), Pondicherry, India, 3–4 July 2020; pp. 1–3. [Google Scholar] [CrossRef]
  86. Patil, S.N.; Paradeshi, J.S.; Chaudhari, P.B.; Mishra, S.J.; Chaudhari, B.L. Bio-therapeutic potential and cytotoxicity assessment of pectin-mediated synthesized nanostructured cerium oxide. Appl. Biochem. Biotechnol. 2016, 180, 638–654. [Google Scholar] [CrossRef] [PubMed]
  87. Mughal, B.; Zohaib, S.; Zaidi, J.; Zhang, X.; Ul Hassan, S. Biogenic Nanoparticles: Synthesis, Characterisation and Applications. Appl. Sci. 2021, 11, 2598. [Google Scholar] [CrossRef]
  88. Abdel-Raouf, N.; Al-Enazi, N.M.; Ibraheem, I.B.M.; Alharbi, R.M.; Alkhulaifi, M.M. Biosynthesis of silver nanoparticles by using of the marine brown alga Padina pavonia and their characterization. Saudi J. Biol. Sci. 2019, 26, 1207–1215. [Google Scholar] [CrossRef]
  89. Nasrollahzadeh, M.; Atarod, M.; Sajjadi, M.; Sajadi, S.M.; Issaabadi, Z. Plant mediated green synthesis of nanostructures: Mechanisms, characterization, and applications. Interface Sci. Technol. 2019, 28, 199–322. [Google Scholar] [CrossRef]
  90. Karunakaran, G.; Jagathambal, M.; Gusev, A.; Torres, J.A.L.; Kolesnikov, E.; Kuznetsov, D. Rapid Biosynthesis of AgNPs Using Soil Bacterium Azotobacter vinelandii With Promising Antioxidant and Antibacterial Activities for Biomedical Applications. JOM 2017, 69, 1206–1212. [Google Scholar] [CrossRef]
  91. Singh, A.; Gautam, P.K.; Verma, A.; Singh, V.; Shivapriya, P.M.; Shivalkar, S.; Sahoo, A.K.; Samanta, S.K. Green synthesis of metallic nanoparticles as effective alternatives to treat antibiotics resistant bacterial infections: A review. Biotechnol. Rep. 2020, 25, e00427. [Google Scholar] [CrossRef]
  92. Ghosh, S.; Ahmad, R.; Banerjee, K.; AlAjmi, M.F.; Rahman, S. Mechanistic Aspects of Microbe-Mediated Nanoparticle Synthesis. Front. Microbiol. 2021, 12, 638068. [Google Scholar] [CrossRef] [PubMed]
  93. Küünal, S.; Rauwel, P.; Rauwel, E. Plant extract mediated synthesis of nanoparticles. In Emerging Applications of Nanoparticles and Architecture Nanostructures; Barhoum, A., Makhlouf, A.S.H., Eds.; Elsevier: Amsterdam, The Netherlands, 2018; pp. 411–446. [Google Scholar] [CrossRef]
  94. Nasrollahzadeh, M.; Mahmoudi-Gom Yek, S.; Motahharifar, N.; Ghafori Gorab, M. Recent developments in the plant-mediated green synthesis of Ag-Based nanoparticles for environmental and catalytic applications. Chem. Rec. 2019, 19, 2436–2479. [Google Scholar] [CrossRef] [PubMed]
  95. Nasrollahzadeh, M.; Sajjadi, M.; Dadashi, J.; Ghafuri, H. Pd-based nanoparticles: Plant-assisted biosynthesis, characterization, mechanism, stability, catalytic and antimicrobial activities. Adv. Colloid Interface Sci. 2020, 276, 102103. [Google Scholar] [CrossRef] [PubMed]
  96. Dwivedi, A.D.; Gopal, K. Biosynthesis of silver and gold nanoparticles using Chenopodium album leaf extract. Colloids Surf. A Physicochem. Eng. Asp. 2010, 369, 27–33. [Google Scholar] [CrossRef]
  97. Rokade, S.S.; Joshi, K.A.; Mahajan, K.; Tomar, G.; Dubal, D.S.; Singh, V.; Kitture, R.; Bellare, J.; Ghosh, S. Novel anticancer platinum and palladium nanoparticles from Barleria prionitis. Micro Nano Lett. 2017, 2, 555600. [Google Scholar] [CrossRef]
  98. Yang, N.; Li, W. Mango peel extract mediated novel route for synthesis of silver nanoparticles and antibacterial application of silver nanoparticles loaded onto non-woven fabrics. Ind. Crops Prod. 2013, 48, 81–88. [Google Scholar] [CrossRef]
  99. Ndikau, M.; Noah, N.M.; Andala, D.M.; Masika, E. Green synthesis and characterization of silver nanoparticles using Citrullus lanatus fruit rind extract. Int. J. Anal. Chem. 2017, 2017, 8108504. [Google Scholar] [CrossRef] [Green Version]
  100. Patil, S.; Chandrasekaran, R. Biogenic nanoparticles: A comprehensive perspective in synthesis, characterization, application and its challenges. J. Genet. Eng. Biotechnol. 2020, 18, 67. [Google Scholar] [CrossRef]
  101. Jacob, J.; Mukherjee, T.; Kapoor, S. A simple approach for facile synthesis of Ag, anisotropic Au and bimetallic (Ag/Au) nanoparticles using cruciferous vegetable extracts. Mater. Sci. Eng. C 2012, 32, 1827–1834. [Google Scholar] [CrossRef]
  102. Das, D.R.K.; Pachapur, V.; Lonappan, L.; Naghdi, M.; Rama, P.; Maiti, S.; Cledón, M.; Larios, A.; Sarma, S.; Brar, S. Biological synthesis of metallic nanoparticles: Plants, animals and microbial aspects Nanotechnol. Environ. Eng. 2017, 2, 18. [Google Scholar] [CrossRef] [Green Version]
  103. Ahmad, T.; Iqbal, J.; Bustam, M.A.; Irfan, M.; Asghar, H.M.A. A critical review on phytosynthesis of gold nanoparticles: Issues, challenges and future perspectives. J. Clean. Prod. 2021, 309, 127460. [Google Scholar] [CrossRef]
  104. Netai, M.M.; Stephen, N.; Musekiwa, C. Synthesis of silver nanoparticles using wild Cucumis anguria: Characterization and antibacterial activity. Afr. J. Biotechnol. 2017, 16, 1911–1921. [Google Scholar] [CrossRef] [Green Version]
  105. El-Desouky, N.; Shoueir, K.; El-Mehasseb, I.; El-Kemary, M. Bio-Inspired Green Manufacturing of Plasmonic Silver Nanoparticles/Degussa using Banana Waste Peduncles: Photocatalytic, Antimicrobial, and Cytotoxicity Evaluation. J. Mater. Res. Technol. 2021, 10, 671–686. [Google Scholar] [CrossRef]
  106. Perveen, K.; Husain, F.M.; Qais, F.A.; Khan, A.; Razak, S.; Afsar, T.; Alam, P.; Almajwal, A.M.; Abulmeaty, M. Microwave-assisted rapid green Synthesis of gold nanoparticles using seed extract of Trachyspermum ammi: ROS mediated biofilm inhibition and anticancer activity. Biomolecules 2021, 11, 197. [Google Scholar] [CrossRef] [PubMed]
  107. Elumalai Abel, E.; Poonga, P.R.J.; Panicker, S.G. Characterization and in vitro studies on anticancer, antioxidant activity against colon cancer cell line of gold nanoparticles capped with Cassia tora SM leaf extract. Appl. Nanosci. 2016, 6, 121–129. [Google Scholar] [CrossRef] [Green Version]
  108. Amina, M.; Al Musayeib, N.M.; Alarfaj, N.A.; El-Tohamy, M.F.; Al-Hamoud, G.A. Antibacterial and immunomodulatory potentials of biosynthesized Ag, Au, AgAu bimetallic alloy nanoparticles using the Asparagus racemosus root extract. Nanomaterials 2020, 10, 2453. [Google Scholar] [CrossRef]
  109. Al Qhtani, M.S.; El-Debaiky, S.A.; Sayed, M. Antifungal and cytotoxic activities of biosynthesized silver, zinc and gold nanoparticles by flower extract of Rhanterium epapposum. Open J. Appl. Sci. 2020, 10, 663. [Google Scholar] [CrossRef]
  110. Rabiee, N.; Bagherzadeh, M.; Kiani, M.; Ghadiri, A.M. Rosmarinus officinalis directed palladium nanoparticle synthesis: Investigation of potential antibacterial, antifungal and Mizoroki-Heck catalytic activities. Adv. Powder Technol. 2020, 31, 1402–1411. [Google Scholar] [CrossRef]
  111. Vinodhini, S.; Vithiya, B.S.M.; Prasad, T.A.A. Green synthesis of palladium nanoparticles using aqueous plant extracts and its biomedical applications. J. King Saud Univ. Sci. 2022, 34, 102017. [Google Scholar] [CrossRef]
  112. Al-Radadi, N.S. Green synthesis of platinum nanoparticles using Saudi’s dates extract and their sage on the cancer cell treatment. Arab. J. Chem. 2019, 12, 330–349. [Google Scholar] [CrossRef]
  113. Mali, S.C.; Dhaka, A.; Githala, C.K.; Trivedi, R. Green synthesis of copper nanoparticles using Celastrus paniculatus Willd. Leaf extract and their photocatalytic and antifungal properties. Biotechnol. Rep. 2020, 27, e00518. [Google Scholar] [CrossRef] [PubMed]
  114. Wu, S.; Rajeshkumar, S.; Madasamy, M.; Mahendran, V. Green synthesis of copper nanoparticles using Cissus vitiginea and its antioxidant and antibacterial activity against urinary tract infection pathogens. Artif. Cells Nanomed. Biotechnol. 2020, 48, 1153–1158. [Google Scholar] [CrossRef] [PubMed]
  115. Dey, A.; Manna, S.; Chattopadhyay, S.; Mondal, D.; Chattopadhyay, D.; Raj, A.; Das, S.; Bag, B.G.; Roy, S. Azadirachta indica leaves mediated green synthesized copper oxide nanoparticles induce apoptosis through activation of TNF-α and caspases signaling pathway against cancer cells. J. Saudi Chem. Soc. 2019, 23, 222–238. [Google Scholar] [CrossRef]
  116. Rabiee, N.; Bagherzadeh, M.; Kiani, M.; Ghadiri, A.M.; Etessamifar, F.; Jaberizadeh, A.H.; Shakeri, A. Biosynthesis of Copper Oxide Nanoparticles with Potential Biomedical Applications. Int. J. Nanomed. 2020, 15, 3983–3999. [Google Scholar] [CrossRef] [PubMed]
  117. Saranya, S. Green synthesis of iron nanoparticles using aqueous extract of Musa rnate flower sheath against pathogenic bacteria. Indian J. Pharm. Sci. 2017, 79, 688–694. [Google Scholar] [CrossRef]
  118. Üstün, E.; Önbas, S.C.; Çelik, S.K.; Ayvaz, M.Ç.; Sahin, N. Green Synthesis of Iron Oxide Nanoparticles by Using Ficus Carica Leaf Extract and Its Antioxidant Activity. Biointerface Res. Appl. Chem. 2022, 2022, 2108–2116. [Google Scholar] [CrossRef]
  119. Devi, H.S.; Boda, M.A.; Shah, M.A.; Parveen, S. Green synthesis of iron oxide nanoparticles using Platanus orientalis leaf extract for antifungal activity. Green Process. Synth. 2019, 8, 38–45. [Google Scholar] [CrossRef]
  120. Ali, J.; Irshad, R.; Li, B.; Tahir, K. Synthesis and characterization of phytochemical fabricated zinc oxide nanoparticles with enhanced antibacterial and catalytic applications. J. Photochem. Photobiol. B Biol. 2018, 183, 349–356. [Google Scholar] [CrossRef]
  121. Suresh, J.; Pradheesh, G.; Alexramani, V. Green synthesis and characterization of zinc oxide nanoparticle using insulin plant (Costus pictus D. Don) and investigation of its antimicrobial as well as anticancer activities. Adv. Nat. Sci. Nanosci. Nanotechnol. 2018, 9, 015008. [Google Scholar] [CrossRef]
  122. Umar, H.; Kavaz, D.; Rizaner, N. Biosynthesisof zinc oxide nanoparticles using Albizia lebbeck stem bark, and evaluation of its antimicrobial, antioxidant, and cytotoxic activities on human breast cancer cell lines. Int. J. Nanomed. 2019, 14, 87–100. [Google Scholar] [CrossRef] [Green Version]
  123. Awan, S.S.; Khan, R.T.; Mehmood, A.; Hafeez, M.; Abass, S.R.; Nazir, M.; Raffi, M. Ailanthus altissima leaf extract mediated green production of zinc oxide (ZnO) nanoparticles for antibacterial and antioxidant activity. Saudi J. Biol. Sci. 2023, 30, 103487. [Google Scholar] [CrossRef]
  124. 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]
  125. Vergheese, M.; Vishal, S.K. Green synthesis of magnesium oxide nanoparticles using Trigonella foenum-graecum leaf extract and its antibacterial activity. J. Pharmacogn. Phytochem. 2018, 7, 1193–1200. [Google Scholar]
  126. Javadi, F.; Yazdi, M.E.T.; Baghani, M.; Es-haghi, A. Biosynthesis, characterization of cerium oxide nanoparticles using Ceratonia siliqua and evaluation of antioxidant and cytotoxicity activities. Mater. Res. Express 2019, 6, 065408. [Google Scholar] [CrossRef]
  127. Thakkar, K.N.; Mhatre, S.S.; Parikh, R.Y. Biological synthesis of metallic nanoparticles. Nanomed. Nanotechnol. Biol. Med. 2010, 6, 257–262. [Google Scholar] [CrossRef]
  128. Nasrollahzadeh, M.; Sajjadi, M.; Iravani, S.; Varmac, R.S. Green-synthesized nanocatalysts and nanomaterials for water treatment: Current challenges and future perspectives. J. Hazard. Mater. 2021, 401, 123401. [Google Scholar] [CrossRef] [PubMed]
  129. Banerjee, A.; Halder, U.; Bandopadhyay, R. Preparations and applications of polysaccharide based green synthesized metal nanoparticles: A state-of-the-art. J. Clust. Sci. 2017, 28, 1803–1813. [Google Scholar] [CrossRef]
  130. Parandhaman, T.; Dey, M.D.; Das, S.K. Biofabrication of supported metal nanoparticles: Exploring the bioinspiration strategy to mitigate the environmental challenges. Green Chem. 2019, 21, 5469–5500. [Google Scholar] [CrossRef]
  131. Durán, M.; Silveira, C.P.; Durán, N. Catalytic role of traditional enzymes for biosynthesis of biogenic metallic nanoparticles: A mini-review. IET Nanobiotechnol. 2015, 9, 314–323. [Google Scholar] [CrossRef] [PubMed]
  132. Shi, L.; Rosso, K.M.; Clarke, T.A.; Richardson, D.J.; Zachara, J.M.; Fredrickson, J.K. Molecular underpinnings of Fe(III) oxide reduction by Shewanella oneidensis MR-1. Front. Microbiol. 2012, 3, 50. [Google Scholar] [CrossRef] [Green Version]
  133. Kang, F.; Alvarez, P.J.; Zhu, D. Microbial Extracellular Polymeric Substances Reduce Ag+ to Silver Nanoparticles and Antagonize Bactericidal Activity. Environ. Sci. Technol. 2013, 48, 316–322. [Google Scholar] [CrossRef]
  134. Yan, L.; Da, H.; Zhang, S.; López, V.M.; Wang, W. Bacterial magnetosome and its potential application. Microbiol. Res. 2017, 203, 19–28. [Google Scholar] [CrossRef]
  135. Boucher, M.; Geffroy, F.; Prévéral, S.; Bellanger, L.; Selingue, E.; Adryanczyk-Perrier, G.; Péan, M.; Lefèvre, C.T.; Pignol, D.; Ginet, N.; et al. Genetically tailored magnetosomes used as MRI probe for molecular imaging of brain tumor. Biomaterials 2017, 121, 167–178. [Google Scholar] [CrossRef]
  136. Vargas, G.; Cypriano, J.; Correa, T.; Leao, P.; Bazylinski, D.A.; Abreu, F. Applications of magnetotactic bacteria, magnetosomes and magnetosome crystals in biotechnology and nanotechnology: Mini-review. Molecules 2018, 23, 2438. [Google Scholar] [CrossRef] [Green Version]
  137. Vaseghi, Z.; Nematollahzadeh, A.; Tavakoli, O. Green methods for the synthesis of metal nanoparticles using biogenic reducing agents: A review. Rev. Chem. Eng. 2018, 34, 529. [Google Scholar] [CrossRef]
  138. Mukherjee, P.; Senapati, S.; Mandal, D.; Ahmad, A.; Khan, M.I.; Kumar, R.; Sastry, M. Extracellular synthesis of gold nanoparticles by the fungus Fusarium oxysporum. ChemBioChem 2002, 3, 461–463. [Google Scholar] [CrossRef] [PubMed]
  139. Lee, K.X.; Shameli, K.; Yew, Y.P.; Teow, S.Y.; Jahangirian, H.; Rafiee-Moghaddam, R.; Webster, T.J. Recent Developments in the Facile Bio-Synthesis of Gold Nanoparticles (AuNPs) and Their Biomedical Applications. Int. J. Nanomed. 2020, 15, 275–300. [Google Scholar] [CrossRef] [PubMed]
  140. Chen, Y.-L.; Tuan, H.-Y.; Tien, C.-W.; Lo, W.-H.; Liang, H.-C.; Hu, Y.-C. Augmented biosynthesis of cadmium sulfide nanoparticles by genetically engineered Escherichia coli. Biotechnol. Prog. 2009, 25, 1260–1266. [Google Scholar] [CrossRef] [PubMed]
  141. Singh, P.; Kim, Y.-J.; Zhang, D.; Yang, D.-C. Biological Synthesis of Nanoparticles from Plants and Microorganisms. Trends Biotechnol. 2016, 34, 588–599. [Google Scholar] [CrossRef]
  142. Narayanan, K.B.; Sakthivel, N. Synthesis and characterization of nano-gold composite using Cylindrocladium floridanum and its heterogeneous catalysis in the degradation of 4-nitrophenol. J. Hazard. Mater. 2011, 189, 519–525. [Google Scholar] [CrossRef]
  143. Wang, L.; Liu, X.; Lee, D.-J.; Tay, J.-H.; Zhang, Y.; Wan, C.-L.; Chen, X.-F. Recent advances on biosorption by aerobic granular sludge. J. Hazard. Mater. 2018, 357, 253–270. [Google Scholar] [CrossRef]
  144. Durán, N.; Marcato, P.D.; Alves, O.L.; De Souza, G.I.H.; Esposito, E. Mechanistic aspects of biosynthesis of silver nanoparticles by several Fusarium oxysporum strains. J. Nanobiotechnol. 2005, 3, 8. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  145. Prabhu, S.; Poulose, E.K. Silver nanoparticles: Mechanism of antimicrobial action, synthesis, medical applications, and toxicity effects. Int. Nano Lett. 2012, 2, 32. [Google Scholar] [CrossRef] [Green Version]
  146. Rajput, S.; Werezuk, R.; Lange, R.M.; McDermott, M.T. Fungal isolate optimized for biogenesis of silver nanoparticles with enhanced colloidal stability. Langmuir 2016, 32, 8688–8697. [Google Scholar] [CrossRef] [PubMed]
  147. El Domany, E.B.; Essam, T.M.; Ahmed, A.E.; Farghali, A.A. Biosynthesis physico-chemical optimization of gold nanoparticles as anti-cancer and synergetic antimicrobial activity using pleurotus ostreatus fungus. J. Appl. Pharm. Sci. 2018, 8, 119–128. [Google Scholar] [CrossRef] [Green Version]
  148. Šebesta, M.; Vojtková, H.; Cyprichová, V.; Ingle, A.P.; Urík, M.; Kolenčík, M. Mycosynthesis of Metal-Containing Nanoparticles—Fungal Metal Resistance and Mechanisms of Synthesis. Int. J. Mol. Sci. 2022, 23, 14084. [Google Scholar] [CrossRef]
  149. Fawcett, D.; Verduin, J.J.; Shah, M.; Sharma, S.B.; Poinern, G.E.J. A Review of Current Research into the Biogenic Synthesis of Metal and Metal Oxide Nanoparticles via Marine Algae and Seagrasses. J. Nanosci. 2017, 2017, 8013850. [Google Scholar] [CrossRef] [Green Version]
  150. Roychoudhury, A. Yeast-mediated Green Synthesis of Nanoparticles for Biological Applications. Indian J. Pharm. Biol. Res. 2020, 8, 26–31. [Google Scholar]
  151. Stegenga, R.W.; Al-Azawi, S.; Bandyopadhyay, D.; Bandyopadhyay, K. Biosynthesis of Gold Nanoparticles by Saccharomyces cerevisiae. FASEB J. 2011, 25, 726.8. [Google Scholar]
  152. Niknejad, F.; Nabili, M.; Daie Ghazvini, R.; Moazeni, M. Green synthesis of silver nanoparticles: Advantages of the yeast Saccharomyces cerevisiae model. Curr. Med. Mycol. 2015, 1, 17–24. [Google Scholar] [CrossRef] [Green Version]
  153. Zhang, X.; Qu, Y.; Shen, W.; Wang, J.; Li, H.; Zhang, Z.; Li, S.; Zhou, J. Biogenic synthesis of gold nanoparticles by yeast Magnusiomyces ingens lh-f1 for catalytic reduction of nitrophenols. Colloids Surf. A 2016, 497, 280–285. [Google Scholar] [CrossRef]
  154. Elahian, F.; Reiisi, S.; Shahidi, A.; Mirzaei, S.A. High-throughput bioaccumulation, biotransformation, and production of silver and selenium nanoparticles using genetically engineered Pichia pastoris. Nanomed. Nanotechnol. Biol. Med. 2017, 13, 853–861. [Google Scholar] [CrossRef] [PubMed]
  155. Sriramulu, M.; Sumathi, S. Biosynthesis of palladium nanoparticles using Saccharomyces cerevisiae extract and its photocatalytic degradation behaviour. Adv. Nat. Sci. Nanosci. Nanotechnol. 2018, 9, 025018. [Google Scholar] [CrossRef]
  156. Zhang, W.; Bao, S.; Fang, T. The neglected nano-specific toxicity of ZnO nanoparticles in the yeast Saccharomyces cerevisiae. Sci. Rep. 2016, 6, 24839. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  157. Peiris, M.; Gunasekara, T.; Jayaweera, P.M.; Fernando, S. TiO2 nanoparticles from baker’s yeast: A potent antimicrobial. J. Microbiol. Biotechnol. 2018, 28, 1664–1670. [Google Scholar] [CrossRef] [Green Version]
  158. Velusamy, P.; Venkat Kumar, G.; Jeyanthi, V.; Das, J.; Pachaiappan, R. Bio-Inspired Green Nanoparticles: Synthesis, Mechanism, and Antibacterial Application. Toxicol. Res. 2016, 32, 95–102. [Google Scholar] [CrossRef] [Green Version]
  159. Shedbalkar, U.; Singh, R.; Wadhwani, S.; Gaidhani, S.; Chopade, B. Microbial synthesis of gold nanoparticles: Current status and future prospects. Adv. Colloid Interface Sci. 2014, 209, 40–48. [Google Scholar] [CrossRef]
  160. Love, A.J.; Makarov, V.V.; Sinitsyna, O.V.; Shaw, J.; Yaminsky, I.V.; Kalinina, N.O.; Taliansky, M.E. A Genetically Modified Tobacco Mosaic Virus that can Produce Gold Nanoparticles from a Metal Salt Precursor. Front. Plant Sci. 2015, 6, 984. [Google Scholar] [CrossRef] [Green Version]
  161. Love, A.; Makarov, V.; Yaminsky, I.; Kalinina, N.; Taliansky, M. The use of tobacco mosaic virus and cowpea mosaic virus for the production of novel metal nanomaterials. Virology 2014, 449, 133–139. [Google Scholar] [CrossRef] [Green Version]
  162. Thangavelu, R.M.; Ganapathy, R.; Ramasamy, P.; Krishnan, K. Fabrication of virus metal hybrid nanomaterials: An ideal reference for bio semiconductor. Arab. J. Chem. 2020, 13, 2750–2765. [Google Scholar] [CrossRef]
  163. Ahiwale, S.S.; Bankar, A.V.; Tagunde, S.; Kapadnis, B.P. A Bacteriophage Mediated Gold Nanoparticles Synthesis and Their Anti-biofilm Activity. Indian J. Microbiol. 2017, 57, 188–194. [Google Scholar] [CrossRef] [PubMed]
  164. Zeng, Q.; Wen, H.; Wen, Q.; Chen, X.; Wang, Y.; Xuan, W.; Liang, J.; Wan, S. Cucumber mosaic virus as drug delivery vehicle for doxorubicin. Biomaterials 2013, 34, 4632–4642. [Google Scholar] [CrossRef] [PubMed]
  165. Everts, M.; Saini, V.; Leddon, J.L.; Kok, R.J.; Stoff-Khalili, M.; Preuss, M.A.; Millican, C.L.; Perlins, G.; Brown, J.M.; Bagaria, H.; et al. Covalently linked Au nanoparticles to a viral vector: Potential for combined photothermal and gene cancer therapy. Nano Lett. 2006, 6, 587–591. [Google Scholar] [CrossRef] [PubMed]
  166. Sharma, A.; Sharma, S.; Sharma, K.; Chetri, S.P.K.; Vashishtha, A.; Singh, P.; Kumar, R.; Rathi, B.; Agrawal, V. Algae as crucial organisms in advancing nanotechnology: A systematic review. J. Appl. Phycol. 2016, 28, 1759–1774. [Google Scholar] [CrossRef]
  167. Hamouda, R.; Hussein, M.; Abo-elmagd, R.; Bawazir, S. Synthesis and biological characterization of silver nanoparticles derived from the cyanobacterium Oscillatoria limnetica. Sci. Rep. 2019, 9, 13071. [Google Scholar] [CrossRef] [Green Version]
  168. Jacob, J.M.; Ravindran, R.; Narayanan, M.; Samuel, S.M.; Pugazhendhi, A.; Kumar, G. Microalgae: A prospective low cost green alternative for nanoparticle synthesis. Curr. Opin. Environ. Sci. Health 2021, 20, 100163. [Google Scholar] [CrossRef]
  169. Patel, V.; Berthold, D.; Puranik, P.; Gantar, M. Screening of cyanobacteria and microalgae for their ability to synthesize silver nanoparticles with antibacterial activity. Biotechnol. Rep. 2015, 5, 112–119. [Google Scholar] [CrossRef] [Green Version]
  170. Al-Katib, M.; Al-Shahri, Y.; Al-Niemi, A. Biosynthesis of silver nanoparticles by cyanobacterium Gloeocapsa sp. Int. J. Enhanc. Res. Sci. Technol. Eng. 2015, 4, 60–73. [Google Scholar]
  171. Rajeshkumar, S. Synthesis of Zinc oxide nanoparticles using algal formulation (Padina tetrastromatica and Turbinaria conoides) and their antibacterial activity against fish pathogens. Res. J. Biotechnol. 2018, 13, 15–19. [Google Scholar]
  172. Mubarak Ali, D.; Sasikala, M.; Gunasekaran, M.; Thajuddin, N. Biosynthesis and characterization of silver nanoparticles using marine cyanobacterium, Oscillatoria willei NTDM01. Dig. J. Nanomater. Biostruct. 2011, 6, 385–390. [Google Scholar]
  173. Khan, F.; Shahid, A.; Zhu, H.; Wang, N.; Javed, M.R.; Ahmad, N.; Xu, J.; Alam, M.A.; Mehmood, M.A. Prospects of algae-based green synthesis of nanoparticles for environmental applications. Chemosphere 2022, 293, 133571. [Google Scholar] [CrossRef] [PubMed]
  174. 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]
  175. Li, S.N.; Wang, R.; Ho, S.H. Algae-mediated biosystems for metallic nanoparticle production: From synthetic mechanisms to aquatic environmental applications. J. Hazard. Mater. 2021, 420, 126625. [Google Scholar] [CrossRef] [PubMed]
  176. Fakayode, O.J.; Oladipo, A.O.; Oluwafemi, O.S.; Songca, S.P. Biopolymer-mediated Green Synthesis of Noble Metal Nanostructures. In Recent Advances in Biopolymers; Perveen, F.K., Ed.; InTech: Houston, TX, USA, 2016. [Google Scholar] [CrossRef] [Green Version]
  177. Rie, V.J.; Thielemans, W. Cellulose–gold nanoparticle hybrid materials. Nanoscale 2017, 9, 8525–8554. [Google Scholar] [CrossRef]
  178. González, P.A.; Zamora-Justo, J.A.; Sotelo-López, A.; Vázquez-Martínez, G.R.; Balderas-López, J.A.; Muñoz-Diosdado, A.; Hernández, M.I. Gold nanoparticles with chitosan, N-acylated chitosan, and chitosanoligosaccharide as DNA carriers. Nanoscale Res. Lett. 2019, 14, 258. [Google Scholar] [CrossRef]
  179. Singh, A.; Hede, S.; Sastry, M. Spider Silk as an Active Scaffold in the Assembly of Gold Nanoparticles and Application of the Gold–Silk Bioconjugate in Vapor Sensing. Small 2007, 3, 466–473. [Google Scholar] [CrossRef]
  180. Muniandy, S.S.; Mohd Kaus, N.H.; Jiang, Z.T.; Altarawneh, M.; Lee, H.L. Green synthesis of mesoporous anatase TiO2 nanoparticles and their photocatalytic activities. RSC Adv. 2017, 7, 48083–48094. [Google Scholar] [CrossRef] [Green Version]
  181. Venkataramanan, N.S.; Matsui, K.; Kawanami, H.; Ikushima, Y. Green synthesis of titania nanowire composites on natural cellulose fibers. Green Chem. 2007, 9, 18–19. [Google Scholar] [CrossRef]
  182. Yan, J.; Wu, G.; Li, L.; Yu, A.; Sun, X.; Guan, N. Synthesis of uniform TiO2 nanoparticles with egg albumen proteins as novel biotemplate. J. Nanosci. Nanotechnol. 2010, 10, 5767–5775. [Google Scholar] [CrossRef]
  183. Mulmi, D.D.; Dahal, B.; Kim, H.-Y.; Nakarmi, M.L.; Panthi, G. Optical and photocatalytic properties of lysozyme mediated titanium dioxide nanoparticles. Optik 2018, 154, 769–776. [Google Scholar] [CrossRef]
  184. Choudhury, A.R.; Malhotra, A.; Bhattacharjee, P.; Prasad, G. Facile and rapid errula-regulated biomineralization of gold by pullulan and study of its thermodynamic parameters. Carbohydr. Polym. 2014, 106, 154–159. [Google Scholar] [CrossRef]
  185. Safat, S.; Buazar, F.; Albukhaty, S.; Matroodi, S. Enhanced sunlight photocatalytic activity and biosafety of marine-driven synthesized cerium oxide nanoparticles. Sci. Rep. 2021, 11, 14734. [Google Scholar] [CrossRef] [PubMed]
  186. Mukherjee, S.; Patra, C.R. Biologically synthesized metal nanoparticles: Recent advancement and future perspectives in cancer theranostics. Future Sci. OA 2017, 3, FSO203. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  187. Maddinedi, S.B.; Mandal, B.K.; Ranjan, S.; Dasgupta, N. Diastase assisted green synthesis of size-controllable gold nanoparticles. RSC Adv. 2015, 5, 26727–26733. [Google Scholar] [CrossRef]
  188. Otari, S.V.; Kumar, M.; Kim, I.-W.; Lee, J.H.; Lee, J.-K. Rapid, thermostable antimicrobial peptide-mediated synthesis gold nanoparticles as highly efficient charge trapping medium for sol-gel-derived thin film. Mater. Lett. 2017, 188, 375–378. [Google Scholar] [CrossRef]
  189. Arib, C.; Spadavecchia, J.; Chapelle, M.L. Enzyme mediated synthesis of hybrid polyedric gold nanoparticles. Sci. Rep. 2021, 11, 3208. [Google Scholar] [CrossRef]
  190. Patil, M.; Kang, M.-J.; Niyonizigiye, I.; Singh, A.; Kim, J.-O.; Seo, Y.B.; Kim, G.-D. Extracellular synthesis of gold nanoparticles using the marine bacterium Paracoccus haeundaensis BC74171T and evaluation of their antioxidant activity and antiproliferative effect on normal and cancer cell lines. Colloids Surf. B 2019, 183, 110455. [Google Scholar] [CrossRef]
  191. Markus, J.; Mathiyalagan, R.; Kim, Y.J.; Abbai, R.; Singh, P.; Ahn, S.; Perez, Z.E.J.; Hurh, J.; Yang, D.C. Intracellular synthesis of gold nanoparticles with antioxidant activity by probiotic Lactobacillus kimchicus DCY51T isolated from Korean kimchi. Enzym. Microb. Technol. 2016, 95, 85–93. [Google Scholar] [CrossRef]
  192. Yuan, Q.; Bomma, M.; Xiao, Z. Enhanced silver nanoparticle synthesis by Escherichia coli transformed with Candida albicans metallothionein gene. Materials 2019, 12, 4180. [Google Scholar] [CrossRef] [Green Version]
  193. Saravanan, M.; Barik, S.K.; MubarakAli, D.; Prakash, P.; Pugazhendhi, A. Synthesis of silver nanoparticles from Bacillus brevis (NCIM 2533) and their antibacterial activity against pathogenic bacteria. Microb. Pathog. 2018, 116, 221–226. [Google Scholar] [CrossRef]
  194. Gaidhani, S.V.; Yeshvekar, R.K.; Shedbalkar, U.U.; Bellare, J.H.; Chopade, B.A. Bio-reduction of hexachloroplatinic acid to platinum nanoparticles employing Acinetobacter calcoaceticus. Process Biochem. 2014, 49, 2313–2319. [Google Scholar] [CrossRef]
  195. Hassan, S.E.L.D.; Salem, S.S.; Fouda, A.; Awad, M.A.; El-Gamal, M.S.; Abdo, A.M. New approach for antimicrobial activity and bio-control of various pathogens by biosynthesized copper nanoparticles using endophytic actinomycetes. J. Radiat. Res. Appl. Sci. 2018, 11, 262–270. [Google Scholar] [CrossRef] [Green Version]
  196. 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]
  197. Jain, D.; Shivani, A.A.B.; Singh, H.; Daima, H.K.; Singh, M.; Mohanty, S.R.; Stephen, B.J.; Singh, A. Microbial fabrication of zinc oxide nanoparticles and evaluation of their antimicrobial and photocatalytic properties. Front. Chem. 2020, 8, 778. [Google Scholar] [CrossRef] [PubMed]
  198. 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] [PubMed] [Green Version]
  199. Ağçeli, G.K.; Hammachi, H.; Kodal, S.P.; Cihangir, N.; Aksu, Z. A novel approach to synthesize tio 2 nanoparticles: Biosynthesis by using Streptomyces sp. Hc1. J. Inorg. Organomet. Polym. Mater. 2020, 30, 3221–3229. [Google Scholar] [CrossRef]
  200. Hassan, S.E.-D.; Fouda, A.; Radwan, A.A.; Salem, S.S.; Barghoth, M.G.; Awad, M.A.; Abdo, A.M.; El-Gamal, M.S. Endophytic actinomycetes Streptomyces spp mediated biosynthesis of copper oxide nanoparticles as a promising tool for biotechnological applications. JBIC J. Biol. Inorg. Chem. 2019, 24, 377–393. [Google Scholar] [CrossRef] [PubMed]
  201. Bukhari, S.I.; Hamed, M.M.; Al-Agamy, M.H.; Gazwi, H.S.; Radwan, H.H.; Youssif, A.M. Biosynthesis of copper oxide nanoparticles using Streptomyces MHM38 and its biological applications. J. Nanomater. 2021, 2021, 6693302. [Google Scholar] [CrossRef]
  202. Majeed, S.; Danish, M.; Ibrahim, M.N.M.; Sekeri, S.H.; Ansari, M.T.; Nanda, A.; Ahmad, G. Bacteria Mediated Synthesis of Iron Oxide Nanoparticles and Their Antibacterial, Antioxidant, Cytocompatibility Properties. J. Clust. Sci. 2021, 32, 1083–1094. [Google Scholar] [CrossRef]
  203. Hulikere, M.M.; Joshi, C.G.; Danagoudar, A.; Poyya, J.; Kudva, A.K.; Dhananjaya, B.L. Biogenic synthesis of gold nanoparticles by marine endophytic fungus-Cladosporium cladosporioides isolated from seaweed and evaluation of their antioxidant and antimicrobial properties. Process Biochem. 2017, 63, 137–144. [Google Scholar] [CrossRef]
  204. Hulikere, M.M.; Joshi, C.G. Characterization, antioxidant and antimicrobial activity of silver nanoparticles synthesized using marine endophytic fungus-Cladosporium cladosporioides. Process Biochem. 2019, 82, 199–204. [Google Scholar] [CrossRef]
  205. Gupta, K.; Chundawat, T.S. Bio-inspired synthesis of platinum nanoparticles from fungus Fusarium oxysporum: Its characteristics, potential antimicrobial, antioxidant and photocatalytic activities. Mater. Res. Express 2019, 6, 1050d6. [Google Scholar] [CrossRef]
  206. Mohana, S.; Sumathi, S. Multi-Functional Biological Effects of Palladium Nanoparticles Synthesized Using Agaricus bisporus. J. Clust. Sci. 2020, 31, 391–400. [Google Scholar] [CrossRef]
  207. Noor, S.; Shah, Z.; Javed, A.; Ali, A.; Bilal Hussain, S.; Zafar, S.; Ali, H.; Muhammad, S.A. A fungal based synthesis method for copper nanoparticles with the determination of anticancer, antidiabetic and antibacterial activities. J. Microbiol. Methods 2020, 174, 105966. [Google Scholar] [CrossRef]
  208. Es-Haghi, A.; Taghavizadeh Yazdi, M.E.; Sharifalhoseini, M.; Baghani, M.; Yousefi, E.; Rahdar, A.; Baino, F. Application of response surface methodology for optimizing the therapeutic activity of ZnO nanoparticles biosynthesized from Aspergillus niger. Biomimetics 2021, 6, 34. [Google Scholar] [CrossRef]
  209. Mohamed, A.A.; Fouda, A.; Abdel-Rahman, M.A.; Hassan, S.E.-D.; El-Gamal, M.S.; Salem, S.S.; Shaheen, T.I. Fungal strain impacts the shape, bioactivity and multifunctional properties of green synthesized zinc oxide nanoparticles. Biocatal. Agric. Biotechnol. 2019, 19, 101103. [Google Scholar] [CrossRef]
  210. Mohamed, A.A.; Abu-Elghait, M.; Ahmed, N.E.; Salem, S.S. Eco-friendly mycogenic synthesis of ZnO and CuO nanoparticles for in vitro antibacterial, antibiofilm, and antifungal applications. Biol. Trace Elem. Res. 2021, 199, 2788–2799. [Google Scholar] [CrossRef] [PubMed]
  211. Manimaran, K.; Murugesan, S.; Ragavendran, C.; Balasubramani, G.; Natarajan, D.; Ganesan, A.; Seedevi, P. Biosynthesis of TiO2 nanoparticles using edible mushroom (Pleurotus djamor) extract: Mosquito larvicidal, histopathological, antibacterial and anticancer effect. J. Clust. Sci. 2021, 32, 1229–1240. [Google Scholar] [CrossRef]
  212. Bhuyar, P.; Rahim, M.H.A.; Sundararaju, S.; Ramaraj, R.; Maniam, G.P.; Govindan, N. Synthesis of silver nanoparticles using marine macroalgae Padina sp. And its antibacterial activity towards pathogenic bacteria. Beni-Suef Univ. J. Basic Appl. Sci. 2020, 9, 3. [Google Scholar] [CrossRef] [Green Version]
  213. Ozturk, B.Y.; Gursu, B.Y.; Dag, I. Antibiofilm and antimicrobial activities of green synthesized silver nanoparticles using marine red algae Gelidium corneum. Process Biochem. 2020, 89, 208–219. [Google Scholar] [CrossRef]
  214. González-Ballesteros, N.; Prado-López, S.; Rodríguez-González, J.B.; Lastra, M.; Rodríguez-Argüelles, M.C. Green synthesis of gold nanoparticles using brown algae Cystoseira baccata: Its activity in colon cancer cells. Colloids Surf. B 2017, 153, 190–198. [Google Scholar] [CrossRef]
  215. Sayadi, M.H.; Salmani, N.; Heidari, A.; Rezaei, M.R. Bio-synthesis of palladium nanoparticle using Spirulina platensis alga extract and its application as adsorbent. Surf. Interfaces 2018, 10, 136–143. [Google Scholar] [CrossRef]
  216. Arsiya, F.; Sayadi, M.H.; Sobhani, S. Green synthesis of palladium nanoparticles using Chlorella vulgaris. Mater. Lett. 2017, 186, 113–115. [Google Scholar] [CrossRef]
  217. Ramkumar, V.S.; Pugazhendhi, A.; Prakash, S.; Ahila, N.K.; Vinoj, G.; Selvam, S.; Kumar, G.; Kannapiran, E.; Rajendran, R.B. Synthesis of platinum nanoparticles using seaweed Padina gymnospora and their catalytic activity as PVP/PtNPs nanocomposite towards biological applications. Biomed. Pharmacother. 2017, 92, 479–490. [Google Scholar] [CrossRef]
  218. Arya, A.; Gupta, K.; Chundawat, T.S.; Vaya, D. Biogenic synthesis of copper and silver nanoparticles using green alga Botryococcus braunii and its antimicrobial activity. Bioinorg. Chem. Appl. 2018, 2018, 7879403. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  219. Salem, D.M.S.A.; Ismail, M.M.; Aly-Eldeen, M.A. Biogenic synthesis and antimicrobial potency of iron oxide (Fe3O4) nanoparticles using algae harvested from the Mediterranean Sea. Egypt. J. Aquat. Res. 2019, 45, 197–204. [Google Scholar] [CrossRef]
  220. El-Belely, E.F.; Farag, M.; Said, H.A.; Amin, A.S.; Azab, E.; Gobouri, A.A.; Fouda, A. Green synthesis of zinc oxide nanoparticles (ZnO-NPs) using Arthrospira platensis (class: Cyanophyceae) and evaluation of their biomedical activities. Nanomaterials 2021, 11, 95. [Google Scholar] [CrossRef]
  221. Kobayashi, M.; Tomita, S.; Sawada, K.; Shiba, K.; Yanagi, H.; Yamashita, I.; Uraoka, Y. Chiral meta-molecules consisting of gold nanoparticles and genetically engineered tobacco mosaic virus. Optics Express 2012, 20, 24856–24863. [Google Scholar] [CrossRef] [PubMed]
  222. Pan, Y.; Blum, A.S.; Simine, L.; Mauzeroll, J. Nanometals templated by tobacco mosaic virus coat protein with enhanced catalytic activity. Appl. Catal. B 2021, 298, 120540. [Google Scholar] [CrossRef]
  223. Singh, A.; Dar, M.Y.; Joshi, B.; Sharma, B.; Shrivastava, S.; Shukla, S. Phytofabrication of silver nanoparticles: Novel drug to overcome hepatocellular ailments. Toxicol. Rep. 2018, 5, 333–342. [Google Scholar] [CrossRef]
  224. Kanwar, R.; Rathee, J.; Salunke, D.B.; Mehta, S.K. Green Nanotechnology-Driven Drug Delivery Assemblies. ACS Omega 2019, 4, 8804–8815. [Google Scholar] [CrossRef] [Green Version]
  225. Namdari, M.; Eatemadi, A.; Soleimaninejad, M.; Hammed, A.T. A brief review on the application of nanoparticle enclosed herbal medicine for the treatment of infective endocarditis. Biomed. Pharm. 2017, 87, 321–331. [Google Scholar] [CrossRef]
  226. Patra, J.K.; Das, G.; Fraceto, L.F.; Campos, E.V.R.; del, P. Rodriguez-Torres, M.; Acosta-Torres, L.S.; Diaz-Torres, L.A.; Grillo, R.; Swamy, M.K.; Sharma, S.; et al. Nano-based drug delivery systems: Recent developments and future prospects. J. Nanobiotech. 2018, 16, 71. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  227. Perez-Herrero, E.; Fernandez-Medarde, A. Advanced targeted therapies in cancer: Drug nanocarriers, the future of chemotherapy. Eur. J. Pharm. Biopharm. 2015, 93, 52–79. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  228. Ibrahim, A.Y.; Mahmoud, K.; El-Hallouty, S.M. Screening of antioxidant and cytotoxicity activities of some plant extracts from Egyptian flora. J. Appl. Sci. Res. 2011, 7, 1246–1258. [Google Scholar]
  229. Pradhan, D. Pharmacological effect of some fractions obtained from Sapindus trifoliatus acting as an antioxidant and against mammary cell proliferation. Afr. J. Pharm. Pharmacol. 2014, 8, 455–463. [Google Scholar] [CrossRef] [Green Version]
  230. Yang, G.; Chen, C.; Zhu, Y.; Liu, Z.; Xue, Y.; Zhong, S.; Wang, C.; Gao, Y.; Zhang, W. GSH-Activatable NIR Nanoplatform with Mitochondria Targeting for Enhancing Tumor-Specific Therapy. ACS Appl. Mater. Interfaces 2019, 11, 44961–44969. [Google Scholar] [CrossRef] [PubMed]
  231. Bruins, M.R.; Kapil, S.; Oehme, F.W. Microbial resistance to metals in the environment. Ecotoxicol. Environ. Saf. 2000, 45, 198–207. [Google Scholar] [CrossRef]
  232. Zhang, P.; Wang, P.; Yan, L.; Liu, L. Synthesis of gold nanoparticles with Solanum xanthocarpum extract and their in vitro anticancer potential on nasopharyngeal carcinoma cells. Int. J. Nanomed. 2018, 13, 7047–7059. [Google Scholar] [CrossRef] [Green Version]
  233. Wang, L.; Xu, J.; Yan, Y.; Liu, H.; Karunakaran, T.; Li, F. Green synthesis of gold nanoparticles from Scutellaria barbata and its anticancer activity in pancreatic cancer cell (PANC-1) activity in pancreatic cancer cell (PANC-1). Artif. Cells Nanomed. Biotechnol. 2019, 47, 1617–1627. [Google Scholar] [CrossRef] [Green Version]
  234. Zhang, Y.; Liu, B.; Wu, H.; Li, B.; Xu, J.; Duan, L.; Jiang, C.; Zhao, X.; Yuan, Y.; Zhang, G.; et al. Anti-tumor activity of verbascoside loaded gold nanoparticles. J. Biomed. Nanotechnol. 2014, 10, 3638–3646. [Google Scholar] [CrossRef] [PubMed]
  235. Todescato, F.; Fortunati, I.; Minotto, A.; Signorini, R.; Jasieniak, J.J.; Bozio, R. Engineering of semiconductor nanocrystals for light emitting applications. Materials 2016, 9, 672. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  236. Martinez-Carmona, M.; Gunko, Y.; Vallet-Regi, M. ZnO Nanostructures for Drug Delivery and Theranostic Applications. Nanomaterials 2018, 8, 268. [Google Scholar] [CrossRef] [Green Version]
  237. Thambiraj, S.; Hema, S.; Shankaran, D.R. Functionalized Gold Nanoparticles for Drug Delivery Applications. Mater. Today Proc. 2018, 5, 16763–16773. [Google Scholar] [CrossRef]
  238. Malathi, S.; Balakumaran, M.D.; Kalaichelvan, P.T.; Balasubramanian, S. Green synthesis of gold nanoparticles for controlled delivery. Adv. Mater. Lett. 2013, 4, 933–940. [Google Scholar] [CrossRef]
  239. Vijayashree, I.; Niranjana, P.; Prabhu, G.; Sureshbabu, V.; Manjanna, J. Conjugation of Au nanoparticles with chlorambucil for improved anticancer activity. J. Clust. Sci. 2017, 28, 133–148. [Google Scholar] [CrossRef] [Green Version]
  240. Alqahtani, M.A.; Al Othman, M.R.; Mohammed, A.E. Bio fabrication of silver nanoparticles with antibacterial and cytotoxic abilities using lichens. Sci. Rep. 2020, 10, 16781. [Google Scholar] [CrossRef]
  241. Thirumurugan, A.; Blessy, V.; Karthikeyan, M. Comparative Study on Doxorubicin Loaded Metallic Nanoparticles in Drug Delivery Against MCF-7 Cell Line. In Applications of Nanomaterials; Bhagyaraj, S.M., Oluwafemi, O.S., Kalarikkal, N., Thomas, S., Eds.; Elsevier Ltd.: Amsterdam, The Netherlands, 2018; pp. 303–313. [Google Scholar] [CrossRef]
  242. Rokade, S.S.; Joshi, K.A.; Mahajan, K.; Patil, S.; Tomar, G.; Dubal, D.S.; Parihar, V.S.; Kitture, R.; Bellare, J.R. Gloriosa superba mediated synthesis of platinum and palladium nanoparticles for induction of apoptosis in breast cancer. Bioinorg. Chem. Appl. 2018, 2018, 4924186. [Google Scholar] [CrossRef] [Green Version]
  243. Azizi, S.; Shahri, M.M.; Rahman, H.S.; Rahim, R.A.; Rasedee, A.; Mohamad, R. Green synthesis palladium nanoparticles mediated by white tea (Camellia sinensis) extract with antioxidant, antibacterial, and antiproliferative activities toward the human leukemia (MOLT-4) cell line. Int. J. Nanomed. 2017, 12, 8841–8853. [Google Scholar] [CrossRef] [Green Version]
  244. Prakashkumar, N.; Vignesh, M.; Brindhadevi, K.; Phuong, N.-T.; Pugazhendhi, A.; Suganthy, N. Enhanced antimicrobial, antibiofilm and anticancer activities of biocompatible neem gum coated palladium nanoparticles. Prog. Org. Coat. 2021, 151, 106098. [Google Scholar] [CrossRef]
  245. Vimala, D.K.; Sundarraj, S.; Paulpandi, M.; Srinivasan, V.; Kannan, S. Green synthesized doxorubicin-loaded zinc oxide nanoparticles regulate the bax and Bcl-2 expression in breast and colon carcinoma. Process Biochem. 2013, 49, 160–172. [Google Scholar] [CrossRef]
  246. Jacob, S.J.P.; Bharathkumar, R.; Ashwathram, G. Aspergillus niger mediated synthesis of ZnO Nanoparticles and their antimicrobial and in vitro Anticancerous activity. World J. Pharm. Res. 2014, 3, 3044–3054. [Google Scholar]
  247. Fadeel, D.A.A.; Hanafy, M.S.; Kelany, N.A.; Elywa, M.A. Novel greenly synthesized titanium dioxide nanoparticles compared to liposomes in drug delivery: In vivo investigation on Ehrlich solid tumor model. Heliyon 2021, 7, 6. [Google Scholar] [CrossRef] [PubMed]
  248. Rao, T.N.; Riyazuddin, B.P.; Ahmad, N.; Khan, R.A.; Hassan, I.; Shahzad, S.A.; Husain, F.M. Green synthesis and structural classification of Acacia nilotica mediated-silver doped titanium oxide (Ag/TiO2) spherical nanoparticles: Assessment of its antimicrobial and anticancer activity. Saudi J. Biol. Sci. 2019, 26, 1385–1391. [Google Scholar] [CrossRef] [PubMed]
  249. Sankar, R.; Maheswari, R.; Karthik, S.; Shivashangari, K.S.; Ravikumar, V. Anticancer activity of Ficus religiosa engineered copper oxide nanoparticles. Mater. Sci. Eng. C 2014, 44, 234–239. [Google Scholar] [CrossRef]
  250. Siddiqi, K.S.; Rahman, A.U.; Tajuddin, H.A. Biogenic Fabrication of Iron/Iron Oxide Nanoparticles and Their Application. Nanoscale Res. Lett. 2016, 11, 498. [Google Scholar] [CrossRef] [Green Version]
  251. Tyagi, N.; Gupta, P.; Khan, Z.; Neupane, Y.R.; Mangla, B.; Mehra, N.; Ralli, T.; Alhalmi, A.; Ali, A.; Al Kamaly, O.; et al. Superparamagnetic Iron-Oxide Nanoparticles Synthesized via Green Chemistry for the Potential Treatment of Breast Cancer. Molecules 2023, 28, 2343. [Google Scholar] [CrossRef]
  252. Fakhri, A.; Tahami, S.; Nejad, P.A. Preparation and Characterization of Fe3O4-Ag2O Quantum Dots Decorated Cellulose Nanofibers as a Carrier of Anticancer Drugs for Skin Cancer. J. Photochem. Photobiol. B 2017, 175, 83–88. [Google Scholar] [CrossRef]
  253. Mollick, M.M.R.; Rana, D.; Dash, S.K.; Chattopadhyay, S.; Bhowmick, B.; Maity, D.; Mondal, D.; Pattanayak, S.; Roy, S.; Chakraborty, M.; et al. Studies on green synthesized silver nanoparticles using Abelmoschus esculentus (L.) pulp extract having anticancer (in vitro) and antimicrobial applications. Arab. J. Chem. 2019, 12, 2572–2584. [Google Scholar] [CrossRef] [Green Version]
  254. Sarkar, S.; Kotteeswaran, V. Green synthesis of silver nanoparticles from aqueous leaf extract of Pomegranate (Punica granatum) and their anticancer activity on human cervical cancer cells. Adv. Nat. Sci. Nanosci. Nanotechnol. 2018, 9, 025014. [Google Scholar] [CrossRef]
  255. Saratale, R.G.; Shin, H.S.; Kumar, G.; Benelli, G.; Kim, D.-S.; Saratale, G.D. Exploiting antidiabetic activity of silver nanoparticles synthesized using Punica granatum leaves and anticancer potential against human liver cancer cells (HepG2). Artif. Cells Nanomed. Biotechnol. 2017, 46, 211–222. [Google Scholar] [CrossRef] [Green Version]
  256. Saratale, R.G.; Benelli, G.; Kumar, G.; Su Kim, D.; Saratale, G.D. Bio-fabrication of silver nanoparticles using the leaf extract of an ancient herbal medicine, dandelion (Taraxacum officinale), evaluation of their antioxidant, anticancer potential, and antimicrobial activity against phytopathogens. Environ. Sci. Pollut. Res. Int. 2018, 25, 10392–10406. [Google Scholar] [CrossRef]
  257. Mukherjee, S.; Chowdhury, D.; Kotcherlakota, R.; Patra, S.; Vinothkumar, B.; Bhadra, M.P.; Sreedhar, B.; Patra, C.R. Potential Theranostics Application of Bio-Synthesized Silver Nanoparticles (4-in-1 System). Theranostics 2014, 4, 316–335. [Google Scholar] [CrossRef] [Green Version]
  258. Mukherjee, S.; Sau, S.; Madhuri, D.; Bollu, V.S.; Madhusudana, K.; Sreedhar, B.; Banerjee, R.; Patra, C.R. Green synthesis and characterization of monodispersed gold nanoparticles: Toxicity study, delivery of doxorubicin and its bio-distribution in mouse model. J. Biomed. Nanotechnol. 2016, 12, 165–181. [Google Scholar] [CrossRef]
  259. Shivashankarappa, A.; Sanjay, K.R. Photodynamic therapy on skin melanoma and epidermoid carcinoma cells using conjugated 5-aminolevulinic acid with microbial errulatem silver nanoparticles. J. Drug Target. 2019, 27, 434–441. [Google Scholar] [CrossRef]
  260. Parida, U.K.; Biswal, S.K.; Bindhani, B.K. Green Synthesis and Characterization of Gold Nanoparticles: Study of Its Biological Mechanism in Human SUDHL-4 Cell Line. Adv. Biol. Chem. 2014, 04, 360–375. [Google Scholar] [CrossRef] [Green Version]
  261. Gul, A.R.; Shaheen, F.; Rafique, R.; Bal, J.; Waseem, S.; Park, T.J. Grass-mediated biogenic synthesis of silver nanoparticles and their drug delivery evaluation: A biocompatible anti-cancer therapy. Chem. Eng. J. 2021, 407, 127202. [Google Scholar] [CrossRef]
  262. Rajan, A.; Rajan, A.R.; Philip, D. Elettaria cardamomum seed mediated rapid synthesis of gold nanoparticles and its biological activities. OpenNano 2017, 2, 1–8. [Google Scholar] [CrossRef]
  263. Devi, P.R.; Kumar, C.S.; Selvamani, P.; Subramanian, N.; Ruckmani, K. Synthesis and characterization of Arabic gum capped gold nanoparticles for tumor-targeted drug delivery. Mater. Lett. 2015, 139, 241–244. [Google Scholar] [CrossRef]
  264. Khoobchandani, M.; Katti, K.K.; Karikachery, A.R.; Thipe, V.C.; Srisrimal, D.; Mohandoss, D.K.D.; Darshakumar, R.D.; Joshi, C.M.; Katti, K.V. New Approaches in Breast Cancer Therapy Through Green Nanotechnology and Nano-Ayurvedic Medicine–Pre-Clinical and Pilot Human Clinical Investigations. Int. J. Nanomed. 2020, 15, 181–197. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  265. Yin, H.Q.; Shao, G.; Gan, F.; Ye, G. One-step, Rapid and Green Synthesis of Multifunctional Gold Nanoparticles for Tumor-Targeted Imaging and Therapy. Nanoscale Res. Lett. 2020, 15, 15–29. [Google Scholar] [CrossRef]
  266. Aboyewa, J.A.; Sibuyi, N.R.S.; Meyer, M.; Oguntibeju, O.O. Gold Nanoparticles Synthesized Using Extracts of Cyclopia intermedia, Commonly Known as Honeybush, Amplify the Cytotoxic Effects of Doxorubicin. Nanomaterials 2021, 11, 132. [Google Scholar] [CrossRef]
  267. Oladipo, A.O.; Iku, S.I.; Ntwasa, M.; Nkambule, T.T.; Mamba, B.B.; Msagati, T.A. Doxorubicin conjugated hydrophilic AuPt bimetallic nanoparticles fabricated from Phragmites australis: Characterization and cytotoxic activity against human cancer cells. J. Drug Deliv. Sci. Technol. 2020, 57, 101749. [Google Scholar] [CrossRef]
  268. Nagajyothi, P.C.; Pandurangan, M.; Kim, D.H.; Sreekanth, T.V.M.; Shim, J. Green synthesis of iron oxide nanoparticles and their catalytic and in vitro anticancer activities. J. Clust. Sci. 2016, 28, 245–257. [Google Scholar] [CrossRef]
  269. Zadeh, F.A.; Jasim, S.A.; Atakhanova, N.E.; Majdi, H.S.; Jawad, M.A.; Hasan, M.K.; Borhani, F.; Khatami, M. Drug delivery and anticancer activity of errulatemmat mesoporous Fe2O3 nanoparticles. IET Nanobiotechnol. 2022, 16, 85. [Google Scholar] [CrossRef] [PubMed]
  270. Naz, S.; Islam, M.; Tabassum, S.; Fernandes, N.F.; De Blanco, E.J.C.; Zia, M. Green synthesis of hematite (α-Fe2O3) nanoparticles using Rhus punjabensis extract and their biomedical prospect in pathogenic diseases and cancer. J. Mol. Struct. 2019, 1185, 1–7. [Google Scholar] [CrossRef]
  271. Rajeswaran, S.; Thirugnanasambandan, S.; Dewangan, N.K.; Moorthy, R.K.; Kandasamy, S.; Vilwanathan, R. Multifarious Pharmacological Applications of Green Routed Eco-Friendly Iron Nanoparticles Synthesized by Streptomyces sp. (SRT12). Biol. Trace Elem. Res. 2019, 194, 273–283. [Google Scholar] [CrossRef] [PubMed]
  272. Nagayothi, P.C.; Muthuraman, P.; Sreekanth, T.V.M.; Kim, D.H.; Shim, J. Green synthesis: In-vitro anticancer activity of copper oxide nanoparticles against human cervical carcinoma cells. Arab. J. Chem. 2017, 10, 215–225. [Google Scholar] [CrossRef] [Green Version]
  273. Kanagamani, K.; Muthukrishnan, P.; Saravanakumar, K.; Shankar, K.; Kathiresan, A. Photocatalytic degradation of environmental perilous gentian violet dye using errulat-mediated zinc oxide nanoparticle and its anticancer activity. Rare Met. 2019, 38, 277–286. [Google Scholar] [CrossRef]
  274. He, F.; Yu, W.; Fan, X.; Jin, B. In vitro cytotoxicity of biosynthesized titanium dioxide nanoparticles in human prostate cancer cell lines. Trop. J. Pharm. Res. 2018, 16, 2793. [Google Scholar] [CrossRef] [Green Version]
  275. Dobrucka, R.; Romaniuk-Drapała, A.; Kaczmarek, M. Anti-Leukemia Activity of Au/CuO/ZnO Nanoparticles Synthesized used Verbena officinalis Extract. J. Inorg. Organomet. Polym. Mater. 2021, 31, 191–202. [Google Scholar] [CrossRef]
  276. 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] [Green Version]
  277. Wang, L.; Hu, C.; Shao, L. The-antimicrobial-activity-of-nanoparticles-present-situati. Int. J. Nanomed. 2017, 12, 1227–1249. [Google Scholar] [CrossRef] [Green Version]
  278. Hamed, A.A.; Kabary, H.; Khedr, M.; Emam, A.N. Antibiofilm, antimicrobial and cytotoxic activity of extracellular green-synthesized silver nanoparticles by two marine-derived actinomycete. RSC Adv. 2020, 10, 10361–10367. [Google Scholar] [CrossRef]
  279. Shaaban, M.; El-Mahdy, A.M. Biosynthesis of Ag, Se, and ZnO nanoparticles with antimicrobial activities against resistant pathogens using waste isolate Streptomyces enissocaesilis. IET Nanobiotechnol. 2018, 12, 741–747. [Google Scholar] [CrossRef]
  280. Tahir, K.; Nazir, S.; Ahmad, A.; Li, B.; Ullah, A.; Ul, Z.; Khan, H.; Ullah, F.; Ullah, Q.; Khan, A.; et al. Facile and green synthesis of phytochemicals capped platinum nanoparticles and in vitro their superior antibacterial activity. J. Photochem. Photobiol. B Biol. 2017, 166, 246–251. [Google Scholar] [CrossRef]
  281. Haider, A.; Ijaz, M.; Ali, S.; Haider, J.; Imran, M.; Majeed, H.; Shahzadi, I.; Ali, M.M.; Khan, J.A.; Ikramet, M. Green synthesized phytochemically (Zingiber officinale and Allium sativum) reduced nickel oxide nanoparticles and confirmed bactericidal and catalytic potential. Nanoscale Res. Lett. 2020, 15, 50. [Google Scholar] [CrossRef]
  282. Ashwini, J.; Aswathy, T.R.; Rahul, A.B.; Thara, G.M.; Nair, A.S. Synthesis and Characterization of Zinc Oxide Nanoparticles Using Acacia caesia Bark Extract and Its Photocatalytic and Antimicrobial Activities. Catalysts 2021, 11, 1507. [Google Scholar] [CrossRef]
  283. Da, E.; Taha, A.; Afkar, E. Green synthesis of iron nanoparticles by Acacia nilotica pods extract and its catalytic, adsorption, and antibacterial activities. Appl. Sci. 2018, 8, 1922. [Google Scholar] [CrossRef] [Green Version]
  284. Yugandhar, P.; Vasavi, T.; Uma, P.; Devi, M. Bioinspired green synthesis of copper oxide nanoparticles from Syzygium alternifolium (Wt.) Walp: Characterization and evaluation of its synergistic antimicrobial and anticancer activity. Appl. Nanosci. 2017, 7, 417–427. [Google Scholar] [CrossRef] [Green Version]
  285. Rajkuberan, C.; Sudha, K.; Sathishkumar, G.; Sivaramakrishnan, S. Antibacterial and cytotoxic potential of silver nanoparticles synthesized using latex of Calotropis gigantea L. Spectrochim. Acta A Mol. Biomol. Spectrosc. 2015, 136, 924–930. [Google Scholar] [CrossRef] [PubMed]
  286. Rajkuberan, C.; Prabukumar, S.; Sathishkumar, G.; Wilson, A.; Ravindran, K.; Sivaramakrishnan, S. Facile synthesis of silver nanoparticles using Euphorbia antiquorum L. latex extract and evaluation of their biomedical perspectives as anticancer agents. J. Saudi Chem. Soc. 2017, 21, 911–919. [Google Scholar] [CrossRef] [Green Version]
  287. Karakoti, A.S.; Hench, L.L.; Seal, S. The potential toxicity of nanomaterials—The role of surfaces. JOM 2006, 58, 77–82. [Google Scholar] [CrossRef]
  288. Romero, M.; Cantón, E.; Pemán, J.; Gobernado, M. Antifúngicos inhibidores de la síntesis del glucano. Rev. Esp. Quimioter. 2005, 18, 281–299. [Google Scholar] [PubMed]
  289. Arciniegas-Grijalba, P.A.; Patiño-Portela, M.C.; Mosquera-Sánchez, L.P.; Guerrero-Vargas, J.A.; Rodríguez-Páez, J.E. ZnO nanoparticles (ZnONPs) and their antifungal activity against coffee fungus Erythricium salmonicolor. Appl. Nanosci. 2017, 7, 225–241. [Google Scholar] [CrossRef] [Green Version]
  290. Kumar, C.G.; Poornachandra, Y. Biodirected synthesis of Miconazole-conjugated bacterial silver nanoparticles and their application as antifungal agents and drug delivery vehicles. Colloids Surf. B Biointerfaces 2015, 125, 110–119. [Google Scholar] [CrossRef] [PubMed]
  291. Kahzad, N.; Salehzadeh, A. Green synthesis of CuFe2O4@Ag nanocomposite using the Chlorella vulgaris and evaluation of its effect on the expression of norA efflux pump gene among Staphylococcus aureus strains. Biol. Trace Elem. Res. 2020, 198, 359–370. [Google Scholar] [CrossRef]
  292. Emmanuel, R.; Palanisamy, S.; Chen, S.-M.; Chelladurai, K.; Padmavathy, S.; Saravanan, M.; Prakash, P.; Ajmal Ali, M.; Al-Hemaid, F.M.A. Antimicrobial efficacy of green synthesized drug blended silver nanoparticles against dental caries and periodontal disease causing microorganisms. Mater. Sci. Eng. C 2015, 56, 374–379. [Google Scholar] [CrossRef]
  293. Agarwal, H.; Nakara, A.; Shanmugam, V.K. Anti-inflammatory mechanism of various metal and metal oxide nanoparticles synthesized using plant extracts: A review. Biomed. Pharmacother. 2019, 109, 2561–2572. [Google Scholar] [CrossRef]
  294. Agarwal, H.; Shanmugam, V. A review on anti-inflammatory activity of green synthesized zinc oxide nanoparticle: Mechanism-based approach. Bioorg. Chem. 2020, 94, 103423. [Google Scholar] [CrossRef]
  295. Kedi, P.B.E.; Meva, F.E.; Kotsedi, L.; Nguemfo, E.L.; Zangueu, C.B.; Ntoumba, A.A.; Mohamed, H.E.A.; Dongmo, A.B.; Maaza, M. Eco-friendly synthesis, characterization, in vitro and in vivo anti-inflammatory activity of silver nanoparticle-mediated Selaginella myosurus aqueous extract. Int. J. Nanomed. 2018, 13, 8537–8548. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  296. Singh, P.; Ahn, S.; Kang, J.P.; Veronika, S.; Huo, Y.; Singh, H.; Chokkaligam, M.; ElAgamy Farh, M.; Aceituno, V.C.; Kim, Y.J.; et al. In vitro errulatemmatory activity of spherical silver nanoparticles and monodisperse hexagonal gold nanoparticles by fruit extract of Prunus errulate: A green synthetic approach. Artif Cells. Nanomed. Biotechnol. 2018, 46, 2022–2032. [Google Scholar] [CrossRef] [Green Version]
  297. Swamy, M.M.; Surendra, B.S.; Mallikarjunaswamy, C.; Pramila, S.; Rekha, N.D. Bio-mediated synthesis of ZnO nanoparticles using Lantana camara flower extract: Its characterizations, photocatalytic, electrochemical and anti-inflammatory applications. Environ. Nanotechnol. Monit. Manag. 2021, 15, 100442. [Google Scholar] [CrossRef]
  298. Abdelbaky, A.S.; Abd El-Mageed, T.A.; Babalghith, A.O.; Selim, S.; Mohamed, A. Green Synthesis and Characterization of ZnO Nanoparticles Using Pelargonium odoratissimum (L.) Aqueous Leaf Extract and Their Antioxidant, Antibacterial and Anti-inflammatory Activities. Antioxidants 2022, 11, 1444. [Google Scholar] [CrossRef]
  299. Eming, S.A.; Martin, P.; Tomic-Canic, M. Wound repair and regeneration mechanisms. Sci. Transl. Med. 2014, 6, 265sr6. [Google Scholar] [CrossRef] [Green Version]
  300. Nosrati, H.; Heydari, M.; Tootiaei, Z.; Ganjbar, S.; Khodaei, M. Delivery of antibacterial agents for wound healing applications using polysaccharide-based scaffolds. J. Drug Deliv. Sci. Technol. 2023, 84, 104516. [Google Scholar] [CrossRef]
  301. Ahn, E.Y.; Jin, H.; Park, Y. Assessing the antioxidant, cytotoxic, apoptotic and wound healing properties of silver nanoparticles green-synthesized by plant extracts. Mater. Sci. Eng. C 2019, 101, 204–216. [Google Scholar] [CrossRef]
  302. Moniri, M.; Moghaddam, A.B.; Azizi, S.; Rahim, R.A.; Saad, W.Z.; Navaderi, M.; Arulselvan, P.; Mohamad, R. Molecular study of wound healing after using biosynthesized BNC/Fe3O4 nanocomposites assisted with a bioinformatics approach. Int. J. Nanomed. 2018, 13, 2955–2971. [Google Scholar] [CrossRef] [Green Version]
  303. Shao, F.; Yang, A.J.; Yu, D.M.; Wang, J.; Gong, X.; Tian, H.X. Bio-synthesis of Barleria gibsoni leaf extract mediated zinc oxide nanoparticles and their formulation gel for wound therapy in nursing care of infants and children. J. Photochem. Photobiol. B 2018, 189, 267–273. [Google Scholar] [CrossRef]
  304. Sankar, R.; Baskaran, A.; Subramanian, K. Inhibition of pathogenic bacterial growth on excision wound by green synthesized copper oxide nanoparticles leads to accelerated wound healing activity in Wistar albino rats. J. Mater. Sci. Mater. Med. 2015, 26, 214. [Google Scholar] [CrossRef]
  305. Augustine, R. Skin bioprinting: A novel approach for creating artificial skin from synthetic and natural building blocks. Prog. Biomater. 2018, 7, 77–92. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  306. Augustine, R.; Hasan, A. Emerging applications of biocompatible phytosynthesized metal/metal oxide nanoparticles in healthcare. J. Drug Deliv. Sci. Technol. 2020, 56, 101516. [Google Scholar] [CrossRef]
  307. Joseph, B.; Augustine, R.; Kalarikkal, N.; Thomas, S.; Seantier, B.; Grohens, Y. Recent advances in electrospun polycaprolactone based scaffolds for wound healing and skin bioengineering applications. Mater. Today Commun. 2019, 19, 319–335. [Google Scholar] [CrossRef]
  308. Augustine, R.; Kalarikkal, N.; Thomas, S. Electrospun PCL membranes incorporated with biosynthesized silver nanoparticles as antibacterial wound dressings. Appl. Nanosci. 2016, 6, 337–344. [Google Scholar] [CrossRef] [Green Version]
  309. Augustine, R.; Hasan, A.; Yadu Nath, V.K.; Thomas, J.; Augustine, A.; Kalarikkal, N.; Moustafa, A.-E.A.; Thomas, S. Electrospun polyvinyl alcohol membranes incorporated with green synthesized silver nanoparticles for wound dressing applications. J. Mater. Sci. Mater. Med. 2018, 29, 205–212. [Google Scholar] [CrossRef]
  310. Jadhav, K.; Rajeshwari, H.; Deshpande, S.; Jagwani, S.; Dhamecha, D.; Jalalpure, S.; Subburayan, K.; Baheti, D. Phytosynthesis of gold nanoparticles: Characterization, biocompatibility, and evaluation of its osteoinductive potential for application in implant dentistry. Mater. Sci. Eng. C 2018, 93, 664–670. [Google Scholar] [CrossRef]
  311. Wang, M.; Wang, L. Plant polyphenols mediated synthesis of gold nanoparticles for pain management in nursing care for dental tissue implantation applications. J. Drug Deliv. Sci. Technol. 2020, 58, 101753. [Google Scholar] [CrossRef]
  312. Tang, Y.; Rajendran, P.; Veeraraghavan, V.P.; Hussain, S.; Balakrishna, J.P.; Chinnathambi, A.; Alharbi, S.A.; Alahmadi, T.A.; Rengarajan, T.; Mohan, S.K. Osteogenic differentiation and mineralization potential of zinc oxide nanoparticles from Scutellaria baicalensis on human osteoblast-like MG-63 cells. Mater. Sci. Eng. C 2021, 119, 111656. [Google Scholar] [CrossRef]
  313. Naito, H.; Iba, T.; Takakura, N. Mechanisms of new blood-vessel formation and proliferative heterogeneity of endothelial cells. Int. Immunol. 2020, 32, 295–305. [Google Scholar] [CrossRef] [Green Version]
  314. Park, M.H.; kyung Kim, A.; Manandhar, S.; Oh, S.Y.; Jang, G.H.; Kang, L.; Lee, D.W.; Lee, S.H.; Lee, H.E.; Huh, T.L.; et al. CCN1 interlinks integrin and hippo pathway to autoregulate tip cell activity. Elife 2019, 8, e46012. [Google Scholar] [CrossRef]
  315. Johnson, K.E.; Wilgus, T.A. Vascular endothelial growth factor and angiogenesis in the regulation of cutaneous wound repair. Adv. Wound Care 2014, 3, 647–661. [Google Scholar] [CrossRef] [Green Version]
  316. Gacche, R.N.; Meshram, R.J. Angiogenic factors as potential drug target: Efficacy and limitations of anti-angiogenic therapy. Biochim. Biophys. Acta Rev. Cancer 2014, 1846, 161–179. [Google Scholar] [CrossRef]
  317. Marew, T.; Birhanu, G. Three dimensional printed nanostructure biomaterials for bone tissue engineering. Regener. Ther. 2021, 18, 102–111. [Google Scholar] [CrossRef] [PubMed]
  318. Xia, Y.; Sun, J.; Zhao, L.; Zhang, F.; Liang, X.-J.; Guo, Y.; Weir, M.D.; Reynolds, M.A.; Gu, N.; Xu, H.H.K. Magnetic field and nano-scaffolds with stem cells to enhance bone regeneration. Biomaterials 2018, 183, 151–170. [Google Scholar] [CrossRef] [PubMed]
  319. Moise, S.; Céspedes, E.; Soukup, D.; Byrne, J.M.; El Haj, A.J.; Telling, N.D. The cellular magnetic response and biocompatibility of biogenic zinc- and cobalt-doped magnetite nanoparticles. Sci. Rep. 2017, 7, 39922. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  320. Atanasov, A.G.; Waltenberger, B.; Pferschy-Wenzig, E.M.; Linder, T.; Wawrosch, C.; Uhrin, P.; Temml, V.; Wang, L.; Schwaiger, S.; Heiss, E.H.; et al. Discovery and resupply of pharmacologically active plant-derived natural products: A review. Biotechnol. Adv. 2015, 33, 1582–1614. [Google Scholar] [CrossRef] [Green Version]
  321. Zhou, J.; Hu, Z.; Zabihi, F.; Chen, Z.; Zhu, M. Progress and perspective of antiviral protective material. Adv. Fiber Mater. 2020, 2, 123–139. [Google Scholar] [CrossRef]
  322. Lysenko, V.; Lozovski, V.; Lokshyn, M.; Gomeniuk, Y.V.; Dorovskih, A.; Rusinchuk, N.; Pankivska, Y.; Povnitsa, O.; Zagorodnya, S.; Tertykh, V. Nanoparticles as antiviral agents against adenoviruses. Adv. Nat. Sci. Nanosci. 2018, 9, 025021. [Google Scholar] [CrossRef]
  323. Fouad, G.I. A proposed insight into the anti-viral potential of metallic nanoparticles against novel coronavirus disease-19 (COVID-19). Bull. Natl. Res. Cent. 2021, 45, 36. [Google Scholar] [CrossRef]
  324. Murugan, K.; Dinesh, D.; Paulpandi, M.; Althbyani, A.D.M.; Subramaniam, J.; Madhiyazhagan, P.; Wang, L.; Suresh, U.; Kumar, P.M.; Mohan, J.; et al. Nanoparticles in the fight against mosquito-borne diseases: Bioactivity of Bruguiera cylindrical-synthesized nanoparticles against dengue virus DEN-2 (in vitro) and its mosquito vector aedes aegypti (Diptera: Culicidae). Parasitol. Res. 2015, 114, 4349–4361. [Google Scholar] [CrossRef]
  325. Sharma, V.; Kaushik, S.; Pandit, P.; Dhull, D.; Yadav, J.P.; Kaushik, S. Green synthesis of silver nanoparticles from medicinal plants and evaluation of their antiviral potential against chikungunya virus. Appl. Microbiol. Biotechnol. 2019, 103, 881–891. [Google Scholar] [CrossRef] [PubMed]
  326. Al-Sanea, M.M.; Abelyan, N.; Abdelgawad, M.A.; Musa, A.; Ghoneim, M.M.; Al-Warhi, T.; Aljaeed, N.; Alotaibi, O.J.; Alnusaire, T.S.; Abdelwahab, S.F.; et al. Strawberry and ginger silver nanoparticles as potential inhibitors for SARS-CoV-2 assisted by in silico modeling and metabolic profiling. Antibiotics 2021, 10, 824. [Google Scholar] [CrossRef] [PubMed]
  327. Fatima, M.; Sadaf Zaidi, N.-u.S.; Amraiz, D.; Afzal, F. In vitro antiviral activity of Cinnamomum cassia and its nanoparticles against H7N3 influenza A virus. J. Microbiol. Biotechnol. 2016, 26, 151–159. [Google Scholar] [CrossRef]
  328. Kumar, S.D.; Singaravelu, G.; Ajithkumar, S.; Murugan, K.; Nicoletti, M.; Benelli, G. Mangrove-Mediated Green Synthesis of Silver Nanoparticles with High HIV-1 Reverse Transcriptase Inhibitory Potential. J. Clust. Sci. 2017, 28, 359–367. [Google Scholar] [CrossRef]
  329. Dev, A.; Binulal, N.S.; Anitha, A.; Nair, S.V.; Furuike, T.; Tamura, H.; Jayakumar, R. Preparation of Poly (Lactic Acid)/Chitosan Nanoparticles for Anti-HIV Drug Delivery Applications. Carbohydr. Polym. 2010, 80, 833–838. [Google Scholar] [CrossRef] [Green Version]
  330. Dhanasezhian, A.; Srivani, S.; Govindaraju, K.; Parija, P.; Sasikala, S.; Kumar, M. Anti-herpes simplex virus (HSV-1 and HSV-2) activity of biogenic gold and silver nanoparticles using seaweed Sargassum wightii. Indian J. Geo-Mar. Sci. 2019, 48, 1252–1257. [Google Scholar]
  331. Yugandhar, P.; Vasavi, T.; Rao, Y.J.; Devi, P.U.M.; Narasimha, G.; Savithramma, N. Cost effective, green synthesis of copper oxide nanoparticles using fruit extract of Syzygium alternifolium (Wt.) Walp: Characterization and evaluation of antiviral activity. J. Clust. Sci. 2018, 29, 743–755. [Google Scholar] [CrossRef]
  332. Srivastava, S.; Usmani, Z.; Atanasov, A.G.; Singh, V.K.; Singh, N.P.; Abdel-Azeem, A.M.; Prasad, R.; Gupta, G.; Sharma, M.; Bhargava, A. Biological nanofactories: Using living forms for metal nanoparticle synthesis. Mini Rev. Med. Chem. 2021, 21, 245–265. [Google Scholar] [CrossRef]
  333. Raj, S.; Sasidharan, S.; Balaji, S.N.; Saudagar, P. An overview of biochemically characterized drug targets in metabolic pathways of Leishmania parasite. Parasitol. Res. 2020, 119, 2025–2037. [Google Scholar] [CrossRef]
  334. Bajwa, H.U.R.; Khan, M.K.; Abbas, Z.; Riaz, R.; ur Rehman, T.; Abbas, R.Z.; Aleem, M.T.; Abbas, A.; Almutairi, M.M.; Alshammari, F.A.; et al. Nanoparticles: Synthesis and Their Role as Potential Drug Candidates for the Treatment of Parasitic Diseases. Life 2022, 12, 750. [Google Scholar] [CrossRef]
  335. Karthik, L.; Kumar, G.; Vishnu Kirthi, A.; Rahuman, A.A.; Rao, K.V.B. Streptomyces sp. LK3 mediated synthesis of silver nanoparticles and its biomedical application. Bioprocess Biosyst. Eng. 2014, 37, 261–267. [Google Scholar] [CrossRef]
  336. Alajmi, R.A.; AL-Megrin, W.A.; Metwally, D.; AL-Subaie, H.; Altamrah, N.; Barakat, A.M.; Moneim, A.E.A.; Al-Otaibi, T.T.; El-Khadragy, M. Anti-Toxoplasma activity of silver nanoparticles green synthesized with Phoenix dactylifera and Ziziphus spina-christi extracts which inhibits inflammation through liver regulation of cytokines in Balb/c mice. Biosci. Rep. 2019, 39, BSR20190379. [Google Scholar] [CrossRef] [Green Version]
  337. Ullah, I.; Cosar, G.; Abamor, E.S.; Bagirova, M.; Shinwari, Z.K.; Allahverdiyev, A.M. Comparative study on the antileishmanial activities of chemically and biologically synthesized silver nanoparticles (AgNPs). Biotech 2018, 8, 98. [Google Scholar] [CrossRef] [PubMed]
  338. Kalangi, S.K.; Dayakar, A.; Gangappa, D.; Sathyavathi, R.; Maurya, R.S.; Narayana, D. Biocompatible silver nanoparticles reduced from Anethum graveolens leaf extract augments the antileishmanial efficacy of miltefosine. Experim. Parasitol. 2016, 170, 184–192. [Google Scholar] [CrossRef]
  339. Jan, H.; Shah, M.; Usman, H.; Khan, M.A.; Zia, M.; Hano, C.; Abbasi, B.H. Biogenic Synthesis and Characterization of Antimicrobial and Antiparasitic Zinc Oxide (ZnO) Nanoparticles Using Aqueous Extracts of the Himalayan Columbine (Aquilegia pubiflora). Front. Mater. 2020, 7, 2020. [Google Scholar] [CrossRef]
  340. Sumbal, A.; Nadeem, S.; Naz, J.S.; Ali, A.; Mannan, M.; Zia, M. Synthesis, characterization and biological activities of monometallic and bimetallic nanoparticles using Mirabilis jalapa leaf extract. Biotechnol. Rep. 2019, 22, e00338. [Google Scholar] [CrossRef]
  341. Ott, M.; Norberg, E.; Zhivotovsky, B.; Orrenius, S. Mitochondrial targeting of tBid/Bax: A role for the TOM complex? Cell Death Differ. 2009, 16, 1075–1082. [Google Scholar] [CrossRef] [Green Version]
  342. Leonard, K.; Ahmad, B.; Okamura, H.; Kurawaki, J. In situ green synthesis of biocompatible ginseng capped gold nanoparticles with remarkable stability. Colloids Surf. B Biointerfaces 2011, 82, 391–396. [Google Scholar] [CrossRef]
  343. Docea, A.O.; Calina, D.; Buga, A.M.; Zlatian, O.; Paoliello, M.M.B.; Mogosanu, G.D.; Streba, C.T.; Popescu, E.L.; Stoica, A.E.; Bîrcă, A.C.; et al. The effect of silver nanoparticles on antioxidant/pro-oxidant balance in a murine model. Int. J. Mol. Sci. 2020, 21, 1233. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  344. Virkutyte, J.; Varma, R.S. Green synthesis of metal nanoparticles: Biodegradable polymers and enzymes in stabilization and surface functionalization. Chem. Sci. 2011, 2, 837. [Google Scholar] [CrossRef]
  345. Sharma, V.K.; Siskova, K.M.; Zboril, R.; Gardea-Torresdey, J.L. Organic-coated silver nanoparticles in biological and environmental conditions: Fate, stability and toxicity. Adv. Colloid Interface Sci. 2014, 204, 15–34. [Google Scholar] [CrossRef]
  346. Liu, W.; Wu, Y.; Wang, C.; Li, H.C.; Wang, T.; Liao, C.Y.; Cui, L.; Zhou, Q.F.; Yan, B.; Jiang, G.B. Impact of silver nanoparticles on human cells: Effect of particle size. Nanotoxicology 2010, 4, 319–330. [Google Scholar] [CrossRef]
  347. Rheder, D.T.; Guilger, M.; Bilesky-José, N.; Germano-Costa, T.; Pasquoto-Stigliani, T.; Gallep, T.B.B.; Grillo, R.; Carvalho, C.; Dos, S.; Fraceto, L.F.; et al. Synthesisof biogenic silver nanoparticles using Althaea officinalis as reducing agent: Evaluation of toxicity and ecotoxicity. Sci. Rep. 2018, 8, 12397. [Google Scholar] [CrossRef] [Green Version]
  348. Pattanayak, S.; Rahaman, M.; Maity, D.; Chakraborty, S.; Kumar, S.; Chattopadhyay, S. Butea monosperma bark extract mediated green synthesis of silver nanoparticles: Characterization and biomedical applications. J. Saudi Chem. Soc. 2017, 21, 673–684. [Google Scholar] [CrossRef] [Green Version]
  349. Murugesan, K.; Koroth, J.; Srinivasan, P.P.; Singh, A.; Mukundan, S.; Karki, S.S.; Choudhary, B.; Gupta, C.M. Effects of green errulate m silver nanoparticles (ST06-AgNPs) using curcumin derivative (ST06) on human cervical cancer cells (HeLa) in vitro and EAC tumor bearing mice models. Int. J. Nanomed. 2019, 14, 5257–5270. [Google Scholar] [CrossRef] [Green Version]
  350. Moldovan, B.; David, L.; Vulcu, A.; Olenic, L.; Perde-Schrepler, M.; Fischer-Fodor, E.; Baldea, I.; Clichici, S.; Filip, G.A. In vitro and in vivo anti-inflammatory properties of green synthesized silver nanoparticles using Viburnum opulus L. fruits extract. Mater. Sci. Eng. C 2017, 79, 720–727. [Google Scholar] [CrossRef]
  351. Yang, L.; Kuang, H.; Zhang, W.; Aguilar, Z.P.; Wei, H.; Xu, H. Comparisons of the Biodistribution and Toxicological Examinations after Repeated Intravenous Administration of Silver and Gold Nanoparticles in Mice. Sci. Rep. 2017, 7, 3303. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  352. Gupta, M.; Mazumder, U.K.; Kumar, R.S.; Kumar, T.S. Antitumor activity and antioxidant role of Bauhinia errulat against Ehrlich ascites carcinoma in Swiss albino mice. Acta Pharmacol. Sin. 2004, 25, 1070–1076. [Google Scholar] [PubMed]
  353. Antony, J.J.; Sithika, M.A.A.; Joseph, T.A.; Suriyakalaa, U.; Sankarganesh, A.; Siva, D.; Kalaiselvi, S.; Achiraman, S. In vivo antitumor activity of biosynthesized silver nanoparticles using Ficus religiosa as a nanofactory in DAL induced mice model. Colloids Surf. B 2013, 108, 185–190. [Google Scholar] [CrossRef]
  354. Vasanth, S.B.; Kurian, G.A. Toxicity evaluation of silver nanoparticles synthesized by chemical and green route in different experimental models. Artif. Cells Nanomed. Biotechnol. 2017, 45, 1720–1726. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  355. Vasquez, R.D.; Apostol, J.G.; de Leon, J.D.; Mariano, J.D.; Mirhan, C.M.C.; Pangan, S.S.; Reyes, A.G.M.; Zamora, E.T. Polysaccharide-mediated green synthesis of silver nanoparticles from Sargassum siliquosum J.G. Agardh: Assessment of toxicity and hepatoprotective activity. OpenNano 2016, 1, 16–24. [Google Scholar] [CrossRef] [Green Version]
  356. Hanan, N.A.; Chiu, H.I.; Ramachandran, M.R.; Tung, W.H.; Nadhirah, N.; Zain, M.; Yahaya, N.; Lim, V. Cytotoxicity of plant-mediated synthesis of metallic nanoparticles: A systematic review. Int. J. Mol. Sci. 2018, 19, 1725. [Google Scholar] [CrossRef] [Green Version]
  357. Foo, Y.Y.; Periasamy, V.; Kiew, L.V.; Kumar, G.G.; Malek, S.N.A. Curcuma mangga-mediated synthesis of gold nanoparticles: Characterization, stability, cytotoxicity, and blood compatibility. Nanomaterials 2017, 7, 123. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  358. Nadhini, J.T.; Ezhilarasan, D.; Rajeshkumar, S. An eco-friendly synthesized gold nanoparticles induces cytotoxicity via apoptosis in HepG2 cells. Environ. Toxicol. 2021, 36, 24–32. [Google Scholar] [CrossRef] [PubMed]
  359. Elia, P.; Zach, R.; Hazan, S.; Kolusheva, S.; Porat, Z.E.; Zeiri, Y. Green synthesis of gold nanoparticles using plant extracts as reducing agents. Int. J. Nanomed. 2014, 9, 4007. [Google Scholar] [CrossRef] [Green Version]
  360. Ganeshkumar, M.; Ponrasu, T.; Raja, M.D.; Subamekala, M.K.; Suguna, L. Green synthesis of pullulan stabilized gold nanoparticles for cancer targeted drug delivery. Spectrochim. Acta Mol. Biomol. Spectrosc. 2014, 130, 64–67. [Google Scholar] [CrossRef]
  361. Lee, Y.J.; Ahn, E.Y.; Park, Y. Shape-dependent cytotoxicity and cellular uptake of gold nanoparticles synthesized using green tea extract. Nanoscale Res. Lett. 2019, 14, 129. [Google Scholar] [CrossRef] [Green Version]
  362. Mironava, T.; Hadjiargyrou, M.; Simon, M.; Jurukovski, V.; Rafailovich, M.H. Gold nanoparticles cellular toxicity and recovery: Effect of size, concentration and exposure time. Nanotoxicology 2010, 4, 120–137. [Google Scholar] [CrossRef]
  363. Eisa, N.; Almansour, S.; Alnaim, I.; Ali, A.; Algrafy, E.; Ortashi, K.; Awad, M.; Virk, P.; Hendi, A.; Eissa, F. Eco-synthesis and characterization of titanium nanoparticles: Testing its cytotoxicity and antibacterial effects. Green Process. Synth. 2020, 9, 462–468. [Google Scholar] [CrossRef]
  364. Selim, Y.A.; Azb, M.A.; Ragab, I.; Abd El-Azim, M.H.M. Green Synthesis of Zinc Oxide Nanoparticles Using Aqueous Extract of Deverra tortuosa and their Cytotoxic Activities. Sci. Rep. 2020, 10, 3445. [Google Scholar] [CrossRef] [Green Version]
Figure 1. Biosynthesis of biogenic NPs using different species, plant extracts, or macromolecules leading to the formation of structures with various compositions, shapes, and sizes.
Figure 1. Biosynthesis of biogenic NPs using different species, plant extracts, or macromolecules leading to the formation of structures with various compositions, shapes, and sizes.
Pharmaceutics 15 01650 g001
Figure 2. A mechanism, participating components, and factors affecting the biosynthesis of metallic nanoparticles.
Figure 2. A mechanism, participating components, and factors affecting the biosynthesis of metallic nanoparticles.
Pharmaceutics 15 01650 g002
Figure 3. Different shapes of biosynthesized nanoparticles.
Figure 3. Different shapes of biosynthesized nanoparticles.
Pharmaceutics 15 01650 g003
Figure 4. Intracellular and extracellular mechanisms of NPs biosynthesis. The intracellular synthesis can take place by cofactors, enzymes or other proteins. The extracellular process can be mediated by functional groups of macromolecules on the cell membrane/cell wall or by secreted biomolecules.
Figure 4. Intracellular and extracellular mechanisms of NPs biosynthesis. The intracellular synthesis can take place by cofactors, enzymes or other proteins. The extracellular process can be mediated by functional groups of macromolecules on the cell membrane/cell wall or by secreted biomolecules.
Pharmaceutics 15 01650 g004
Figure 5. Cytotoxic mechanisms of action of biogenic NPs that trigger apoptosis or necrosis in cancer cells.
Figure 5. Cytotoxic mechanisms of action of biogenic NPs that trigger apoptosis or necrosis in cancer cells.
Pharmaceutics 15 01650 g005
Figure 6. Mechanisms of antimicrobial activity of biosynthesized NPs.
Figure 6. Mechanisms of antimicrobial activity of biosynthesized NPs.
Pharmaceutics 15 01650 g006
Figure 7. Anti-inflammation mechanisms of biosynthesized metal and metal oxide NPs.
Figure 7. Anti-inflammation mechanisms of biosynthesized metal and metal oxide NPs.
Pharmaceutics 15 01650 g007
Table 1. Plant-derived biosynthesized metallic NPs.
Table 1. Plant-derived biosynthesized metallic NPs.
PlantSourceNPsSize, nmMorphologyApplication/
Tested Objects
Ref
Cucumis anguriaLeafAg11–27SphericalAntibacterial against S. aureus, E. coli[104]
Madhuca longifoliaFlowerAg30–50Spherical and ovalAntibacterial against E. coli, S. typhimurium, B. cereus, S. saprophyticus[28]
Banana wastePeduncleAg14.1SphericalAntimicrobial against S. aureus, E. coli, C. albicans, A. niger[105]
Trachyspermum ammiSeedAu16.6SpheroidalAnti-biofilm against L. monocytogenes and S. marcescens; anticancer against HepG2 cell lines[106]
Garcinia mangostanaPericarpAu, Ag13.7 ± 5.1
31.1 ± 4
Nanodumbbell shapesAnticancer, antioxidant against colon cancer Col320 cell line[107]
Asparagus racemosusRootAg, Au, Ag-Au10–50SphericalAntibacterial against P. aeurgnosia and S. aureus, anti-inflammatory in NK92 cells[108]
Rhanterium epapposumFlowerAg
Au
Zn
16.1
17.9
23.5
Spherical, triangular, hexagonalAntifungal against C. albicans and A. melleus, cytotoxic analysis against breast adenocarcinoma MCF-7, hepatocellular carcinoma HepG-2 and colorectal carcinoma HCT 116 human cell lines[109]
Rosmarinus officinalisLeafPd15–90Semi-sphericalAntibacterial against E. coli, S. aureus, S. epidermidis and M. lutens; antifungal against C. albicans, C. parapsilolis, C. glabrata and C. krusei[110]
Allium fistulosum, Tabernaemontanadivaricata, and Basella albaLeafPd20–50SphericalAntibacterial against Basella alba, Allium fistulosum; antifungal against C. albicans, A. flavus, and Penicillinium sp.; antioxidant (free radical scavenging), antidiabetic (α-amylase inhibition)[111]
Saudi’s datesFruitPt1.3–6.6SphericalAnticancer against colon cancer HCT-116, breast cancer MCF-7 and hepatocellular carcinoma HePG-2 cell line[112]
Celastrus paniculatus Willd.LeafCu2–10SphericalAntifungal against F. oxysporum[113]
Cissus vitigineaLeafCu2–20SphericalAntioxidant, antibacterial against E. coli, Enterococcus sp., Proteus sp. Klebsiella sp.[114]
Syzygium aromaticumBudCu12–15SphericalAntimicrobial against Staphylococcus spp., Bacillus spp., Pseudomonas spp., E. coli, A. niger, A. flavus and Penicillium spp.[56]
Azadirachta indicaLeafCuONASphericalAnticancer against breast cancer MCF-7 and Hela cell lines[115]
Achillea millefoliumLeafCuO28Semi-sphericalAntibacterial against S. aureus, M. tuberculosis, E. coli, K. pneumoniae, P. mirabili, C. diphtheriae, S. pyogenens; antifungal against C. albicans, A. flavus, M. canis and G. glabrata[116]
Musa ornateFlowerFe43.7NAAntibacterial against S. aureus, S. agalactiae, S. enterica, E. coli[117]
Ficus caricaLeafFe3O443–57MultiformAntioxidant (DPPH free radical scavenging)[118]
Platanus orientalisLeafIron oxide30–40SphericalAntifungal against A. niger and M. piriformis[119]
Mentha pulegium (L.)LeafZnO38–49Semi-sphericalAntimicrobial against E. coli and S. aureus[68]
Conyza canadensisLeafZnO-SphericalAntibacterial against E. coli and S. aureus[120]
Costus pictus D. DonLeafZnO40Elongated, hexagonal, and rod shapeAntimicrobial against S. aureus, B. subtilis, E. coli, S. parathyphi, A. niger, C. albicans anticancer against Dalton’s ascites cells[121]
Albizia lebbeckStem barkZnO66.6Irregular sphericalAntimicrobial against B. cereus, S. aureus, E. coli, K. pneumoniae and S. typhi, antioxidant (hydrogen peroxide free radical scavenging), anticancer against breast cancer MDA-MB 231 and MCF-7 cell lines[122]
Ailanthus altissimaPlant extractZnO13.3SphericalAntibacterial against S. aureus, K. pneumonia, E. coli, S. pyogenes; antioxidant (DPPH free radical scavenging)[123]
Punica granatumPeelZnO118.6Irregular nanorodsAntibacterial against S. aureus, E. aerogenes, P. aeruginosa, and K. pneumoniae[124]
Trigonella foenum-graecumLeafMgO13SphericalAntibacterial against E. coli, Bacillus spp., and S. aureus[125]
Ceratonia siliquaLeafCeO222NAAntioxidant (DPPH free radical scavenging), cytotoxic against breast cancer MCF-7 cell lines[126]
Table 2. An overview of different nano-vector microorganisms used for the production of metallic NPs employed in biomedical applications.
Table 2. An overview of different nano-vector microorganisms used for the production of metallic NPs employed in biomedical applications.
Class OrganismMetallic NPsIntra/Extracellular SynthesisSpeciesMorphologySize, nmApplication/
Tested Objects
Ref.
BacteriaAuExtracellularParacoccus haeundaensisspherical20.9 ± 3.5Antioxidant (DPPH free radical scavenging), cytotoxicity against normal kidney cells HEK 293, anticancer against lung carcinoma A549 and gastric adenocarcinoma AGS cell lines[190]
AuIntracellularLactobacillus kimchikusspherical5–30Antioxidant (DPPH free radical scavenging)[191]
AgIntracellularTransformed E. colispherical20NA[192]
AgExtracellularBacillus brevisspherical41–68Antibacterial against S. aureus and S. typhi[193]
AgExtracellularBacillus subtilisspherical3–20Antimicrobial against S. aureus, S. epidermidis, K. pneumoniae, E. coli and C. albicans[29]
PtIntracellularAcinetobacter calcoaceticuscubic2–3NA[194]
CuIntercellularStreptomyces capillispiralis
Ca-1
spherical3.6–59Antimicrobial against S. aureus, B. subtilis, B. dimenuta, P. aeruginosa, E. coli;
Antifungal against C. albicans and A. brasiliensis
[195]
ZnOExtracellularPseudomonas putidaspherical44.5Anti-biofilm against B. cereus and E. faecalis, antibacterial against P. otitidis, P. oleovorans, A. baumannii, B. cereus, E. faecalis[196]
ZnOExtracellularStaphilococcus aureusirregular10–50Antibacterial against E. coli, S. aureus, S. epidermis, E. faecalis, K. pneumonia, P. aeruginosa[69]
ZnOIntercellularSerratia nematodipilaspherical15–30Antibacterial against X. oryzae and antifungal against Alternaria sp.[197]
TiO2
ZnO
ExtracellularHalomonas elongataspherical104.6
18.1
Antibacterial against E. coli and S. aureus[198]
TiO2ExtracellularStreptomyces sp.spherical30–70Antimicrobial activity against E. coli, S. aureus, C. albicans, A. niger; antibiofilm against P. aeruginosa[199]
CuOExtracellularStreptomyces sp.spherical78–80Antifungal against F. oxysporum, P. ultimum, A. niger, A. alternata
cytotoxicity against Vero and Caco-2 cell lines
[200]
CuOExtracellularStreptomyces sp.spherical1.7–13.5Antibacterial against E. faecalis, S. typhimurium, P. aeruginosa, E. coli;
Antifungal against C. albicans, R. solani, A. niger
[201]
Iron oxideExtracellularProteus vulgaris ATCC-29905spherical19.2–30.5Antibacterial against methicillin resistant S. aureus;
anticancer against U87MG-glioblastoma cancer and HT-29 cancer cell line
[202]
Fungi and YeastAuExtracellularCladosporium cladosporioidesquasi-spherical60Antioxidant (DPPH and FRAP assay), antimicrobial against S. aureus, B. subtilis, P. aeruninosa, A. niger[203]
AgExtracellularCladosporium cladosporioidesspherical30–60Antimicrobial against S. aureus, B. subtilis, S. E. coli, S. epidermidis, and C. albicans; antioxidant (DPPH assay)[204]
AgExtracellularBeauveria bassianatriangular, circular, hexagonal10–50Antibacterial against E. coli, P. aeruginosa, S. aureus[30]
PtExtracellularFusarium oxysporumspherical25Antibacterial against E. coli
Antioxidant (DPPH assay)
[205]
PdIntracellularAgaricus bisporustriangular spherical13–18Antibacterial against S. aureus, S. pyrogens, B. subtilis, E. aerogenes, K. pneomoniae, P. vulgaris, anticancer against PK13 cell lines, anti-inflammatory with RBC cells
Antioxidant (DPPH method)
[206]
CuExtracellularAspergillus nigerround500–800Antidiabetic (α-glucosidase assay), anticancer against human hepatocellular carcinoma Huh-7 cell lines; antibacterial against E. coli, S. aureus, K. pneumoniae, M. luteus, B. subtilis[207]
ZnOExtracellularAspergillus nigerspherical35Antibacterial against E. coli; cytotoxicity against breast cancer MCF-7 cell lines[208]
ZnOExtracellularXylaria acutahexagonal34–55Antimicrobial against S. aureus, B. subtilis, P. aeruginosa, E. coli[70]
ZnOExtracellularAspergillus niger and
F. keratoplasticum
nanorods
hexagonal
8–32
10–42
Antimicrobial against B. subtilis, S. aureus, P. aeruginosa, E. coli[209]
CuO
ZnO
ExtracellularPenicillium chryogenumspherical
hexagonal
10.5–59.7
9–35
Antibiofilm against S. aureus and P. aeruginosa
Antimicrobial against S. aureus, B. subtilis, P. aeruginosa, E. coli, S. typhimurium and fungi F. solani, F. oxysporum, S. sclerotia, A. terreus
[210]
TiO2ExtracellularPleurotus djamorspherical31Anticancer against A549 cell lines,
Antibacterial against P. fluorescens, S. aureus, C. diphtheriae
[211]
AlgaeAgIntracellularPadina sp.spherical25–60Antimicrobial against S. aureus, B. subtilis, E. coli, S. typhi, P. aeruginosa[212]
AgExtracellularGelidium corneumspherical20–50Anti-biofilm and antibacterial against C. albicans and E. coli[213]
AuExtracellularCystosuira baccataspherical8.4Anticancer against colon cancer cell lines HT-29 and Caco-2[214]
PdExtracellularSpirulina patensisspherical10–20-[215]
PdExtracellularChlorella vulgarisspherical5–20-[216]
PtExtracellularPadina gymnosporaspherical, octahedral10–60Antimicrobial against E. coli, L. lactis, K. pneumoniae with no hemolytic activity in vitro[217]
CuExtracellularBotryococcus brauniicubic, spherical, triangular10–70Anti-microbial against P. aeruginosa, E. coli, K. pneumoniae, S. aureus and F. oxysporum[218]
Fe3O4ExtracellularColpomenia sinuosa and Pterocladia capillaceananospheres11.2–33.7
16.9–22.5
Antibacterial against E. coli, P. aeruginosa, S. typhi, V. cholera, B. subtilis, S. aureus
Antifungal against A. flavus and F. oxysporum
[219]
ZnOExtracellularSargassium muticumhexagonal30–57Anti-angiogenetic and apoptotic effect on human liver cancer cell line HepG2[71]
ZnOExtracellularArthrospira platensisspherical30–55Antimicrobial against B. subtilis, S. aureus, P. aeruginosa, E. coli and C. albicans[220]
VirusesAuintracellularTobacco mosaic virus (TMV)spherical5-[221]
AuintracellularGenetically modified Tobacco mosaic virus (TMV)spherical9–33-[160]
Au
Ag
intracellularPlant pathogenic virus of Squash leaf curl China virusspherical5–12
5–20
Cytotoxicity against A549 lung cancer cell line[162]
Pd
Pt
Au
Intra- and extracellularTobacco mosaic virus (TMV)spherical3.4 ± 0.5
3.1 ± 0.5
2.9 ± 0.5
Catalytic (allyl alcohol hydrogenation)[222]
Table 3. Biosynthesized NPs from different species with anticancer activity, their size, shape, drug conjugation and the main outcome of the research.
Table 3. Biosynthesized NPs from different species with anticancer activity, their size, shape, drug conjugation and the main outcome of the research.
NPsSourceSize, nmShapeDrugTested ObjectMain OutcomeRef.
AgAbelmoschus esculentus (L.)21.3Spherical-Jurkat cells (human T-cell lymphoma)-Antiproliferative effect in a dose-dependent manner with an IC50 value of 16.15 μg/mL
-Triggered high levels of ROS;
-Loss of integrity of mitochondrial membrane;
[253]
AgPunica granatul leaf extract41.7–69.6spherical-Cervical cancer HeLa cell line-IC50 for inhibiting 50% of the Hela cell line was 100 μg/mL;
-at a concentration of 100 μg/mL Ag NPs induced apoptosis by fragmentation of DNA
[254]
AgPunica granulatum leaf extract35–60Spherical-Liver cancer cell line HepG2-IC50 for the HepG2 cell line was 70 μg/mL;
-In vitro free radical scavenging activity.
[255]
AgTaraxacum officinale5–30Spherical-Liver cancer cell line HepG2-High cytotoxic effect against HepG2[256]
AgOlax scandens leaf extract20–60Spherical-B16 mouse melanoma cell line; A549 human lung cancer cell line; MCF7 human breast cancer cells-Inhibition of proliferation of both A549 and B16 in a dose-dependent manner;
-Compared to chemically synthesized Ag NPs, the biogenic showed biocompatible nature with normal cells;
[257]
AuPeltophorumpterocarpum65–149SphericalDoxLung (A549) and melanoma (B16F10) cancer cell line, C57BL6/J mice-Nano-conjugate cellular uptake and drug release were faster than pure Dox;
-No significant changes in hematology, serum clinical biochemistry, or histopathology in C57BL6/J mice were observed after 7 days
[258]
AgBacillus licheniformis20–80Triangular5-aminoevulinic acid (ALA)Skin melanoma (B16F10), epidermoid carcinoma (A431)-Higher cytotoxicity on both cell lines than the pure ALA and Ag NPs;[259]
AuSyzygium aromaticum12–20Spherical-Human
lymphoma SUDHL-4 cell line
-The IC50 of Au NPs after 48 h was 30 μM;
-Au NPs induced ROS accumulation and dose and time-dependent manner;
-Apoptosis mechanisms involved caspase activation, formation of pores in the mitochondrial membrane, and release of cytochrome c;
-The NPs decrease the growth and increase apoptosis of cells
[260]
Ag (Starch coated)Meadow grass (Poa anua)36.7 ± 7.9SphericalEuphorbia dracunculoides Lam. Plant extractHuman embryonic kidney (HEK293), Hep3B liver cancer, SCC-7 murine cancer cell line, Sprague Dawley rats-Up to 96% drug loading efficacy was observed;
-Good biocompatibility (90–100%) with HE293 at high concentrations (up to 1000 μg/mL)
-at concentration 400 μg/mL of Hep3B and SCC-7 decreased to 20 and 30%, respectively after 72 h;
-IC50 of Au NPs was 113.3 μg/mL;
-Significantly low cytotoxicity in vivo of the bio-drug NP conjugates
[261]
AuElettaria cardamomum seeds15.2Spherical-Cervical cancer (HeLa) cell line-the IC50 of Au NPs was 42.6 μL;
-Penetration of Au NPs through the cell membrane;
-Shrinking, rounding granulation morphological changes
[262]
AuVitex negundo (reducing agents) and Arabic gum (capping agent)98.7 ± 1.9SphericalEpirubicinLung adenocarcinoma A549 cell line-The amount of epirubicin bound was 85%;
-The IC50 for value NP conjugate was around 4 μg/mL whereas for the drug alone was 24 μg/mL
-The drug release from the conjugate was slow, thus the cytotoxicity and therapeutic efficacy were higher
[263]
AuParacoccus haeundaensis BC74171I bacterium20.9 ± 3.5Spherical-A549 and AGS cancer cell lines-Au NPs exert an antiproliferative effect on cancer cells and did not affect the normal cells HEK293 and HaCaT up to a concentration of 200 μg/mL;
-The percentage of cell growth inhibition of AGS cells was higher than that of A549 indicating selective cytotoxicity;
[190]
AuMagnifera indica55.5–65.5Spherical-MDA-MB-231 and breast cancer mice-Highly effective in controlling the growth of breast cancer in a dose-dependent fashion;[264]
AuPeptides100–150SphericalDoxHela and nude mice-Tumor inhibition in nude mice via tumor and tail vain was 66.7 and 57.7%, respectively;
-Efficient Dox transport vector for cervical cancer therapy.
[265]
AuCyclopia intemedia and magniferin20Spherical and TriangularDoxHuman colon cancer (Caco-2), prostate cancer (PC-3) and glioblastoma (U87) cell lines; normal breast epithelial cell lines-In combination with Dox, antitumor effects are augmented.
-At the same time, relatively low cytotoxicity to normal cells was observed
[266]
Au/PtPhragmites
australis leaves
35.1 ± 2.7Flower-like shapeDoxBreast (MCF-7) and lung (A459)
cancer cell line
-pH-controlled dependency response under acidic tumor conditions;
-A three-fold cell death to cancer cells compared to Dox alone;
-Dox-conjugated Au/Pt NPs exhibited time-release phosphatidylserine exposure;
-Cell-specific response against MCF-7 compared to A459
[267]
Fe2O3Psoralea corylifolia39Spherical-Renal tumor MDCK and Caki-2 cell lines-Strong cancer cell growth inhibiting in a dose-dependent manner[268]
Fe2O3Mentha piperita leaves17.9Polygonal (rhombic and hexagonal)DoxBreast MCF-7 cancer cell line-The IC50 of Fe2O3 NPs against cancer cells was 0.5 μg/mL while at 1 μg/mL their toxicity to normal was negligible;
-71% Dox loading efficacy;
-NPs-Dox conjugates significantly inhibited the tested cancer cells compared to free drug;
[269]
Fe2O3Rhus punjabensis extract41.5 ± 5Rhombohedral HL-60 leukemia and DU-145 prostate cancer cell line-The NPs were active against both cancer cells and leaved health control cell unaffected;[270]
FeStreptomyces sp.65–87Spherical DU145 and prostate cancer (PC3) cell line-The IC50 for value NPs was around 65 μg/mL;
-The color staining revealed an antiproliferative effect of the NPs on tumor cells;
[271]
CuOBean extract26.6Spherical, hexagonal, uneven shape-Human cervical carcinoma (HeLa) cell line-The IC50 for value NPs was around 13 μg/mL increased ROS and lipid peroxidation of liposomal membrane;
-Alteration in the mitochondrial structure;
-Inability of cells to proliferate.
[272]
ZnOLeucaena leucocephala50–200Hexagonal-Breast cancer (MCF-7) and prostate cancer (PC3) cell line;
Dalton lymphoma ascites (DLA) cell line
-The IC50 for value NPs was around 103 μg/mL
-Better cytotoxic activity on PC-3 than MCF-7 cell line;
-Against DLA cells the biosynthesized ZnO NPs revealed 92% inhibition with a concentration of 200 μg/mL;
[273]
TiO2Cinnamomum tamala23Irregular-Human prostate cancer (D145) cell line-TiO2 NPs showed a dose-dependent anticancer effect[274]
Au/
CuO/
ZnO
Verbena officinalis L. extract35Spherical-Jurkat cell line-The IC50 for value NPs was around 0.64 μmol;
-Within 24 h over 80% of the cells cultured in the presence of Au/CuO/ZnO NPs at a concentration of 10 μmol exhibited sighs of late apoptosis;
-about 60% of the cells at a concentration of 100 μmol underwent necrosis.
[275]
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

Nikolova, M.P.; Joshi, P.B.; Chavali, M.S. Updates on Biogenic Metallic and Metal Oxide Nanoparticles: Therapy, Drug Delivery and Cytotoxicity. Pharmaceutics 2023, 15, 1650. https://doi.org/10.3390/pharmaceutics15061650

AMA Style

Nikolova MP, Joshi PB, Chavali MS. Updates on Biogenic Metallic and Metal Oxide Nanoparticles: Therapy, Drug Delivery and Cytotoxicity. Pharmaceutics. 2023; 15(6):1650. https://doi.org/10.3390/pharmaceutics15061650

Chicago/Turabian Style

Nikolova, Maria P., Payal B. Joshi, and Murthy S. Chavali. 2023. "Updates on Biogenic Metallic and Metal Oxide Nanoparticles: Therapy, Drug Delivery and Cytotoxicity" Pharmaceutics 15, no. 6: 1650. https://doi.org/10.3390/pharmaceutics15061650

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