Next Article in Journal
Pulse Electrolysis Technique for Preparation of Bimetal Tin-Containing Electrocatalytic Materials
Next Article in Special Issue
One Step Synthesis of Oxygen Defective Bi@Ba2TiO4/BaBi4Ti4O15 Microsheet with Efficient Photocatalytic Activity for NO Removal
Previous Article in Journal
Annealing and Plasma Effects on the Structural and Photocatalytic Properties of TiO2 Fibers Produced by Electrospinning
Previous Article in Special Issue
Electrochemical Activity of Original and Infiltrated Fe-Doped Ba(Ce,Zr,Y)O3-Based Electrodes to Be Used for Protonic Ceramic Fuel Cells
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

An Evaluation of the Biocatalyst for the Synthesis and Application of Zinc Oxide Nanoparticles for Water Remediation—A Review

by
Nkosingiphile E. Zikalala
1,2,
Shohreh Azizi
1,2,*,
Sithembela A. Zikalala
3,
Ilunga Kamika
3,
Malik Maaza
1,2,
Ali Akbar Zinatizadeh
4,
Touhami Mokrani
5 and
Kasinathan Kaviyarasu
1,2,*
1
UNESCO-UNISA Africa Chair in Nanosciences & Nanotechnology Laboratory, College of Graduate Studies, University of South Africa, Muckleneuk Ridge, P.O. Box 392, Pretoria 0002, South Africa
2
Nanosciences African Network (NANOAFNET)-Materials Research Department, iThemba LABS-National Research Foundation, Somerset West, P.O. Box 722, Cape Town 7129, South Africa
3
Institute for Nanotechnology and Water Sustainability, College of Science, Engineering and Technology, University of South Africa, Florida, Johannesburg 1709, South Africa
4
Environmental Research Center, Department of Applied Chemistry, Razi University, Kermanshah P.O. Box 67144-14971, Iran
5
Department of Chemical Engineering, University of South Africa, Private Bag X6, Florida, Johannesburg 1709, South Africa
*
Authors to whom correspondence should be addressed.
Catalysts 2022, 12(11), 1442; https://doi.org/10.3390/catal12111442
Submission received: 30 September 2022 / Revised: 1 November 2022 / Accepted: 7 November 2022 / Published: 15 November 2022

Abstract

:
Global water scarcity is threatening the lives of humans, and it is exacerbated by the contamination of water, which occurs because of increased industrialization and soaring population density. The available conventional physical and chemical water treatment techniques are hazardous to living organisms and are not environmentally friendly, as toxic chemical elements are used during these processes. Nanotechnology has presented a possible way in which to solve these issues by using unique materials with desirable properties. Zinc oxide nanoparticles (ZnO NPs) can be used effectively and efficiently for water treatment, along with other nanotechnologies. Owing to rising concerns regarding the environmental unfriendliness and toxicity of nanomaterials, ZnO NPs have recently been synthesized through biologically available and replenishable sources using a green chemistry or green synthesis protocol. The green-synthesized ZnO NPs are less toxic, more eco-friendly, and more biocompatible than other chemically and physically synthesized materials. In this article, the biogenic synthesis and characterization techniques of ZnO NPs using plants, bacteria, fungi, algae, and biological derivatives are reviewed and discussed. The applications of the biologically prepared ZnO NPs, when used for water treatment, are outlined. Additionally, their mechanisms of action, such as the photocatalytic degradation of dyes, the production of reactive oxygen species (ROS), the generation of compounds such as hydrogen peroxide and superoxide, Zn2+ release to degrade microbes, as well as their adsorbent properties with regard to heavy metals and other contaminants in water bodies, are explained. Furthermore, challenges facing the green synthesis of these nanomaterials are outlined. Future research should focus on how nanomaterials should reach the commercialization stage, and suggestions as to how this ought to be achieved are presented.

Graphical Abstract

1. Introduction

The concurrent occurrence of organic dyes and inorganic toxic pollutants, such as heavy metals (HMs), pharmaceuticals, and pathogenic microorganisms in wastewater is a major worldwide concern [1]. Although the effects of each of the pollutants are well documented, the toxicity that arises from the co-occurrence of, and interactions between, these pollutants remains unknown; therefore, this presents a greater concern than the singular effects of each of the pollutants, with respect to the negative effect on the balance of the ecosystem, plants, and animals [2]. Synthetic dyes cause water to become unaesthetic, and they adversely compromise the health of humans and aquatic organisms by altering the biochemical oxygen demand (BOD), chemical oxygen demand (COD), salt concentration, light penetration, and rate of photosynthesis [3]. HMs are extremely poisonous due to the failure of living systems to remove them from the bodies of animals, thus leading to bioaccumulation. As such, HMs are responsible for neurological complications, carcinogenesis, mutations, and multiorgan failure in different organisms, even at low concentrations [4]. Pathogenic microorganisms comprise another class of public health concerns due to their short lifecycle, high reproductive rate per lifecycle, and ultrafast mutation rates that produce resistant strains. Around 500 waterborne microorganisms have been identified in wastewater treatment plants, including viral, bacterial, parasitic protozoan, and fungal pathogens [5]. These organisms invade a range of animal tissues, and they compromise the immune system through the release of toxins [6]; therefore, such detrimental pollutants must be efficiently eradicated from effluents before they are discharged into surface water bodies for the protection of ecological and biological systems.
Nanomaterials can address bottlenecks in conventional wastewater treatment systems because the nano-dimensions create a high surface area, quantum confinement, an attenuated surface charge, and other peculiar physical and chemical characteristics [7]. As such, various nanostructures have been synthesized for the purification of wastewater. Among the nanomaterials developed are: metallic nanoparticles such as Ag [8], Au [9], and Cu [10], and nanometal oxides such as TiO2 [11], Fe2O3 [12], SnO [13], AgO [14], CuO [15], and ZnO [16].
ZnO is among one of the most used semiconductors because of its ease of synthesis and its photo-driven electronic response. ZnO is a crystalline n-type semiconductor, it is a member of the II–VI group, and it exhibits a wide bandgap (3.37 eV) in the near UV spectrum [17,18]. Moreover, it exhibits high exciton binding energy (60 meV) in standard atmospheric conditions [19]. With respect to its crystallographic structures, ZnO has three polymorphs, namely: zincblende, cubic rock salt, and the hexagonal wurtzite. Of these polymorphs, wurtzite is the most stable, hence its synthesis at room temperature (RT) under pressure. The rock salt polymorph is formed under high pressure which results in the close packing of the Zn and O atoms, as shown in Figure 1. The zincblende polymorph ZnO is metastable, and therefore, it cannot exist on its own in ambient conditions [20,21]. With respect to the morphology of ZnO, as seen under transmission electron microscopy (TEM) and scanning electron microscopy (SEM), there is a wide variety of nanoparticulate ZnO morphologies. These include nanorods [22], spheres [23], nanosheets [24], hexagons [25], tripods [26], nanowires [27], nanotubes [28], nanostars [29], nanoflowers [30], and nanocubes [31], as shown in Figure 2.
Each morphology displays unique physical and chemical properties, and therefore, each one has a unique application or an added advantage over the other morphologies, with respect to a particular application [38]. These properties allow for the wide applicability of nanostructures in various fields, such as photocatalysis [39], antimicrobial action [40], adsorption [41], energy storage [42], gas sensors [43], cosmetics [44], cancer treatment [45], solar cells [46], agriculture [47], and wound healing [48], to name but a few. Owing to their multifunctional properties, such as better biocompatibility and the diverse applicability and solubility of ZnO in alkaline media, a wide range of physical and chemical synthesis procedures have been explored [49]. Regarding chemical synthesis, altering the precursor or the reaction parameters, such as temperature, pH, precursor concentration, type of precursor solvent, and the reaction pressure, produces a wide variety of morphologies, sizes, and geometries, with regard to the ZnO NPs. Jamkhande et al. [50] and Srujana et al. categorized chemical methods into the following groups: (i) liquid phase methods, which include sol-gel, hydrothermal, solvothermal, colloidal, and precipitation methods; and (ii) liquid–gas phase methods which include the spray pyrolysis and gas condensation methods [51]. Physical or mechanical synthesis techniques for ZnO include ball milling, physical/chemical/solid vapor deposition, laser ablation, sputter deposition, electric arc deposition, and ion implantation [52]. Some advantages of the chemical and physical methods include the high production rates, the fact that they produce controllable shapes, and they produce high purity levels [50]. Physical methods, however, suffer from non-uniformity in particle size and they have a high instrumentation cost. Moreover, chemical methods yield NPs that are agglomerated, unstable, as well as polydisperse. To counter these limitations, green synthesis routes have become the focus in current research that concerns the development of nanomaterials.

2. Green Synthesis

Over the last few decades, the “green synthesis” approach has received a lot of interest in the field of nanomaterial synthesis. Indeed, the green synthesis technique for nanomaterials is based on twelve green chemistry principles, namely: (i) to improve the reaction efficiency of atoms (atom economy); (ii) to improve the energy efficiency of reaction processes (i.e., avoidance of long synthesis times, high temperatures, and high pressure environments); (iii) to use harmless chemicals in order to reduce the toxicity of the procedures and the synthesized nanomaterials; (iv) to reduce waste throughout the synthesis process; (v) to use easily available renewable precursors; (vi) to design non-toxic biodegradable products; (vii) to follow safer synthesis routes and replace hazardous substances with safe (or at least less toxic) reagents; (viii) to avoid the use of derivatives such as stabilizers; (ix) to avoid the release of toxic waste; (x) to employ eco-friendly solvents, such as water, instead of organic solvents; (xi) to accelerate reactions by including a catalyst, so as to reduce total energy demands in order to improve sustainability and efficiency; and (xii) to limit accidents during and after the synthesis process [53,54,55]. In traditional chemical synthesis, organic capping agents such as oleic acid, linolenic acid, oleylamine, hexadecylamine, and 2-mercaptoethanol need to be used to attain particles with the desired morphology and with a high degree of uniformity [56]. Biological synthesis, on the other hand, encourages the use of readily available resources, such as non-toxic solvents and auxiliaries, for a benign synthesis protocol [57]; therefore, it is in accordance with the principles of green synthesis. Biological synthesis promotes the development of nanoparticles using nature-derived materials such as plants [58], bacteria [59], fungi [60], algae [61], and biological derivatives [62]. These bio-sources (enzymes, polysaccharides, phytochemicals, biomolecules, vitamins, etc.) serve as green capping, stabilizing, or chelating agents [7]. During biosynthesis, a zinc salt such as chloride, acetate, sulfate, or nitrate is generally added to a previously prepared biological extract (from plants, bacteria, fungi, algae, yeast, etc.) to facilitate the complexation of metal ions with the available functional groups in the biological extract (Figure 3). This is achieved by heating and stirring the extract in an alkaline environment. The temperature, pH, and ratio of the precursors determine the growth phase, wherein the nucleated seeds grow either isotropically or anisotropically. Moreover, as is the case in conventional chemical synthesis processes, the synthesis parameters in green synthesis procedures determine the morphology, size, agglomeration, and size distribution of the resultant NPs [63]. After the metal oxide precipitates, it is washed, dried, and/or calcined to produce ZnO NPs, as indicated in Figure 3.
Several reviews point out the potential of biosynthesized ZnO NPs for various applications. These applications can be broadly grouped into medical, environmental remediation, and agricultural applications [7,57,64,65,66]. The interest in green-synthesized ZnO has increased due to its biocompatibility, photocatalytic capability, antimicrobial effectiveness, physical and chemical stability, and cost effectiveness. The current review discusses the biosynthesis of ZnO by focusing on the synthesis conditions, factors influencing the properties of the synthesized NPs, and the mechanisms of ZnO formation using various biological sources. The key characterization techniques for ZnO NPs are also discussed. Recent developments on the use of the ZnO NPs, regarding water remediation (photocatalysis, adsorption, and antimicrobial applications), are reviewed with an emphasis on the mechanisms of action. Lastly, the prospective outcomes of the green synthesis of ZnO are discussed.

2.1. Plant-Mediated/Phytosynthesis of ZnO NPs

2.1.1. Phytosynthesis Synthesis of Undoped ZnO NPs

In comparison with other ZnO green synthesis techniques, the utilization of plant extracts (phytosynthesis) has dominated scientific research, hence the large number of reports available for review [67,68,69,70]. Researchers prefer using parts of plants over microbes since the use of plants eliminates laborious and delicate procedures such as microbe isolation, growth, and maintenance. Furthermore, phytosynthesis is safer, as it eliminates the health risks associated with the use of hazardous microorganisms [65]. The phytosynthesis approach is preferred for large-scale production, as plant extracts can easily be obtained through the use of eco-friendly solvents such as distilled water or alcohols [71]. Additionally, plant sources are easily accessible, abundant, and hence more sustainable [72]. Since plant extracts naturally possess secondary metabolites, including amino acids, alkaloids, and flavonoids, among others, they are commonly employed as both chelating and reducing agents in the fabrication of nanoparticles [73]. Although the explicit mechanism of their formation cannot yet be explained, Bala et al. and Senthilkumar et al. concluded that phenols, flavonoids, and proteins are the main participants in the development of metallic and metal oxide NPs [74,75].
Various parts of a known plant species, including the leaves, seeds, pericarps, peels, roots, bark, or even the whole plant, have been used to synthesize ZnO NPs. Zinc salts are said to be oxidized by phenols and flavonoids into ZnO NPs with the help of the numerous functional groups that these molecules possess. These functional groups attach onto the surface of Zn2+, and they oxidize the molecule into ZnO NPs [63,75]. Since the coordination chemistry of proteins is specific to that of zinc ions, they are the predominant chelating agents for Zn2+. The H2O molecules of the aqua complexes are exchanged when they bind to the protein ligands [76]. During their interaction with the acidic [Zn(H2O)6]2+, an electron from the COO initiates a nucleophilic reaction, which results in the formation of the [Zn(OH) (H2O)5]+ complex [77]. The [Zn(OH) (H2O)5]+ is then transmuted into ZnO NPs [78]. According to Singh et al., the −OH and/or −COO− of the flavonoids react with the ZnO precursor to generate a distinctive Zn flavonoid complex via the electrostatic interaction between negatively charged flavonoids and positively ZnO nuclei. This complex is then later converted to ZnO through thermal treatment [79]. An alkaline pH favors the formation of metal oxide nanoparticles that are stabilized by the phytochemicals in the plant extract [80]; thus, it follows that during the phytosynthesis of metal oxides such as ZnO, a combination of polyphenols, flavonoids, and proteins results in the creation of ZnO NPs [81]. Krol et al. [77] presented a probable reaction mechanism for a plant mediated ZnO, as shown in Figure 4.
Table 1 provides an overview of the plant-based synthesis of ZnO NPs. The use of the waste pericarp of Ananas comosus yielded spherical NPs, the sizes of which varied between 10 and 60 nm after sonicating for 1 h at 40 °C [82]. The seed extract of Eriobutria japonica produced hexagonal NPs when a mixture of the seed extract and Zn (CH3COO)2·2H2O was heated at 60 °C for 1 h. Using Fourier-transform infrared spectroscopy (FTIR), the reducing and capping agents were revealed to be phenolic groups, alcohols, and aliphatic amines. Recently, fruit-derived aqueous extracts have been employed to prepare ZnO NPs [22,83,84,85]. Golmohammadi et al. [85] obtained spherical NPs with an average size of 29 nm when they refluxed Zizyphus jujuba fruit and Zn(NO3)2·6H2O with water at 80 °C for 30 min prior to calcination. Recently, Faisal et al. [86] produced semi spherical NPs when synthesizing with the fruit extract of Myristica fragrans.
Biosynthesizing ZnO NPs using flower extracts have also been reported [87,88]. The extract of Hibiscus sabdariffa served as a reducing and stabilizing agent in order to produce wurtzite-phased NPs. The concentration of the employed extract determined the surface structure and size of the NP. The sizes ranged between 8.71 and 88.63 nm [83]. Khara et al. produced spherical and irregular ZnO NPs using Peltophorum pterocarpum, with a mean diameter of 69.45 nm [88]. This flower extract possesses phenolic compounds and flavonoids, among other properties, that potentially reduce and stabilize ZnO NPs. Moreover, Diallo and colleagues obtained very small (1–8.5 nm) quasi-spherical ZnO nanomaterials by applying the flower extract of Aspalathus linearis as a reducing/oxidizing chemical agent [89].
Regarding phytosynthesis, leaves are the most frequently used parts of a plant during extraction. This is likely due to their accessibility, abundance, and ease of grinding. Several authors have reported on the phytosynthesis of various nanomaterials through the use of leaf extracts from various species of plants [90,91,92]. Mostly, the attained NPs were spherical and irregularly shaped. Ekennia and colleagues, however, attained flower-shaped NPs (31 nm) when using an extract of Euphorbia sanguinea and ZnCl2 as the salt precursor [93]. The appearance of the 1638 cm−1 absorption band in the FTIR spectra was ascribed to the N H bending of amines in alkaloids or C=O stretching in polyphenols, flavonoids, and reducing sugars, which were thought to be responsible for reducing and stabilizing the synthesized NPs. An extract of quince seed mucilage was kept for 4 h in the presence of Zn(NO3)2·6H2O, before it was concurrently heated and stirred at 80 °C for 2 h to produce 25 nm spherical ZnO NPs [94]. Prior to this study, Prasad and colleagues [95] mixed Abelmoschus esculentus mucilage extract with Zn(OOCCH3)2·2(H2O) salt and quince seed mucilage in water, and the mixture was stirred for 4 h, at RT, to obtain spherical ZnO NPs with an average diameter of 29 nm and 70 nm long nanorods. The reducing and capping agents for the NPs were carbohydrates, which are available in excess in plants. The heterogeneous sizes and shapes were ascribed to the method employed and the high calcination temperature (700 °C for 2 h); this poses a challenge with regard to controlling the morphology of the prepared particles.
There have also been reports on the utilization of the aqueous extracts in roots, regarding the biogenic preparation of ZnO NPs with various morphologies [96]. A methanolic extract was employed by Verma et al. [68] to obtain hexagonal and rod-shaped NPs after stirring Zn(CH3COO)2·2H2O in the presence of Salvadora persica root extract for 3 h at 80 °C; it was then heated at 400 °C for an appropriate length of time. Matinise et al. prepared ZnO NPs by capping the NPs with an aqueous solution of Moringa oleifera leaf. The extract completely dissolved the zinc salts at an ambient temperature; this was evidenced by the fact that the solution changed color, from clear to brown, after 18 h without any precipitation. Spherically shaped ZnO NPs were formed after drying the concentrated solution at 100 °C for 1 h [97]. Highly crystalline single phase zincite ZnO NPs were produced by adding an aqueous solution of Agathosma betulina, in order to reduce the presence of zinc nitrate hexahydrate salt; in this scenario, the temperature was set at 100 °C for 2 h. The resultant precipitate was further annealed at 500 °C for 2 h to obtain (15.8 nm) quasi-spherical NPs [98]. When Ngom et al. prepared the same nanoparticles by employing Moringa oleifera leaves as the chelating agent, wurtzite phased ZnO nanomaterials, with an average diameter of 10.81 nm, were produced [99].
Table 1. Plant-based synthesis of ZnO nanoparticles.
Table 1. Plant-based synthesis of ZnO nanoparticles.
Serial No.Plant NamePlant PartExtraction Solvent
Temp (°C)
Time
Zinc PrecursorBiosynthesis
Time (h)/Temp (°C)
Calcination
Temp (°C)/Time (h)
ShapeSize (nm)Ref
1Pithecellobium dulcePeelWater, 70 °C for 50 minZn(CH₃CO₂)₂ 400/2Spherical30[83]
2Ananas comosusPericarpWater, 80 °C for 20 minZn(CH3COO)2·2H2O,Sonicated 1/40Spherical10–60[82]
3Punica granatumFruit peelWater, 65 °C for 1 hZn(NO3)2·6H2O–/90600/1Spherical and hexagonal38.98[84]
4Zizyphus jujubaFruitRefluxed with water, 80 °C for 30 minZn(NO3)2·6H2O4/80550/3Spherical29[85]
5Myristica fragransFruitWater, 150 °C for 20 minZn(NO3)2·2H2O2/60500/2Semispherical41.23[86]
6Hibiscus sabdariffaFlowerWater, 60 °C, for 1 h after 2 h stirring at RTZn(NO3)21/RT.400/1Semi spherical8.7–88.6[87]
7Peltophorum pterocarpumFlowerWater, 80 °C for 1 hZn(NO3)2–/80400/2Spherical and irregular69.45[88]
8Aspalathus linearisFlowerWater, 25 °C for 48 hZn(NO3)2·6H2O−/−300/2Quasi-spherical1–8.5[89]
9Eucalyptus spp. LeavesWater, 90 °C for 90 minZn(NO3)2·6H2O2/110250/6Irregular shapes32.71[90]
10Capparis zeylanicaLeavesWater, 60 °C for 20 minZn(OOCCH3)2·2(H2O),2/80400/2Spherical20–40[91]
11Ocimum tenuiflorumLeavesPowderZnSO40.5/RT600/0.5Spherical and granular50–63[92]
12Euphorbia sanguineaLeavesWater, 80 °C for 45 minZnCl22/90Flower-like34[93]
13Moringa oleiferaLeavesWater, 50 °C for 105 minZn(NO3)2·6H2O18/RT then 100500/1Spherical16–31.9[97]
Gathosma betulinaLeavesWater, 100 °C for 60 minZn(NO3)2·6H2O2/100500/2Quasi-spherical15.8[98]
14Quince seedsMucilageWater, 60 °C for 4 hZn(NO3)2·6H2O2/80 after keeping for 4 h at RT600/2Spherical25[94]
15Abelmoschus esculentusMucilageSoaked for 24 h at RTZn(OOCCH3)2·2(H2O),4/RT700/2Spherical
Rods
29
70
[95]
16Panax spp.Roots and stemsWater, 85 °C for 8 hZn(NO3)22.5/85500/2Leafy flower480[96]
17Salvadora persicaRootsWater and methanol, 80 °C for 3 hZn(CH3COO)2·2H2O2–3/80400/2Hexagonal and rods24–25[68]
18Rubus fairholmianusRootsAcetone using Soxhlet apparatusZn(NO3)248/80Spherical1–100 clusters[100]

2.1.2. Phytosynthesis of Doped ZnO Nanoparticles

Recently, the number of biogenically doped ZnO NPs has been increasing, which speaks to the efficiency of plant metabolites in nanomaterial synthesis. Doping occurs as a result of the wide band gap of ZnO, which causes ZnO to only absorb light that falls within the UV spectrum. Doping is a process wherein a point defect in a nanomaterial is formed by inserting a specific ion into the crystal lattice of a nanomaterial to modify its electronic band structure. The band structure is modified by the formation of shallow or deep energy level, known as mid-gap states, that extend the absorption range from shorter to longer wavelengths; this can even extend to the visible range of the electromagnetic spectrum [101,102]. The physicochemical properties of doped ZnO significantly depend on the dopant type and its concentration. p-type dopants tend to withdraw electrons from the ZnO, whereas n-type dopants increase the number of electrons. As the dopant concentration increases, the energy band gap E g of ZnO decreases until an optimal dopant concentration is reached, beyond which, the dopant ions act as charge recombination centers and reduce the efficiency of ZnO in photocatalytic processes. The dopants also become active sites that are different from the ZnO active sites, thereby increasing the mass transfer of pollutants from the solution to the ZnO surface. This phenomenon is not only useful for increasing the adsorption efficiency in cases where ZnO is used as an adsorbent, but it also helps to increase the level of light that penetrates the polluted water, thereby increasing the photocatalytic efficiency; this will be discussed further in Section 4.1 [103]. As well as induced defects, ZnO exhibits intrinsic defects, which include zinc vacancy (VZn), oxygen vacancy (VO), zinc interstitial (Zni), oxygen interstitial (Oi), or antisite oxygen (OZn) defects [104]. As with dopants, it is generally accepted that surface defects (i.e., surface VO, VZn, Oi, and Zni, especially the anionic Vo) are beneficial to photocatalytic activity because they can serve as active sites for photocatalysis. These defects create mid-gap states that alter the electronic band structure of photocatalysts to facilitate charge separation with longer lifetimes [101]. Conversely, deep level defects (VO and VZn), and other nonradiative defects, have negative impacts on photocatalysis because they create recombination centers for photogenerated charges; this causes electronic delocalization and low photocatalytic efficiency. These defects are repaired via annealing at appropriate temperatures [105].
The most used doping agents in the biogenetic synthesis of ZnO include Ag, Cu, Mg, Ce, Fe, and Co, though Ag doping is the most common [33,58,106]. Saeed et al. synthesized AgNO3-ZnO NPs at an ambient temperature using different plant species and parts. Owing to its strong reduction potential, the extract of Calotropis gigantea resulted in the formation of rod-shaped NPs [33]. Moringa olifera seed extract yielded flower-like NPs [58], whereas Ocimum tenuiflorum seed extract produced hexagonal and spherical NPs ranging between 50 and 60 nm in diameter [106]. Mn-ZnO NPs were prepared by mixing a salt precursor with Melastoma malabathricum. The mixture was heated at 60 °C until a paste formed, and it was further heated at 400 °C to obtain spherical NPs with an average diameter of 222 nm [107]. Fe-ZnO NPs were produced by Jan et al. This was achieved by reducing zinc ions with the flavonoids, phenolics, terpenoids, triterpenes, tannins, and fatty acids present in the leaf extract of Myrtus communis [108]. The NPs were produced after 3 h of stirring and heating at 60 °C, after which they were calcinated at 400 °C for 3 h. Their diameters were small, at 17 nm on average, with a pseudo-spherical shape. Okeke et al. prepared Mg-ZnO NPs by reacting Zn(NO3)2 and Mg(NO3)2 with Piper guineense leaf extract, which was then stirred and heated for 15 min at 80 °C before calcining the paste at 350 °C for 2 h [109]. A dual doping of ZnO NPs with Cu and Mg was successfully completed by Rahman et al. [110], who utilized a Ziziphus mauritiana solution to reduce Zn(NO3)2·6H2O, as depicted in Figure 5. The synthesized NPs were nearly spherical, but relatively large in size (0.1–1 µum), depending on the Cu/Mg dopant concentration. The FTIR suggests that the reduction of ions was controlled by tannins, flavonoids, saponins, and phenols. An eco-friendly fabrication of Nb-ZnO using Vernonia amygdalina leaf extract was reported by Nguyen et al., and the obtained Nb-ZnO NPs showed an improved photocatalytic degradation of tetracycline under natural light in comparison with pristine ZnO. The composite reached a 93.2% degradation efficiency after 3 h [111]. A summary of the synthetic procedures and the results of the doped-ZnO NPs are given in Table 2.

2.2. Microbial Synthesis of ZnO Nanoparticles

Microorganisms, or microbes comprising bacteria, fungus, and algae, have also been exploited for the biological synthesis of metal oxide NPs. Although many metal oxides have been synthesized biologically using microorganisms, the microbial production of ZnO has not been thoroughly investigated. An advantage of the microbial synthesis of NPs over phytosynthesis is the ease with which microorganisms can be reproduced, as they are easily grown in a laboratory. During microbial synthesis, metal reduction is made possible with the available enzymes, carbohydrates, proteins, and other metabolites that are intrinsically present in the microorganisms. The biomolecules that are secreted in the growth medium influence the shape, size, and dispersity of the produced NPs [112]. Under stressful conditions, or in the presence of heavy metals, microbes reduce these ions to metals in order to enable their own survival; thus, they act as natural nano-factories [113].

2.2.1. Bacterial Synthesis

Prokaryotic microorganisms, such as bacteria and algae, have been extensively studied because they are easy to handle and genetically modify [114]. Bacteria offer the advantage of being able to quickly multiply, thus creating a readily available source of secondary metabolites that are involved in the fabrication of ZnO. The most used microbes in the synthesis of NPs comprise Actinobacter sp., Corynebacterium sp., Klebsiella pneumonia, Lactobacillus sp., and Pseudomonas sp. [115]. The biological preparation of metal, and metal oxide nanomaterials, using bacterial cultures may occur in the intra- or extracellular environment [116]. The mechanism of formation during intracellular synthesis is challenging due to the intricacy of the bacterial cell make-up and its metabolic processes. The mechanism for the intracellular generation of ZnO in microbes is illustrated in Figure 6a. The cell walls of microbes have different types of polysaccharides and proteins that provide active sites for the binding of metal oxides through electrostatic interactions [117]. Then, reductases (NADH and NADPH) that are either present in the cell wall or secreted as soluble enzymes [118] reduce the trapped ions into elemental atoms [119]. Finally, the nuclei develop and assemble in the cytoplasm or cell wall, while the peptides and amino acids stabilize the NPs [120]. Bacteria-mediated nanoparticle synthesis creates less hazardous metal oxide nanoparticles, such as TiO2, CuO, and ZnO; however, the downside to this is that it involves isolating, screening, and culturing potential microbes, which are time consuming processes. In addition, the process may involve the use of expensive chemicals, such as growth media [119]. The size distribution, shape, and crystallinity of microbe-synthesized NPs are not easy to monitor [121]. Moreover, the intracellularly formed nanomaterials can only be released by lysing the bacterial cells; thus, time consuming and laborious steps are unavoidable [122]. As a result, research on bacterial synthesis has, to date, focused on extracellular synthesis. Regarding extracellular synthesis, the enzymes that are released by microorganisms into the growth culture are responsible for the bio-reduction of metal ions into NPs. The bio-reduction of zinc ions occurs when an electron is transferred from NADH with NADH reductase [123]. The extracellular synthesis pathways are depicted diagrammatically in Figure 6b.
Table 3 presents the current research on ZnO NP microbial synthesis. A variety of bacterial species are used in the green synthesis of ZnO NPs, including Cyanobacterium Nostoc sp. [29], Aeromonas hydrophila [59], Serratia nematodiphila [105], Bacillus subtillis [124], Bacillus haynesii [125], and Alkalibacillus sp. [126], to name but a few. Jayabalan et al. [127] reduced Zn (NO3)2 salt by incubating a mixture of the salt with a Pseudomonas putida cell culture solution at 37 °C for 24 h before calcining the resultant paste at 400 °C for 2 h. The resultant NPs were spherical agglomerates with diameters ranging between 25 and 45 nm. A green Taguchi method was employed to obtain optimum conditions in the extracellular bacterial-based production of ZnO NPs, using a Halomonas elongate extract, in order to reduce the presence of ZnCl2 salt. Spherical NPs, with an average diameter of 18 nm, were obtained [128]. To produce ZnO NPs, a biomass of zinc-tolerant Lactobacillus plantarum TA4 was used as a nanofactory. The presence of protein, carboxyl, and hydroxyl on the surfaces of the resulting ZnO NPs was verified using FTIR analysis. Furthermore, TEM analysis of the generated NPs revealed a flower-like morphology for ZnO NP (supernatant) samples, and an irregular shape for ZnO NP (cell biomass) samples, with particle sizes of 291.1 and 191.8 nm, respectively. The presence of protein, carboxyl, and hydroxyl on the surfaces of the resulting ZnO NPs were verified using FTIR analysis. Furthermore, TEM analysis of the generated NPs revealed a flower-like morphology for ZnO NP (supernatant) samples, and an irregular shape for ZnO NP (cell biomass) samples, with particle sizes of 291.1 and 191.8 nm, respectively [129]. The presence of protein, carboxyl, and hydroxyl on the surfaces of the resulting ZnO NPs was verified using FTIR analysis. In addition, TEM analysis of the generated NPs revealed a flower-like morphology for ZnO NP (supernatant) samples and an irregular shape for ZnO NP (cell biomass) samples, with particle sizes of 291.1 nm and 191.8 nm, respectively [129]. The biosynthesis of ZnO NPs was carried out using Lactococcus lactis NCDO1281(T) and Bacillus sp. PTCC 1538. A pure hexagonal crystalline structure of the particles was confirmed using XRD spectra. The SEM showed that ZnO NPs produced by Bacillus sp. were nanorods with an average size of 99 nm, whereas L. lactis produced NPs that were spherical, with diameter sizes ranging between 55   and   60.5 nm. The presence of amino and carbonyl groups in biosynthesized ZnO was confirmed using FTIR, and it was most likely related to residues of cell membrane proteins involved in NP biosynthesis [30].

2.2.2. Fungal Synthesis

Fungi can release higher concentrations of secondary metabolites than bacteria. In addition, they exhibit a higher tolerance to metal concentrations, stronger binding capabilities, and better metal bioaccumulation than bacteria [144]. It is therefore possible that fungi have more potential than bacteria with regard to the upscaling of green-produced NPs. Moreover, fungal cells are more tolerant to intermediate products during the synthesis process; hence, they are more suitable for large-scale synthesis [145]. The mechanisms for generating metal and metal oxide NPs using fungal biomass or cultures are similar to the one discussed for green synthesis using bacteria in Section 2.2.1. Different Aspergillus species have already been employed in the green synthesis of ZnO NPs, such as Aspergillus aeneus [60], Aspergillus fumigatus [131], Aspergillus fumigatus [132], and Aspergillus niger [133], as shown in Table 3. The ability to reduce and cap the NPs is conferred by the alcohols, phenols, and the aromatic and primary amines that are present in the fungal extracts of the various Aspergillus species [132]. During extracellular synthesis, fungi secrete enzymes that produce pure and monodispersed NPs. These particles are free from the cellular components that are associated with organisms [131]. The biomass of A. fumigatus reduced ZnSO4 and chelated the formed ZnO NPs after 72 h at 32 °C. The formed NPs were spherical with diameters ranging between 60 and 80 nm. The attained ZnO exhibited antimicrobial activities against Klebsiella pneumonia, P. aeruginosa, E. coli, S. aureus, and B. subtilis, respectively, as indicated in Figure 7. The synthesis of ZnO NPs, using the aqueous extract of Agaricus bisporus, was carried out in [134]. According to the FTIR results, the phenol groups contained the prominent ingredients for reducing and capping ZnO NPs. The zeta potential (−20.5 mV) validated the NPs’ stability, whereas the SEM and TEM revealed that the NPs were spheroids. ZNO NPs were produced with Zn(NO3)2·6H2O and Xylaria acuta fungal filtrate using the combustion method at 400 °C. The resultant sample was then calcined for 2 h at 700 °C to obtain hexagonal shaped NPs with diameters ranging between 34 and 55 nm [135].
Moghaddam et al. [136] utilized Pichia kudriavzevii in the mycogenic synthesis of ZnO. The reaction proceeded for 36 h at 35 °C, the resultant NPs were hexagonal, and they exhibited diameters that ranged between 10 and 61 nm. Small, quasi-spherical particles (2–6 nm) were prepared in [137] using Cochliobolus geniculatus. Fluorescence spectroscopy confirmed the presence of proteins on the prepared ZnO in the form of tyrosine moieties, as per the emission peak at 350 nm. The FTIR absorption peaks confirmed the role of aromatic, aliphatic amines (amide I and amide II) in the fungal extract, and potentially in the stabilization of the ZnO NPs as well. Motazedi et al. [138] prepared 20   nm   to   30 nm spherical ZnO NPs using a cell-free extract of Saccharomyces cerevisiae. The FTIR results suggested that the presence of alcohols, proteins, the phenolic group, fatty acids, and carbohydrates in the extracellular excretions of S. cerevisiae were potentially responsible for reducing the zinc ions so that ZnO NPs could be obtained. Mani et al. [139] prepared ZnO NPs at a pH of 7.2 using an endophytic fungus, Cladosporium tenuissimum FCBGr, which was isolated from a Vilva tree; bouquet-like NPs were obtained after 36 h of continuous stirring at RT.

2.2.3. Algal Synthesis

Algae are mono and multicellular aquatic photosynthesizing organisms, but unlike plants, algae are without roots and leaves [115]; however, similarly to plants, macroalgae have active metabolites such as alkaloids, peptides, polysaccharides, proteins, tannins, quinones, lipids, and glycerol [146]. These metabolites have –OH and –COOH functional groups that can chelate and stabilize ZnO NPs [146]; therefore, the chemical formation route of ZnO NPs during algal synthesis is similar to that of plants [57], as discussed in Section 2.1.1. Khalafi et al. [140] propose a mechanism of ZnO NP formation using a Chlorella extract (Figure 8). Based on the assumption that chlorella contains a significant amount of carbohydrates (20%), it was presumed that carbohydrates predominantly served as Zn2+ reducing agents and as stabilizers of the prepared ZnO NPs. Moreover, the biosynthesis of green ZnO NPs was assumed to occur through a donor–acceptor mechanism that occurred between the oxygen atoms of functional moieties in chlorella and Zn2+ wherein the −OH groups on the carbohydrates transfers an electron to the electrotrophilic Zn species, thus resulting in the oxidation of the −OH group and Zn2+ reduction to Zn atoms.
Algal species have been used to prepare ZnO NPs by reacting the extract with various zinc salts to obtain spheres, flower-like rods, and hexagonal shapes (Table 3). Recently, [141] produced 30   to   55 nm spherical NPs by reducing Zn(CH3COO)2·2H2O with Arthrospira platensis extract at 30 °C for 24 h. The emergence of Zn O characteristics, peaking at 503 cm−1 according to the FTIR results, confirmed the presence of ZnO particles. Algal extracts that were investigated for their ability to be synthesized with doped ZnO NPs have been successfully produced (Table 3). For instance, Rajaboopathi and Thambidurai prepared Ag-ZnO NPs by mixing separate solutions of AgNO3 and Zn(NO3)2 salts with the leaf extract of Padina gymnospora. The solution was stirred at 60 °C for 30 min which led to a brown–white precipitate emerging from the light brown solution. The stretching and bending vibrations of water molecules (−OH) occurred at 3435 and 1638 cm−1, respectively, whereas the stretching and bending vibrational bands of carbonate ions appeared at 1448 and 860 cm−1, respectively. For the pristine ZnO NPs, the Zn-O band appeared at 420 cm−1, whereas for Ag-ZnO, the Zn-O stretching vibrational bands appeared at 417 cm−1 [142]. Sargassum species extract was used to reduce the Zn (NO3)2·6H2O and Co(NO3)2·6H2O salts, and to stabilize the resultant Co-ZnO NPs with the aid of microwave radiation at pH 10.5. The obtained NPs were nearly spherical (5.4–6.8 nm). The interaction between Zn2+, Co2+, and the algae was ascertained as a result of the variation in the intensity of the FTIR spectra that corresponded with the Metal–O bond [143].

2.2.4. Preparation of ZnO NPs Using Biological Derivatives Synthesis

Apart from macro- and microorganisms, several biological derivatives have been used as precursors to synthesize NPs; however, only a handful reports discuss the use of biological derivatives for the green fabrication of ZnO, as shown in Table 4. Moreover, the polysaccharide pullulan was stirred together with Zn(NO3)2·6H2O for 5 h at RT. This was followed by a 1 h heat treatment at 400 °C to yield spherical and hexagonal ZnO NPs with an average diameter of 58.13 nm [147]. El-Saied and Ibrahim, 2020, used chitosan to synthesize ZnO NPs, with average size of 55 to 70 nm [148]. Vijayakumar et al. [149] prepared egg albumen wrapped ZnO NPs, with diameters ranging between 20   and   60 nm, and with spherical and diagonal platelets. Additionally, artemia eggshells were used to hydrothermally produce ellipsoidal ZnO NPs. The prepared NPs had a 50 nm diameter on average, according to TEM [150]. Highly crystalline crustin-capped ZnO NPs were produced after the coprecipitation of the salt with an extract of crustin at RT for 2 h. The average size of the resultant NPs were 50 nm [151]. Amino acid-capped NPs were produced via a wet chemical synthesis, prior to calcining at 550 °C for 3.5 h. The produced sheet-like NPs had diameters ranging between 22.46   and   40.29 nm [152]. Smaller ZnO NPs were produced with the aid of tannic acid, as both the size influencer and capping agent were treated at 70 °C for 1.5 h without further heat treatment. The particles had a spherical morphology, with diameters ranging between 26 and 34 nm [153]. A separate study reported that honey and cow urine, respectively, yielded spherical and hexagonal leaf-like ZnO NPs [154].

2.3. Factors Influencing ZnO Nanoparticle Biosynthesis

To control the size distribution, shapes, and formation rate of NPs, the optimization of the reaction media and physical parameters is crucial. Biosynthesized NPs can be commercially considered if the NPs’ yield is high and if they exhibit the desired size, size distribution, and morphology [119].

2.3.1. Temperature

The synthesis temperature influences the rate of crystallization in the wet synthesis procedures of semiconductor nanoparticles, and ZnO is no exception. The increased kinetic energy within the reaction vessel speeds up the rate at which nanoparticles collide, therefore forming crystalline nuclei. In cases of calcination, the calcination temperature influences the degree to which the amorphous ZnO matrix transforms into crystalline ZnO. The rate of nuclei formation increases with temperature, thus enhancing the degree of crystallinity, especially during biosynthetic processes involving ZnO [64]. In their investigation on the effect of temperature on the quality of the ZnO when using Hibiscus subdariffa leaf extract, Bala et al. [74] found that the UV-Vis spectra did not show the characteristic absorption band at 355 nm for the ZnO that was synthesized at 30 °C, thus suggesting that the material was only an amorphous ZnO; however, for the ZnO synthesized at 60 and 100 °C, the presence of ZnO NPs was confirmed by the appearance of the absorption band and by the appearance of plasmonic bands at 377 nm. Moreover, the intensity of the bands became stronger as the synthesis temperature increased [74]. The effect of the synthesis temperature was further illustrated by the variation in ZnO morphology. The ZnO NPs that were synthesized at 30 °C were generally irregularly shaped, and they exhibited polydisperse-characteristics that are common in amorphous semiconductor matrices. As the temperature increased, the morphology became clearer, such that the ZnO NPs that synthesized at 60 °C were spherical, and those that synthesized at 100 °C were dumbbell shaped. Previously, Das et al. [158] noted that the effects of different temperatures could be observed through the degree to which the molecules associated with the synthesized biological media remained attached to the surface of the nanoparticles. The authors found that synthesis at 60 °C enabled numerous bio-compounds on gold to remain attached to the surface of NPs to a greater extent than synthesis at 100 °C. When Sukri et al. [84] prepared ZnO NPs using Punica granatum fruit peel, the TEM revealed mostly spherical and hexagonal ZnO NPs, and the mean sizes were dependent on calcination temperatures. The sizes were 32.98 nm and 81.84 nm when the samples were calcined at 600 °C and 700 °C, respectively.

2.3.2. Effect of pH

The metabolites that exist in biological extracts are pH sensitive; hence, variation in terms of medium tampers with their reducing and chelating capabilities, which directly affects the growth of the nanomaterials [159]. Furthermore, pH influences the redox activities of enzymes and other metabolites in the cell-free extract [123]. The changes in the acidity of the reaction medium also influences the NPs’ yield and morphology. Al-Kordy et al. [126] reported that the average diameter of ZnO decreased gradually as the pH increased, but at higher pH levels, a slight increase in size was noticed when an extract of Alkalibacillus sp. W7 was used. This was due to the fact that the chelating/growth controlling compounds, that were present during the formation of the Alk-ZnO NPs, were denatured at an extreme pH [160]. Guan et al. [161] reported that a pH value of 8.5 encouraged the production of almost spherical hybrids, with an average diameter of 60 nm, because of the low OH ion concentration; however, at pH values of 10.5 and 11.0, sheet-like and flower-like hybrids were obtained because of the large quantity of OH ions. The hybrid nanomaterials exhibited a high degree of crystallinity and a high ZnO content, thus leading to higher thermal stability.

2.3.3. Effect of the Concentration of the Precursor

The concentration of biomass/extract and salt influences the morphology, and enhances the production, of nanoparticles [162]. When the zinc ion concentration increased from 0.005 M to 0.3 M, the formation of NPs occurred more quickly, but a lower yield of hexagonal ZnO NPs was obtained, with diameter sizes ranging between 20.7   and   96.5 nm. SEM images confirmed that the concentration of biosynthesized ZnO NPs occurred rapidly, and larger particles formed due to the competition between Zn2+ and the functional moieties found in the cherry extract, which served to accelerate the reduction [163]. This finding confirms the observations made by Jamdagni et al., who observed an increase in the absorbance of the UV-vis spectra as the precursor concentration increased [164]; however, further increases in concentration yielded a broad and low absorbance peak that signified a low yield of ZnO NPs. The sizes of the ZnO NPs obtained with 1% Hibiscus sabdariffa extracts ranged between 20 and 40 nm, whereas the 8% extract samples exhibited homogenous shapes and smaller NPs (5–12 nm) due to the abundant chelating compounds in the extract, thus resulting in very stable ZnO NPs [87].

2.3.4. Effect of Reaction Time

According to Fagier et al. [69], the biosynthesis of ZnO NPs is entirely determined by the temperature, pH, biological source, and zinc precursor, rather than the reaction time; however, Dhandapani et al. [165] observed that the heating duration at a particular temperature determines the formation of aggregates. For instance, upon increasing the heating duration for a reaction conducted at 50 °C, from 30 min to 90 min, the agglomeration of the particles increased. Prolonging the reaction time (36 h) during the fabrication of ZnO NPs using Pichia kudriavzevii resulted in aggregates with irregularly shaped NPs, whereas reaction times of 12 h and 24 h yielded smaller NPs [136]. These findings are consistent with observations of chemical methods, where extended nucleation times resulted in the generation of larger ZnO NPs [166].

2.3.5. Effect of Different Biological Species

Mahdi et al. prepared ZnO NPs using different bacteria, namely, Lactococcus lactis and Bacillus sp. [30]. The particles synthesized by L. lactis showed an absorption band at 444 cm−1 with a distinct shoulder at 486 cm−1, which is a typical ZnO absorption band, whereas the Bacillus sp. sample displayed absorption bands at 560 cm−1 and 601 cm−1. These changes in wavenumber, peak shape, and intensity indicate the changes in ZnO morphology and the size of particles. This finding was confirmed by the SEM, which revealed that the ZnO NPs synthesized by L. lactis were 55 nm spheres, whereas the ZnO NPs produced by Bacillus sp. exhibited diameters of 99 nm and they were rod shaped. In a separate study, Mohamed et al. [167] prepared ZnO NPs using extracts of two fungal strains, Fusarium keratoplasticum (A1–3) and Aspergillus niger (G3–1). The F. keratoplasticum strain (A1–3) yielded hexagonal (10–42 nm) NPs, whereas the A. niger strain (G3–1) resulted in nanorod-shaped (8 38 nm) NPs with −39 mV and −35 mV Zeta potential values for the hexagonal and nanorod ZnO NPs, respectively.

3. Characterization Techniques for ZnO

Having demonstrated the fact that a wide spectrum of morphological possibilities arises from the variation of synthesis parameters and plant extracts, it is necessary to study the ZnO NPs that are formed in greater detail. Although there are a wide variety of characterization techniques, only a few are needed to illustrate the story of ZnO green synthesis juxtaposed with the classical synthesis techniques. Among these techniques, the most important include the ultraviolet visible spectroscopy (UV-vis), X-ray diffraction (XRD), SEM, TEM, energy-dispersive X-ray spectroscopy (EDX), and FTIR. These are discussed in relation to the information that they provide regarding the uniqueness of green-synthesized ZnO NPs.

3.1. UV-Vis Spectroscopy

UV–Vis is a preliminary technique that is used to confirm the formation of the nanomaterial [140]. Ideal ZnO NPs display a characteristic absorbance peak between 320–390 nm which corresponds to the E g (3.37 eV) of the material [133]. The inherent   E g absorption is attributable to the electron transitioning from the valence band to the conduction band (O2p → Zn3d) [168]. The peak position, shape, and width is influenced by the structure and particle size, oxygen vacancies, E g , and so on. [110]. The strong absorption of ZnO in the UV region warrants its applicability in the photocatalytic degradation of pollutants. Abdelhakim et al. [34] produced an absorbance peak of 369 nm with samples that were prepared using a culture filtrate of the endophytic fungus Alternaria tenuissima. The absorbance peak for Phoenix dactylifera waste-ZnO-NPs was found to be 381 nm. The red-shift was ascribable to the surface plasmonic resonance of the sample, which was determined by its morphology and size [39]. Doped ZnO samples also displayed the red-shifted position of the absorbance peak, as well as an improved intensity. This is a result of the decreased E g , which is facilitated by the dopant, as observed when using Nb-ZnO that is prepared using the leaf extract of Vernonia amygdalina [111]. The same finding was reported in [110], which mentioned that increasing the dopant concentration reduced the E g in Mg and Cu co-doped ZnO NPs; this is because the dopants introduce dopant-induced states that provide additional pathways for electronic transitions. Although broad peaks indicate polydisperse NPs, narrow peaks suggest monodispersed ZnO NPs [110].

3.2. X-ray Diffraction (XRD)

XRD is used to ascertain crystallinity, confirm the respective crystallographic planes, and determine the purity of nanomaterials. Most ZnO diffraction peaks are indexed to the hexagonal phase in accordance with the wurtzite structure, which is the most stable phase in ambient conditions [169]. The wurtzite structure’s peaks are approximately located at 2θ~31.7, 34.4, 36.3, 47.6, 56.7, 62.8, 63m and 69.5°, which are reflected in the lattice planes of (100), (002), (101), (102), (110), (103), (112), and (202), respectively [25,170,171]. In the case of bio-prepared samples, additional peaks are often located, and these arise from the crystallization of the biomolecules. As such, a peak was observed at 23°, which was due to crystallized chitosan [171]. In a separate study, Kaningini et al. [170] also identified extra peaks that were thought to have been reflected from excess crystallized extract. Moreover, ions such as K, S, and Cl, which were previously absorbed by the plant from the soil, were also identified; however, annealing at high temperatures (800 °C) removes the additional diffraction peaks from the various ions. Doped ZnO samples display high intensity peaks compared with undoped ZnO. The enhanced intensity suggests an increase in the packing of atoms in their respective planes. Moreover, the doping of ZnO samples usually leads to a red shift in the 2θ position, as observed when doping ZnO with Ce [170]. The red shift is consistent with the reduction of E g , as the dopants are incorporated into the lattice of ZnO. Annealing the obtained samples also improves the crystallinity, and thus, the intensity, of the Bragg’s peaks since high temperatures repair the structural defects of the material [172]. Similarly, Hafiz et al. [173] observed that the (002) plane diffraction peak for Al-ZnO had a higher intensity compared with the unannealed sample.
The X-ray diffraction peaks of the biosynthesized ZnO samples displayed broadening, which can be attributed to the nano-scaled size of the samples; this was accompanied by fewer internal stresses within the sample [58]. In most of the green synthesis processes that were reported, the procedure ends with the calcination of amorphous ZnO in order to transform it into crystalline ZnO; as such, the high degree of crystallinity that was observed is more likely to be due to the high temperature than the use of green biosynthetic protocols. This is expected since annealing repairs the structural defects of synthesized ZnO; however, a few studies have developed highly crystalline ZnO without post synthesis calcination, or by just drying the amorphous ZnO at temperatures as low as 50 °C. For example, Khalifu et al. synthesized ZnO NPs using a Chlorella microalgae extract, and they only dried the resultant ZnO at 50 °C [140]. The XRD spectra showed intense peaks that were attributed to ZnO, and no traces of amorphous material, as indicated in Figure 9a. Another report showed that drying the amorphous ZnO at a low temperature (60 °C), after being treated with zinc acetate and Cyanobacteria extract, produces ZnO with good crystallinity, albeit with some degree of amorphous composition [29] (Figure 9b). XRD spectra has also unearthed the dependence of the catalytic efficiency of ZnO on the ZnO facet that is most exposed. It has been shown that ZnO NPs with limited growth in the 0001 direction show a higher catalytic activity due to the facets on the 0001 plane that are rich with terminal oxygens [102,174,175] (Figure 9c). In the XRD spectra, the growth in the 0001 direction can be inferred from the intensity of the 002 peak, as when the 002 peak has a low intensity relative to the 101 peak, there is limited crystalline growth in the 0001 direction, and the ZnO has high catalytic activity.

3.3. Scanning Electron Microscopy (SEM) and Transmission Electron Microscopy (TEM)

SEM provides information on the topography and morphology of the NPs; however, currently, researchers prefer to use TEM because it offers greater magnification and resolution capabilities than SEM, and as a result, the information regarding the size, shape and crystallography of the nanoparticles from TEM is more accurate [64]. The size and morphology of ZnO NPs play crucial roles in terms of their properties and applications. Smaller NPs possess a higher surface area than larger ones; however, ZnO NPs form aggregates because smaller particles possess high surface energy that allows them to interact with each other, possibly via hydrogen, intermolecular Van der Waals, and electrostatic bonding during the synthesis and drying processes [25]. With respect to antimicrobial properties, small NPs penetrate microbial membranes with ease, thus enhancing their killing efficiency. Moreover, the release of the toxic Zn2+ and hydrogen peroxide is also enhanced by larger surface areas. The photocatalytic degradation of pollutants also relies on the size and uniformity of the nanomaterials. The morphology of the nanomaterials is affected by several factors, including the concentration of the salt precursor, as observed in the SEM analysis in [40], wherein small dispersed spherical NPs at 5 mmol/Kg, agglomerated sponge-like morphologies at 10 mmol/Kg, and nano capsules at 50 mmol/Kg were obtained, respectively. The TEM images were similar to those obtained using SEM, but using high resolution TEM (HRTEM), at 5 mmol/Kg, hexagonal crystalline ZnO NPs were attained; at 10 mmol/Kg, hexagonal ZnO NPs agglomerated to form short rods; and at 50 mmol/Kg, nanorods were formed. In retrospect, the precursor tunes the structure and morphology of ZnO NPs without altering the crystallite structure [40]. In another study, the SEM revealed uniformly dispersed spherical NPs. The diameters of the NPs decreased from 78 to 64 nm, with an increase in the concentration of Syzygium cumini leaf extract, from 10 to 25 mL. The size reduction occurred as a result of the capping efficiency of the extract during the formation of NPs [47]. Using the TEM technique [105], polydisperse ZnO NPs were obtained with sizes ranging between 15 and 30 nm. According to the surface area electron diffraction (SAED), the NPs exhibited a hexagonal crystalline structure. The high crystallinity of the ZnO NPs was confirmed by the presence of bright spots that formed a hexagonal pattern.

3.4. Energy-Dispersive X-ray Spectroscopy (EDX)

EDX is commonly used to analyze the elemental constituents and composition of NPs; therefore, in bio-prepared samples it is mainly used confirm the sample’s purity. In ultrapure ZnO samples, only the elements Zn and O are observed in the EDX spectrum; however, it is not uncommon for the EDX to detect the presence of Cu, which, in most cases, is due to the copper grid upon which the samples are often loaded. Additional C signals may originate from the carbon used to coat the grid, or alternatively, from the extra biocompounds that are adsorbed onto the ZnO NPs [169]. In samples synthesized with nitrogen-containing extracts such as chitosan and algae, it is common to observe the N signal arising from the (-NH2) group [171]. If impurities and/or dopants are present, they are detected and shown in the spectra [108]. Samples that are annealed with oxygen reveal high levels of O compared with unannealed samples, or those annealed under controlled conditions [173].

3.5. Fourier-Transform Infrared Spectroscopy (FTIR)

FTIR provides information on the surface chemical functional groups, as well as bonding between the nanomaterials and biological extracts. The characteristic band for undoped and doped ZnO NPs lies between 400 and 700 cm−1. This peak is due to the presence of the Zn-O stretching vibration [29]. The specific peak position and intensity of the absorption bands depends on the synthesis method and reducing agents. In the case of biosynthesized ZnO NPs, the position is influenced by the available biocompounds in the extract. Chemically prepared ZnO samples display a sharp absorption band at 536.86 cm−1 [176], whereas this characteristic peak appeared at 547.38 cm−1 for ZnO samples reduced with the extract of Cyanobacterium Nostoc sp. EA03 [29]; however, Shabaani et al. [25] prepared ZnO using the Eriobutria japonica seed, and they observed the peak at 465.24 cm−1, whereas it was positioned at 457 cm−1 in the study presented in [177]. The other dominant absorption bands in the FTIR for bio-prepared ZnO NPs are found in the region of 800–1500 cm−1; these are attributable to organic compounds (phenolic groups, alcohols, and aliphatic amines). The absorption peaks in the region of 1600–3400 cm−1 are the result of the stretching bands of C=O in the amide group, and the presence of the −COOH and −OH groups, respectively [29]. If the extract predominantly contains proteins, bands at around 1650 cm−1 (amide I, C=O stretching) and 1550 cm−1 (amide II, C-N stretching and N-H deformation) become apparent. The N-O bands appear at 1380 cm−1 [25]. Band shifts and changes in intensities indicate some form of chemical bonding between the NPs and the reducing agents.

3.6. Other Techniques for ZnO Analysis

The use of biomolecules to tailor ZnO also necessitates the determination of other aspects of ZnO NPs such as the chemical composition, bond order, surface area, interactions with liquids, and pollutants. X-ray photoelectron spectroscopy (XPS) is useful in determining the intrinsic chemical structure of ZnO NPs, and this is particularly the case regarding doped ZnOs. With XPS, the bond order of ZnO and the presence of defects can be determined with accuracy, and therefore, doping efficiency in green synthetic processes in comparison to conventional synthesis processes can be determined. The surface area of nanomaterials is determined using Brunauer–Emmett-Teller (BET) analysis with an inert gas such as N2 to probe the available surface area [178]. The only limitation in terms of surface area, as represented by BET analysis, is that it is only representative of the surface area of the dry sample yet ZnO is generally used in aqueous conditions in cases of water remediation. To counteract this limitation, the BET surface area is used in conjunction with the particle size and point of zero charge determination using dynamic light scattering (DLS) techniques. Although the DLS technique does not provide the size of the surface area, it does determine the size of the ZnO NP and the aggregation index in water. The BET technique also enables the prediction of the type of interaction between the ZnO and the pollutant (i.e., whether it is chemisorption or physisorption).

4. Application of Green ZnO NPs in Water Remediation

Traditional water treatment approaches such as chlorination, ultraviolet irradiation, and ozonation are undertaken in water treatment plants to reduce the pathogenic bacterial load. Although these techniques effectively decontaminate the water, the oxidizing agents typically react with natural organic matter to form disinfection by-products, especially halogenated organics which are renowned for their carcinogenicity [179]. Nanotechnology, on the other hand, can potentially enhance the treatment efficiency of wastewater and eliminate the threat of incurring toxic by-products, thereby nudging water treatment technologies a step closer to becoming green technologies. Since ZnO NPs exhibit salient physicochemical properties, they have been explored for water decontamination purposes, including the photocatalytic and adsorptive removal of pollutants such as dyes, heavy metals, pharmaceuticals, and microorganisms [180].

4.1. Photocatalytic Degradation of Pollutants

The decontamination of wastewater via the photocatalytic break down of microbes is superior to physical, chemical, and biological decontamination methods. This is because photocatalytic degradation is cost-effective and environmentally benign; it involves a rapid oxidation/reduction process and it can eliminate polycyclic products through mineralization [181]. In this sense, the catalytic process fulfils one of the twelve pillars of green synthesis and one of the green processes, specifically that which relates to the acceleration of a reaction using a catalyst and reduction of energy demands. ZnO nanostructures have been extensively applied as semiconductor photocatalysts owing to their high reactivity, abundant reactive sites, high yields of hydrogen peroxide, affordability, and non-toxicity [182]. The photocatalytic mechanism of action for ZnO is represented in Figure 10. Briefly, when irradiated by photons with an energy that is above the photo-threshold, the electrons (e) are excited from the valence band (VB) to the conduction band (CB). This forms positively charged holes (h+) for every excited e [183] (Equation (1)). The newly formed electron-hole (e/h+) pairs migrate to the surface of the semiconductor (ZnO) where the e reduces the adsorbed O2 to generate superoxide anion ( O 2 . ) (Equation (2)). The h+ reduces H2O and OH to generate H2O and H2O radicals, respectively (Equations (3) and (4)). These radicals and the superoxide ions are known as reactive oxygen species (ROS) and they are responsible for the degradation of pollutants [47].
Photoexcitation :   ZnO + hv e cb + h vb +
Oxygen   adsorption :   O 2   abs + e O 2
Water   ionization :   H 2 O OH + H +
Superoxide   protonation :   O 2 . + H +   HOO .
e +   HOO .   +   H +   H 2 O 2
H 2 O 2 +   e   OH . +   O H
ZnO photocatalysts are mainly applied to decolorize water contaminated by synthetic, industrial dyes; this is because dyes cannot be completely degraded via aerobic approaches, thus rendering biotreatment ineffective [184]. ZnO-based photocatalysis has been shown to be capable of deactivating a wide variety of microbes and degrading organic pollutants, as depicted in Figure 10. Biosynthesized ZnO NPs have demonstrated photocatalytic efficiency matching that of conventionally synthesized ZnO, and reports concerning the photocatalytic degradation of pollutants using biosynthesized ZnO NPs are detailed in Table 5.
Figure 10. Mechanism of the degradation of pollutants by ZnO NPs. (Modified with permission from Ref. [185] © 2013, Elsevier B.V.).
Figure 10. Mechanism of the degradation of pollutants by ZnO NPs. (Modified with permission from Ref. [185] © 2013, Elsevier B.V.).
Catalysts 12 01442 g010
ZnO nanoparticles generated by phytosynthesis, using plants such as Lycopersicon esculentum [186], Calliandra haematocephala-L. [187], and Abelmoschus esculentus L., have been applied in the photocatalytic decontamination of methylene blue (MB) [188]. Spherical ZnO NPs were synthesized by Nava et al. using Camelia sinesnsis extract and degraded MB at under a UV light after 120 min [189] with an 84.37% degradation efficiency. Other dyes that have been degraded by plant-based synthesized ZnO NPs include Congo red (CR), which was degraded by Artocarpus heterophyllus-reduced NPs [190], and Rhodamine 6 G (R6G), which was degraded by Syzygium cumini-stabilized ZnO NPs [47]. In addition to dye degradation, ZnO NPs that were synthesized using plant-mediated methods were used by Razanamahandry et al. [191] in order to break down free cyanide, a very toxic pollutant that is generated by metallurgical activities. When spherical ZnO NPs (9.8 nm) were synthesized using Eucalyptus globulus leaves, a 98% efficiency degradation of 3 g CN L1 under solar light was achieved, and a 45% efficiency degradation occurred under UV light within 20 min.
Table 5. Summary of the reported photocatalytic activities of ZnO (doped and undoped) nanoparticles synthesized from various biological sources.
Table 5. Summary of the reported photocatalytic activities of ZnO (doped and undoped) nanoparticles synthesized from various biological sources.
Name of Biological SourceZnO MorphologyPollutantCatalyst ConcentrationPollutant ConcentrationRadiationExposure Time (min)pHEfficiency (%)Ref
Plant based
Lycopersicon esculentum,PolyhedralMethylene blue (MB)_15 mg/L10 W
UV light
180697[186]
Calliandra haematocephala-L. Nanoflower-likeMB50 mg/100 mL20 mg/LSunlight270688[187]
Abelmoschus esculentus L. SphericalMB and methyl orange (MO)15 mg/mL10 mg/10 mLUV light270496[188]
Camellia SinensisSphericalMB15 mg/L200 mg/L10 W
UV light
120684.37[189]
Artocarpus HeterophyllusSphericalCongo red (CR)24 g/L20 mg/L30 W
UV light = 2
609 > 90[190]
Eucalyptus globulusSphericalFree cyanide0.04 g/100 mL0.3 g/LUV light20unspecified98[191]
Azadirachta indicaSpherical and flower-likeMB10 mg/mL10 mg/LSunlight120697[192]
Rosmarinus officinalisQuasi-hexagonalIndustrial dye effluent50 mg/50 mL TE20 W
Vis light
1005.2863[193]
Synadium grantiiSphericalMB
Indigo carmine
RhB
100 mg/L in 100 mL25 mg/L30 W
UV light
75
75
75
unspecified79.2%
80.8%
75%
[194]
Algae
ChlorellaSphericalDBT0.01 g/20 mL10 mgL/20 mLUV light180797[140]
Sargassum specieIrregularMalachite green (MG)0.06 g5 mg/L150 W
Vis light
60892.5[143]
Fungi
Cordyceps militarisFlower-likeMB0.25 mg/mL10 mg/LUV light180697[195]
Saccharomyces cerevisiaeSphericalMB12 mg/mL5 mg/LUV light200367[138]
Saccharomyces cerevisiaeSpherical4-nitrophenol50 mg/L1.5 × 10−4 MUV-Vis light100678[196]
Aspergillus niger.Near sphericalBismarck Brown 1 × 10−4 MVis light72 hunspecified89[197]
Biological derivatives
PullulanSpherical and hexagonalAmoxicillin
Paracetamol
50 mg/100 mL30 mg/L9 W
Mercury lamp
1209
5
85.7
96.8
[147]
Artemia eggshellSpherical or ellipsoidalRhB
MB
Neutral red
10 mg/100 mL4 mg/10 mg/mL
(10 mg/mL)
1203–1188.75
93.3
86.6
[150]
ChitosanFlower-likehexavalent chromium25–100 mg50 mg/L (50 mL)32 W
UV light
365
605.299.2[198]
Cellulose FibersFlower-likeMethyl orange
Phenol
0.3 g/15 mL0.06 mmol/L (15 mL)
60 mg/L (15 mL)
6.6 W
UV light
365
9 hDye100[199]
Co-doped ZnO supported in algae (Sargassum sp.) was synthesized for the photocatalytic decolorization of malachite green (MG) dye under visible light. The photocatalytic performances were enhanced by 58% for ZnO/algae, and by approximately 75% for Co–ZnO/algae, compared with pure ZnO without support, for 60 min. Rajaboopathi and Thambidurai [142] degraded MB and reactive blue 198 (RB-198) over 150 min using Padina gymnospora Ag-doped ZnO NPs, with 98.7% and 99.8% efficiency, respectively. Ag-doped ZnO NPs, developed with the aid of Azadirachta indica extract via a hydrothermal method, were obtained by Patil et al. [192]. The degradation efficiency of 10 mg/L MB dye and a 10 mg/mL catalyst concentration was 97% over 120 min under natural light, as shown in Figure 11a. The rate constant for degrading MB using the Ag- ZnO NPs was 5.9668 × 10−2 min−1, and it was higher compared with that of the undoped ZnO (2.527 × 10−2 min−1) in Figure 11b. Rosmarinus officinalis yielded quasi-hexagonal Ag-doped ZnO NPs that were used to degrade an industrial dye under visible light at pH 5.28 with a 63% efficiency [193]. Cu-ZnO NPs were synthesized from Synadium grantii leaf extract. The resultant composites degraded MB, Indigo Carmine (IC) and RhB dyes with a 79.2%, 80.8%, and 75% efficiency, respectively [194].
ZnO NPs that are prepared with the aid of fungi are known to be viable for pollutant degradation. As such, Cordyceps militaris chelated flower-shaped ZnO NPs degraded 10 mg/L MB, with a 97% efficiency over 180 min, when irradiated with UV light [195], whereas Saccharomyces cerevisiae-stabilized NPs achieved only a 67% efficiency when degrading 5 mg/L MB under UV irradiation over 200 min [144]. Previously, Kumar et al. prepared 50 mg/L ZnO NPs using the biomass of Saccharomyces cerevisiae as a template to break down 1.5 × 10−4 M 4-nitrophenol (4-NP) [196]. The pollutant was degraded in 100 min under UV light with 78% efficiency. The precipitation process was used to prepare biogenic ZnO NPs utilizing pullulan as the biomaterial. The material photodegraded amoxicillin and paracetamol with 85.7% and 96.8% efficiency, respectively. The pharmaceuticals were destroyed with ring opening, oxidation, and/or the breaking down of molecules [147]. Biological chitosan was used to chelate ZnO NPs for the photocatalytic breakdown of Cr(VI) [198], whereas cellulose-stabilized ZnO NPs have been observed to completely degrade both phenol and MO after 9 h [188].

4.2. Adsorption of Heavy Metals and Organic Dyes

Adsorption is renowned for its simplicity and its versatility, and it can remove contaminants, whether in trace amounts or high concentrations, from wastewater [200]. Adsorption depends on the surface area and the type of adsorbate as well as the conditions under which adsorption is taking place [201]. The operation conditions (pH, temperature, contact time, and so on.) influence the surface charge of the adsorbent and the target pollutant, and as such, they provide a tailored adsorbent surface to target specific contaminants. During the adsorption process, the adhesion of adsorbates or pollutants to reactive sites on the adsorbent surface takes place through either physisorption or chemisorption [4]. ZnO nanoadsorbents are suitable candidates for removing pollutants from water because they are environmentally friendly, cost-efficient, and biocompatible [202]. Mahdav et al. state that ZnO NPs are potential adsorbents due to their high adsorption capacity [203]. Non-functionalized wurtzite ZnO NPs bear a positive charge; however, the surface functionalization that takes place during biogenic synthesis results in NPs with negatively charged surfaces; this is due to the negatively charged functional groups [204]. Electrostatic attraction occurs between these negatively charged surfaces and positively charged pollutants, as indicated in Figure 12a. The mechanism of ZnO nanoadsorbents, with regard to dye removal, depends on the surface charge of the dye. It can be characterized using a point of zero charge (pHpzc). For instance, [205] discovered that the pHpzc of Ocimum sanctum which stabilized the ZnO nanoadsorbents was pH 6.2; therefore, if the pH was above 6.2, the surfaces of the ZnO NPs were negatively charged, thus resulting in electrostatic repulsion with the anionic Congo Red dye molecules, but when the pH was below the pHpzc, the surface of ZnO NPs was positively charged. The strong electrostatic interactions between ZnO NPs and anionic dye molecules are shown in Figure 12b. Nanostructure aggregates are superior nanoadsorbents to simple nanopowders because of their high porosity and ease of separation from a solution [206].
Adsorption isotherms are crucial for explaining the process of distribution concerning the adsorbates between the solid and liquid phases. The parameters of isotherm models are generally used to determine the adsorption capacity of adsorbents, and the suitability of adsorption under a batch system [208]. The adsorption of HM ions such as Cr6+ Hg2+, Cd2+, Co2+, and Pb2+ using ZnO is best explained with the Langmuir isotherm. This suggests that the HM ions are distributed homogeneously on the ZnO adsorbent, which subsequently forms a monolayer. Modifying ZnO via doping and surface functionalization alters the isotherm model to Freundlich, Dubinin–Radushkevich (pore filling), and Redlich–Peterson (features of Langmuir and Freundlich) isotherms for Cd2+, Co2+, Cr6+, and Pb2+. In most cases, ZnO nanoadsorbents follow pseudo-second-order (PSO) kinetics, thus suggesting that chemical interactions between the contaminants and the ZnO adsorbent are favored [209].

4.2.1. Heavy Metal Adsorption Using ZnO NPs

Biosynthesized ZnO NPs have been applied in order to remove heavy metals. The natural occurrence of hexavalent chromium (Cr(VI)) is not common; therefore, the main contributors to Cr(VI) pollution are industrial processes [210]. An excess of the Cr(VI) ion in water causes respiratory tract irritation, different types of cancers, and even death; therefore, it is imperative to eradicate Cr(VI) from wastewater before it is discharged into the environment. When used with a 25 mg/L pollutant, guar gum biosynthesized ZnO NPs and removed Cr(VI) with a 98.63% removal efficiency (RE) and a 55.56 mg/g adsorption capacity. Most importantly, optimal adsorption occurred at pH 7, which is the natural pH of natural water in bodies [211]. Parthenium hysterophorus L. activated carbon-prepared ZnO NPs, and 1000 mg/L Cr(VI) at pH 4 was removed with a 98.8% removal efficiency over 180 min. MB was removed from industrial wastewater with the same nanoadsorbent over 150 min [212]. A lead ion (Pb2+) is a very toxic HM. Pb2+ exposure has the potential to induce gastrointestinal, cardiovascular, neurological, hematological, and renal problems [213]. As a result, the World Health Organization (WHO) requires that the content of Pb2+ in industrial effluents is reduced to an acceptable level (0.01 mg/L) [214]. Somu and Paul prepared ZnO NPs, using casein as a reducing agent, and reported a 95.3% RE with (2 g/L) ZnO NPs in 100 mL water with 100 ppm Pd2+ at pH 8 over 120 min [215]. Based on the Langmuir model, the maximal adsorption capacity was 194.93 mg/g. This adsorption capacity was higher than the 3.41 mg/g reported by [216], wherein the NPs were reduced and stabilized with Mangifera indica leaves at pH 6 in 50 min. When Shorea robusta capped NPs were employed for the removal of Pb2+, with 0.8 g adsorbed in 1000 mg/L pollutant at pH 6, the RE was 92.9% [208]. Table 6 presents a summary of the reported adsorption activities of ZnO nanoparticles synthesized from various biological sources.
Cd2+ is another extremely poisonous HM. The set maximum limit of Cd2+ in drinking water is 0.003 μg/L [214]. Excess Cd2+ consumption causes high blood pressure, bone marrow malignancy, renal damage, gastrointestinal issues, and an increased chance of stillbirth, among other problems [217]. In a casein-based ZnO nanoadsorbent, the adsorptive capacity of Cd2+ was found to be 156.74 mg/g with an 85.63% RE [215]. Joshi and Singh [208] reported an RE of 84.6% within 50 min when removing Cd2+ from water containing 50 mg/L of the pollutant and 1 g of the nanoadsorbent at pH 6. The RE reported by [216] when the adsorbent was synthesized with Shorea robusta (leaf) was 89.9% at pH 6 with a pollutant concentration of 1000 mg/L.

4.2.2. Adsorption of Organic Dyes Using ZnO NPs

ZnO NPs synthesized with leaf extracts of Hibiscus rosa-sinensis were utilized in the removal of Congo red (CR) from water. The optimum adsorption occurred when 4 mg/L of the dye was used, the adsorbent dosage was 0.05 g, and at pH 4. [218]; however, Ocimum sanctum leaf extract prepared with ZnO removed CR at an adsorbent dosage of 0.2 g/L concentration and at pH 4, with a removal efficiency of 97% and a 74.07 mg/g adsorption capacity [205]. Curcumin-conjugated ZnO NPs were synthesized in [219] for CR removal, and it was discovered that the addition of KOH to the reaction media improved the stability of the formed nanoparticles. Moreover, the addition of chitosan further enhanced the NPs’ stability with only a 10% loss in mass. The registered adsorption capacity of Zn(Cur)O was 89.85 mg/g, and the pseudo-second-order model was followed. Aloe vera-stabilized ZnO nanosheets removed 76.73% of the initial MG dye concentration after 180 min, whereas the adsorption kinetics fitted a pseudo-first-order model with an r2 factor of 0.98984 [220]. Gum Arabic-grafted polyacrylamide (GA-cl-PAM) served as a self-template for the in-situ synthesis of ZnO NPs in order to remove MG from an aqueous solution. The nanocomposite adsorbed ~99% MG dye with a 0.4 g/L adsorbent dose at a neutral pH. A high adsorption capacity of 766.52 mg of dye per gram of adsorbent was attained at 25 °C [221]. An aqueous leaf extract of Becium grandiflorum was used in the friendly synthesis of ZnO NPs. The reported MB removal efficiency was 71.53% when both the adsorbent mass and dye concentration was 25 mg/L. The synthesized NPs showed an adsorption capacity of 143.6 mg/g after 180 min [222]. Sawdust-derived cellulose nanocrystals, incorporated with ZnO NPs, were prepared in [223] at pH 4 in order to act as an adsorbent for MB removal. The MB adsorption was better characterized by the Langmuir isotherm which registered an adsorption capacity of 64.93 mg/g after 12 h. ZnO NPs were impregnated into Ananas comosus waste (Acp) for the adsorption of Celestine Blue (CEB), a dye that is used in the nuclear and textile industries. The adsorption process was best characterized by the Langmuir and pseudo-second-order models (R2 > 0.9518). The biomass enhanced the process and exhibited endothermic characteristics. An RE of 70.2% was recorded at pH 9 after 120 min of reaction [224].
Table 6. Summary of the reported adsorption activities of ZnO nanoparticles synthesized from various biological sources.
Table 6. Summary of the reported adsorption activities of ZnO nanoparticles synthesized from various biological sources.
Name of Biological SourceMorphology of ZnOPollutantConcentration of AdsorbentConcentration of PollutantTime
(min)
pHRemoval Efficiency (%)Isotherm
Model
Adsorption Capacity (mg/g)Ref.
Heavy metals
Guar gumSphericalCr (VI)1.0 g25 mg/L50798.63Freundlich55.56[211]
Parthenium hysterophorus L.Hexagonal and sphericalCr (VI)
MB
50 mg/L100 mg/L180
150
4
6
98.8 [212]
CeisenQuasi-sphericalCd
Co
Pb
(2 g/L)/100 mL50–500 mg/L1207
8
85.63
71.23
95.33
Langmuir156.74
67.93
194.93
[215]
Mangifera indica (leaves)Semi-sphericalCu2+
Pb2+
Cd2+
1.0 g50 mg/L50689.1
92.1
84.6
Freundlich2.35
3.41
2.65
[208]
Shorea robusta (Leaf)Smaller clustersPb2+
Cd2+
0.8 g1000 mg/L50692.9
89.9
[216]
Organic dyes
Hibiscus rosa-sinensis.Spherical, cubic, and hexagonalCR0.05 g4 mg/L20495.55Langmuir [218]
Ocimum sanctum (leaf)SphericalCR0.2 g40 mg/l30497Langmuir74.07[205]
CucuminSphericalCR0.15 g11.15 mg/L24 h 89.85[219]
Aloe vera (Leaf)NanosheetsMG10 mg/500 mL10 mg/L18076.23 [220]
Gum ArabicFlower-likeMG0.4 g/L50–800 mg/L60799 766.52[221]
Becium grandiflorum (Leaf)SphericalMB25 mg/L25 mg/L,18071.53 143.6[222]
CelluloseSphericalMB0.15100 mg/L1440497.5Langmuir64.93[223]
Ananas comosusIrregularCelestine blue (CEB)0.1550 mg/L120970.2Langmuir6.52[224]

4.3. Antimicrobial Activity of ZnO NPs

Microorganism-infested bodies of water are the main sources of infectious diseases affecting people and animals. The infectious microorganisms are bacteria, fungi, algae, viruses, and parasitic protozoans. Antimicrobial activity refers to the ability of a material to either kill microbes or inhibit the growth of infectious agents and antimicrobial agents [7]. Green-synthesized antimicrobial nanomaterials meet this criterion, and they have the added advantage of being non-toxic to living organisms. These green synthesized nanomaterials, such as ZnO, owe their non-toxicity to their small size, which favors internalization into the cells of microorganisms [225]. In addition, the residual surface functional groups comprising the biological capping agents may imbue the ZnO with enhanced capabilities to attach onto the cell walls/cell membranes of the disease-causing microorganisms.
The reaction mechanisms of ZnO NPs, regarding their function as antimicrobial agents, are broadly classified into three types: (1) the generation of ROS via photocatalysis, as discussed in Section 4.1; (2) the dissolution of ZnO NPs into Zn2+; and (3) the direct interaction between ZnO NPs and cell membranes. Generation of ROS within and in the periphery of the bacterial cell is arguably the main contributor to antibacterial activities as shown in Figure 13a. In addition, the H2O2 generated during photocatalysis penetrates the bacterial cell membrane and damages organelles through the hydrolysis of the constituent molecules such as lipids, proteins, and DNA, thus causing cell lysis [49] (Figure 13b). With respect to the dissolution of ZnO, studies suggest that the dissolution of Zn2+ is caused by size and morphology; indeed, morphologies that offer a large surface area, such as nanosheets and nanorods, have a greater surface area that is in contact with the water, and therefore, they can release the Zn2+ ion more quickly (Figure 13c). The better morphologies for use as bactericides therefore tend to be smaller nanoparticles. Secondly, the ZnO NPs with a large, exposed surface area easily penetrate bacterial cells and tamper with active transport and enzyme activity, thereby inhibiting the growth of the bacteria [226] (Figure 13a). Moreover, a large number of NPs is required to envelope a bacterial colony, which leads to innumerable ROSs [227]. With respect to the direct interactions between ZnO NPs and cell membranes, Stoimenov et al. [228] concluded that Gram-negative bacteria are more susceptible to inactivation by ZnO NPs because they harbor an overall negative charge due to the presence of the –COOH and –OH groups on their cell walls. ZnO NPs, on the other hand, have a positive charge when placed in an aqueous medium due to the fact that they become partially dissolved. These opposite charges encourage electrostatic interactions (Figure 13d) that damage the plasma membrane and cause a leakage of organelles. Furthermore, the build-up of ZnO NPs in the cytoplasm disrupt bacterial metabolic activities, thus leading to their inactivation.

4.3.1. Antibacterial Activity of Biologically Prepared ZnO NPs

The antibacterial activity of biogenic ZnO NPs is not exempt from the selectivity of conventionally synthesized ZnO NPs with regard to its use in Gram-negative bacteria. This is because Gram-positive bacteria are more resistant due to their thick peptidoglycan layer around the cell walls. Moreover, the electrostatic interactions that exist between the positively charged Zn2+ and the negatively charged surfaces of the Gram-negative bacteria promote the penetration of Zn2+ NPs into Gram-negative bacteria [169] (Figure 13c). Sharma et al. [73] prepared spherical ZnO NPs that inhibited the growth of B. subtilus (21.3 mm), E coli (18.6 mm), S. typhi (20.6 mm), Klebsiella pneumoniae (21.6 mm), and S. aureus. Rajendran et al. [100] recorded a minimum inhibitory concentration (MIC) of 157.2 µg/mL against S. aureus when the bacterial concentration was 1 × 105 CFU/mL and the root extract of Rubus fairholmianus synthesized with ZnO NPs (100 µL) was employed. The production of H 2 O 2 that disturbed the lipid and protein bilayers was largely responsible for the inactivation of the bacterial cells [70,229,230] (Figure 13b). An extract from Phoenix roebelenii leaves was used to chelate ZnO NPs during their synthesis, the inhibition zones for which were 21.4 mm for S. typhi, 15.8 mm for E. coli, 15.1 mm for S. pneumoniae, and 15 mm for S. aureus. The prepared samples exhibited a greater antibacterial performance against Gram-negative bacteria than against Gram-positive bacteria. The extract of Prunus dulcis was employed by Anand and colleagues for the synthesis of ZnO NPs. The high-resolution SEM (HRSEM) showed aggregated, but homogeneous, ZnO NPs with a mean size of 25 nm. The antibacterial activity of the produced nanoparticles was evaluated using the agar diffusion technique. The green prepared ZnO NPs showed bactericidal effects on S. aureus, E. coli, and S. paratyphi, with zone diameters of 18 mm, 32 mm, and 25 mm, respectively. Interestingly, there was no zone of inhibition (ZOI) observed on the Gram-negative bacteria, Klebsiella pneumoniae, and P. mirabilius [231]. Table 7 summarizes recent findings on the plant-mediated production of ZnO NPs which aimed to inactivate both Gram-positive and Gram-negative bacteria.
Bacteria have also been inactivated by algae-synthesized ZnO NPs [232,233]. When Debnath et al. [218] prepared spherical Hypnea musciformis-capped ZnO NPs, the growth of S. aureus and E. coli was inhibited via ROS generation on the surface of the NPs and the interactions between zinc and the cell membranes of bacteria through the adhesion of ZnO NPs. Jamil et al. [233] inhibited the growth of E. coli (20 mm), S. aureus (14.5 mm), Klebsiella pneumoniae (11 mm), C. freundii (20 mm), and M. luteus (12 mm) by capping the NPs with Phormidium sp., Viridiplantae sp., and Cosmarium sp. When an aqueous extract from Sargassum muticum was used as a bioreducing reagent for the preparation of ZnO NPs, the obtained NPs showed enhanced antibacterial activity against Gram-negative, drug-resistant Pseudomonas stutzeri and Acinetobacter baumannii, when compared with the antibacterial activity against Gram-positive Bacillus filamentosus and Bacillus flexus. This is because the cell walls of the Gram-positive bacteria are complex and they have a very thick peptidoglycan that restricts the NPs from entering the bacterial cells [234].
The fungal-mediated synthesis of ZnO NPs that are active against bacteria has been reported, and the findings are summarized in Table 7. Aspergillus fumigatus synthesized ZnO NPs are used for eradicating Klebsiella pneumoniae (30 mm), P. aeruginosa (27 mm), E. coli (25 mm), S. aureus (30 mm), and B. subtilis (20 mm) when following a well diffusion method [126]. The bactericidal activity of Pichia kudriavzevii-capped ZnO NPs on both Gram-positive bacteria (B. subtilis, S. epidermidis and S. aurous) and Gram-negative bacteria (E. coli and Serratia marcescens) was reported. A zone of inhibition (~19 mm) was observed for S. epidermidis at a concentration of 100 µg/mL (ZnO NPs) for a sample prepared at RT for over 4 h [130]. Spherical ZnO NPs, stabilized with Saccharomyces cerevisiae, were reported to kill E. coli [196]. Results on the bactericidal activity of biological derivatives that mediated ZnO have been amply reported [151,157,171] and are summarized in Table 7.

4.3.2. Antifungal Activity of ZnO NPs

Phyto-meditated ZnO NPs that exhibited antifungal activity were reported by Madhumitha, who prepared ZnO NPs from the peels of Pithecellobium dulce, using a broth dilution procedure in order to inhibit the growth of A. flavus and A. niger with 63.57% and 57.17% efficiency, respectively, when the NP concentration was 1000 mg/L [83]. Nilavukkarasi et al. [91] used a Capparis zeylanica leaf extract to cap NPs, and they incorporated 100 µg/mL ZnO NPs using a well diffusion method to inactivate C. albicans and A niger. Moreover, they realized that the zone of inhibition was 30 mm and 12 mm for C. albicans and A. niger after 24 h, respectively. The fungi were killed by oxidative stress, which resulted in cell death by apoptosis. Spherical and almond shaped ZnO NPs were attained by using different parts of Bixa orellana. These ZnO NPs demonstrated antifungal activity against Penicillium sp., Aspergillus flavus, Fusarium oxysporum, and Rhizoctonia solani [235]. Gelatin-coated ZnO NPs were synthesized via co-precipitation, which effectively inhibited the biofilm growth of the fungus Candida albicans at 50 μg/mL [62]. ZnO NPs were produced on the surface of bacterial cellulose (BC) films with the aid of an ultrasound. These were later impregnated with the extracts of ethanolic propolis (EEP). The BC-ZnO-100US film exhibited an MIC value of 1.3 mg/mL against C. albicans. Propolis’ antibacterial efficacy was shown to be reliant on the presence of phenolic chemicals, such as flavonoids, which allow the nanomaterial to break down fungal cell membranes, thus resulting in a loss of K+ and subsequent cell autolysis [236].

5. Conclusions

The attainment of ZnO NPs via the green protocol has been a focal point for research in recent times. Green sources, such as plants, bacteria, fungi, and algae, operate as both reducing and stabilizing agents to regulate the morphologies of the NPs. This review has provided insights on green synthesis and its advantages over traditional synthesis techniques. Recent advances in the biosynthesis of ZnO NPs, which uses plants, bacteria, fungus, and algae, show the ability of biosynthesis protocols to produce nanoparticles in the same size range as ZnO NPs that are synthesized using conventional chemical synthesis procedures. The morphology and particle size are shown to depend on factors such as reaction time, pH, temperature, and the biogenic precursor used. The key characterization techniques investigate the structural and chemical properties of the biosynthesized ZnO NPs and most importantly, the crystallinity. Furthermore, the use of the produced ZnO NPs for water treatment and disinfection were also explored. In retrospect, green, synthetic approaches for ZnO NPs have, so far, managed to satisfactorily meet five of the twelve pillars of green synthesis. These pillars include the use of readily available precursors; the avoidance of the use of derivatives such as stabilizers; the employment of ecofriendly solvents; and the use of harmless chemicals to evade toxicity. In place of chemical stabilizers and solvents, it is evident that natural metabolites (plant, bacterial, and fungal) are able to satisfactorily be used as stabilizers, capping agents, and reducing agents. Indeed, the particle sizes for the ZnO NPs that are manufactured using green, synthetic approaches are comparable to the ZnO NPs that are synthesized using conventional approaches. There is still a great deal of research that needs to be undertaken in order to create greener ZnO green synthesis approaches with respect to the atom economy and reducing energy expenditure during synthesis. A great deal of progress has been made with regard to the reduction of energy requirements when formulating microbe-mediated, greener approaches, as shown by the elimination of post-synthesis calcination, albeit with less crystalline ZnO; however, this progress is unsatisfactory.

6. Future Perspectives

  • The likelihood of integrating cheap, readily available biological sources with one another, and gaining control of the parameters concerning environmentally benign syntheses, justifies a future focus on the sustainable and scalable production of ZnO NPs; however, to completely maximize and make use of the advantages that biogenic ZnONPs provide, several areas need more research to enable a better understanding of the topic.
  • As there are few studies on ZnO nanomaterial synthesis using microorganisms, biopolymers, marine fungi, algae, and waste materials, there exists a need for further exploration into biogenic synthesis protocols, largely because it offers more advantages over conventional chemical and physical synthesis.
  • It is difficult to use ZnO NPs for biosynthesis on an industrial level as the underlying reaction mechanisms and kinetics of the process are not yet fully understood, especially microbial synthesis processes. Currently, reproducing bio-prepared NPs is difficult and may be time consuming because the mechanisms are insufficiently understood. The development of predictable synthesis protocols thus remains a challenge.
  • A thorough understanding of the manner in which the most active of the bioprecursors that cap and reduce the nanomaterials are isolated and traced, in addition to an understanding of their structure, is necessary. Understanding the structure and chemistry of the reducing agents will speed up the commercialization of the green synthesis protocol. Moreover, the green protocol method must be explored in greater depth in order to understand the effects of the various biological extracts on the surface of ZnO, and to understand the deep level defects that occur, especially with respect to the photocatalytic break down of pollutants. Understanding this critical information may assist in eliminating the need to anneal the resultant samples, which consumes a great deal of energy.
  • The manner in which the yield of the ZnO NPs corresponds with the precursors has not been reported, and process cost comparisons between biologically synthesized NPs and traditionally produced NPs have not been reported on as the former has not yet been upscaled. Understanding how the yield corresponds with the precursors should be a priority in order to obtain realistic expectations regarding this synthesis protocol. This requires a full understanding of the life cycle of specific bioprecursors in order to ensure continuous availability. This should prevent the production of the material from stopping or from being significantly reduced once the commercialization of green synthesis has been fully adopted. Such studies would also enable biosynthesis protocols to improve the atom economy in synthetic processes, thus meeting another pillar of green synthesis.
The eventual release of a high concentration of the green synthesized ZnO NPs into bodies of water would likely be toxic; therefore, it is imperative that these NPs are embedded onto solid substrates that can easily be separated from bodies of water. Furthermore, the toxicity of the NPs should not only be studied in vitro, but also in vivo, in order to obtain a comprehensible idea of the material’s toxicity on living things. Lastly, ZnO, being an amphoteric oxide, is soluble in water at extreme pH levels, thus the Zn2+ ion can be placed in water. There is a need to investigate the influence of biosynthetic approaches on the stabilization of ZnO NPs with a range of pH levels.

Author Contributions

All the listed authors contributed to the writing of the review manuscript. N.E.Z.—wrote prepared the original draft, reviewed, and edited the manuscript as advised; S.A.Z.—contributed on the manuscript drafting, reviewed, and edited the whole document; S.A.—Supervised the work, sourced the funding, I.K.—Read, reviewed, and edited the manuscript; A.A.Z.—Reviewed and edited the document; M.M.—Supervised, provided funding, and reviewed the document; K.K.—Visualization and Validation, and edited the manuscript, T.M.—Visualization and edited the manuscript. All authors have read and agreed to the published version of the manuscript.

Funding

This work is supported by the SASOL-NRF University Collaborative Research Grant (No. 138626).

Data Availability Statement

Not applicable.

Acknowledgments

The authors are grateful to SASOL-NRF University Collaborative Research Grants (Grant No. 138626) and to the Unesco-Unisa Africa Chair for Nanoscience and nanotechnology for their support.

Conflicts of Interest

The authors declare no conflict of interests.

References

  1. Herrmann, S.; De Matteis, L.; de la Fuente, J.M.; Mitchell, S.G.; Streb, C. Removal of Multiple Contaminants from Water by Polyoxometalate Supported Ionic Liquid Phases (POM-SILPs). Angew. Chem.-Int. Ed. 2017, 56, 1667–1670. [Google Scholar] [CrossRef]
  2. Sun, H.; Cao, L.; Lu, L. Magnetite/Reduced Graphene Oxide Nanocomposites: One Step Solvothermal Synthesis and Use as a Novel Platform for Removal of Dye Pollutants. Nano Res. 2011, 4, 550–562. [Google Scholar] [CrossRef]
  3. Mok, C.F.; Ching, Y.C.; Muhamad, F.; Abu Osman, N.A.; Hai, N.D.; Che Hassan, C.R. Adsorption of Dyes Using Poly(Vinyl Alcohol) (PVA) and PVA-Based Polymer Composite Adsorbents: A Review. J. Polym. Environ. 2020, 28, 775–793. [Google Scholar] [CrossRef]
  4. Dhiman, V.; Kondal, N. ZnO Nanoadsorbents: A Potent Material for Removal of Heavy Metal Ions from Wastewater. Colloids Interface Sci. Commun. 2021, 41, 100380. [Google Scholar] [CrossRef]
  5. Ashbolt, N.J. Microbial Contamination of Drinking Water and Human Health from Community Water Systems. Curr. Environ. Heal. Rep. 2015, 2, 95–106. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  6. Roberts, K.; Alberts, B.; Johnson, A.; Lewis, J.; Raff, M.; Walter, P.; Edition, F.; Roberts, K.; Alberts, B.; Johnson, A.; et al. Molecular Biology of the Cell, Fourth Edition Description, 6th ed.; W.W. Norton & Company: Boca Raton, FL, USA, 2017; ISBN 9781315735368. [Google Scholar]
  7. Prasad, A.R.; Williams, L.; Garvasis, J.; Shamsheera, K.O.; Basheer, S.M.; Kuruvilla, M.; Joseph, A. Applications of Phytogenic ZnO Nanoparticles: A Review on Recent Advancements. J. Mol. Liq. 2021, 331, 115805. [Google Scholar] [CrossRef]
  8. Oluwafemi, O.S.; Anyik, J.L.; Zikalala, N.E.; Sakho, E.H.M. Biosynthesis of Silver Nanoparticles from Water Hyacinth Plant Leaves Extract for Colourimetric Sensing of Heavy Metals. Nano-Struct. Nano-Objects 2019, 20, 7–9. [Google Scholar] [CrossRef]
  9. Singh, A.K.; Srivastava, O.N. One-Step Green Synthesis of Gold Nanoparticles Using Black Cardamom and Effect of pH on Its Synthesis. Nanoscale Res. Lett. 2015, 10, 1–12. [Google Scholar] [CrossRef] [Green Version]
  10. Akintelu, S.A.; Folorunso, A.S.; Folorunso, F.A.; Oyebamiji, A.K. Green Synthesis of Copper Oxide Nanoparticles for Biomedical Application and Environmental Remediation. Heliyon 2020, 6, e04508. [Google Scholar] [CrossRef]
  11. Zikalala, S.A.; Kuvarega, A.T.; Mamba, B.B.; Mhlanga, S.D.; Nxumalo, E.N. The Effect of Synthetic Routes on the Physicochemical Properties and Optical Response of N-Doped Titania–Oxidized Carbon Nanotube Nanohybrids. Mater. Today Chem. 2018, 10, 1–18. [Google Scholar] [CrossRef]
  12. Ling, D.; Lee, N.; Hyeon, T. Chemical Synthesis and Assembly of Uniformly Sized Iron Oxide Nanoparticles for Medical Applications. Acc. Chem. Res. 2015, 48, 1276–1285. [Google Scholar] [CrossRef] [PubMed]
  13. Kuriganova, A.B.; Vlaic, C.A.; Ivanov, S.; Leontyeva, D.V.; Bund, A.; Smirnova, N.V. Electrochemical Dispersion Method for the Synthesis of SnO2 as Anode Material for Lithium Ion Batteries. J. Appl. Electrochem. 2016, 46, 527–538. [Google Scholar] [CrossRef]
  14. Sobhani-Nasab, A.; Behpour, M. Synthesis and Characterization of AgO Nanostructures by Precipitation Method and Its Photocatalyst Application. J. Mater. Sci. Mater. Electron. 2016, 27, 1191–1196. [Google Scholar] [CrossRef]
  15. Siddiqui, H.; Qureshi, M.S.; Haque, F.Z. Surfactant Assisted Wet Chemical Synthesis of Copper Oxide (CuO) Nanostructures and Their Spectroscopic Analysis. Optik 2016, 127, 2740–2747. [Google Scholar] [CrossRef]
  16. Sahu, K.; Choudhary, S.; Singh, J.; Kuriakose, S.; Singhal, R.; Mohapatra, S. Facile Wet Chemical Synthesis of ZnO Nanosheets: Effects of Counter Ions on the Morphological, Structural, Optical and Photocatalytic Properties. Ceram. Int. 2018, 44, 23094–23101. [Google Scholar] [CrossRef]
  17. Champi, A.; Marques, F.C. Structural Changes in Amorphous Carbon Nitride Films Due to Bias Voltage. Thin Solid Film. 2006, 501, 362–365. [Google Scholar] [CrossRef]
  18. Aminuzzaman, M.; Ying, L.P.; Goh, W.S.; Watanabe, A. Green Synthesis of Zinc Oxide Nanoparticles Using Aqueous Extract of Garcinia mangostana Fruit Pericarp and Their Photocatalytic Activity. Bull. Mater. Sci. 2018, 41, 1–10. [Google Scholar] [CrossRef] [Green Version]
  19. Janotti, A.; Van De Walle, C.G. Fundamentals of Zinc Oxide as a Semiconductor. Rep. Prog. Phys. 2009, 72, 12–15. [Google Scholar] [CrossRef] [Green Version]
  20. Xu, S.; Weintraub, B.; Wang, Z.L. Zinc Oxide Nanowire Arrays on Flexible Substrates. Wet Chemical Growth and Applications in Energy Conversion, 1st ed.; Elsevier Ltd.: Amsterdam, The Netherlands, 2010; ISBN 9781437778236. [Google Scholar]
  21. Ayoub, I.; Kumar, V.; Abolhassani, R.; Sehgal, R.; Sharma, V.; Sehgal, R.; Swart, H.C.; Mishra, Y.K. Advances in ZnO: Manipulation of Defects for Enhancing Their Technological Potentials. Nanotechnol. Rev. 2022, 11, 575–619. [Google Scholar] [CrossRef]
  22. Adeyemi, J.O.; Elemike, E.E.; Onwudiwe, D.C. ZnO Nanoparticles Mediated by Aqueous Extracts of Dovyalis Caffra Fruits and the Photocatalytic Evaluations. Mater. Res. Express 2019, 6, 11–14. [Google Scholar] [CrossRef]
  23. Narendra Kumar, H.K.; Chandra Mohana, N.; Nuthan, B.R.; Ramesha, K.P.; Rakshith, D.; Geetha, N.; Satish, S. Phyto-Mediated Synthesis of Zinc Oxide Nanoparticles Using Aqueous Plant Extract of Ocimum americanum and Evaluation of Its Bioactivity. SN Appl. Sci. 2019, 1, 1–9. [Google Scholar] [CrossRef] [Green Version]
  24. Awwad, A.M.; Albiss, B.; Ahmad, A.L. Green Synthesis, Characterization and Optical Properties of Zinc Oxide Nanosheets Using Olea europea Leaf Extract. Adv. Mater. Lett. 2014, 5, 520–524. [Google Scholar] [CrossRef]
  25. Shabaani, M.; Rahaiee, S.; Zare, M.; Jafari, S.M. Green Synthesis of ZnO Nanoparticles Using Loquat Seed Extract; Biological Functions and Photocatalytic Degradation Properties. LWT 2020, 134, 110133. [Google Scholar] [CrossRef]
  26. Rajabairavi, N.; Raju, C.S.; Karthikeyan, C.; Varutharaju, K.; Nethaji, S.; Hameed, A.S.H.; Shajahan, A. Biosynthesis of Novel Zinc Oxide Nanoparticles (ZnO NPs) Using Endophytic Bacteria Sphingobacterium thalpophilum. Springer Proc. Phys. 2017, 189, 245–254. [Google Scholar] [CrossRef]
  27. Wu, J.; Yin, C.; Zhou, J.; Li, H.; Liu, Y.; Shen, Y.; Garner, S.; Fu, Y.; Duan, H. Ultrathin Glass-Based Flexible, Transparent, and Ultrasensitive Surface Acoustic Wave Humidity Sensor with ZnO Nanowires and Graphene Quantum Dots. ACS Appl. Mater. Interfaces 2020, 12, 39817–39825. [Google Scholar] [CrossRef]
  28. Timerkaev, B.A.; Felsinger, V.S.; Kaleeva, A.A.; Erlingayte, E.A.; Uktamov, J.A.; Nuriddinov, H.S. Plasma-Chemical Synthesis of Zinc Oxide Nanotubes. J. Phys. Conf. Ser. 2021, 1870, 012003. [Google Scholar] [CrossRef]
  29. Ebadi, M.; Zolfaghari, M.R.; Aghaei, S.S.; Zargar, M.; Shafiei, M.; Zahiri, H.S.; Noghabi, K.A. A Bio-Inspired Strategy for the Synthesis of Zinc Oxide Nanoparticles (ZnO NPs) Using the Cell Extract of Cyanobacterium: Nostoc Sp. EA03: From Biological Function to Toxicity Evaluation. RSC Adv. 2019, 9, 23508–23525. [Google Scholar] [CrossRef] [Green Version]
  30. Mahdi, Z.S.; Talebnia Roshan, F.; Nikzad, M.; Ezoji, H. Biosynthesis of Zinc Oxide Nanoparticles Using Bacteria: A Study on the Characterization and Application for Electrochemical Determination of Bisphenol A. Inorg. Nano-Met. Chem. 2021, 51, 1249–1257. [Google Scholar] [CrossRef]
  31. Kiani, A.; Dastafkan, K. Zinc Oxide Nanocubes as a Destructive Nanoadsorbent for the Neutralization Chemistry of 2-Chloroethyl Phenyl Sulfide: A Sulfur Mustard Simulant. J. Colloid Interface Sci. 2016, 478, 271–279. [Google Scholar] [CrossRef]
  32. Mustapha, S.; Ndamitso, M.M.; Abdulkareem, A.S.; Tijani, J.O.; Shuaib, D.T.; Ajala, A.O.; Mohammed, A.K. Application of TiO2 and ZnO Nanoparticles Immobilized on Clay in Wastewater Treatment: A Review. Appl. Water Sci. 2020, 10, 1–36. [Google Scholar] [CrossRef]
  33. Saeed, M.; Siddique, M.; Ibrahim, M.; Akram, N.; Usman, M.; Aleem, M.A.; Baig, A. Calotropis Gigantea Leaves Assisted Biosynthesis of ZnO and Ag@ZnO Catalysts for Degradation of Rhodamine B Dye in Aqueous Medium. Environ. Prog. Sustain. Energy 2020, 39, e13408. [Google Scholar] [CrossRef]
  34. Abdelhakim, H.K.; El-Sayed, E.R.; Rashidi, F.B. Biosynthesis of Zinc Oxide Nanoparticles with Antimicrobial, Anticancer, Antioxidant and Photocatalytic Activities by the Endophytic Alternaria tenuissima. J. Appl. Microbiol. 2020, 128, 1634–1646. [Google Scholar] [CrossRef] [PubMed]
  35. Liu, Y.; Liu, H.; Zhang, Q.; Li, T. Adjusting the Proportions of {0001} Facets and High-Index Facets of ZnO Hexagonal Prisms and Their Photocatalytic Activity. RSC Adv. 2017, 7, 3515–3520. [Google Scholar] [CrossRef] [Green Version]
  36. Zhang, J.; Zhao, B.; Pan, Z.; Gu, M.; Punnoose, A. Synthesis of ZnO Nanoparticles with Controlled Shapes, Sizes, Aggregations, and Surface Complex Compounds for Tuning or Switching the Photoluminescence. Cryst. Growth Des. 2015, 15, 3144–3149. [Google Scholar] [CrossRef]
  37. Tănase, M.A.; Marinescu, M.; Oancea, P.; Răducan, A.; Mihaescu, C.I.; Alexandrescu, E.; Nistor, C.L.; Jinga, L.I.; Diţu, L.M.; Petcu, C.; et al. Antibacterial and Photocatalytic Properties of ZnO Nanoparticles Obtained from Chemical versus Saponaria Officinalis Extract-Mediated Synthesis. Molecules 2021, 26, 2072. [Google Scholar] [CrossRef] [PubMed]
  38. Ramirez-Vick, J.E. Nanostructured ZnO for Electrochemical Biosensors. J. Biosens. Bioelectron. 2012, 3, 1. [Google Scholar] [CrossRef] [Green Version]
  39. Rambabu, K.; Bharath, G.; Banat, F.; Show, P.L. Green Synthesis of Zinc Oxide Nanoparticles Using Phoenix dactylifera Waste as Bioreductant for Effective Dye Degradation and Antibacterial Performance in Wastewater Treatment. J. Hazard. Mater. 2021, 402, 123560. [Google Scholar] [CrossRef] [PubMed]
  40. Sharma, S.; Kumar, K.; Thakur, N.; Chauhan, M.S. Ocimum tenuiflorum Leaf Extract as a Green Mediator for the Synthesis of ZnO Nanocapsules Inactivating Bacterial Pathogens. Chem. Pap. 2020, 74, 3431–3444. [Google Scholar] [CrossRef]
  41. Fazlzadeh, M.; Khosravi, R.; Zarei, A. Green Synthesis of Zinc Oxide Nanoparticles Using Peganum harmala Seed Extract, and Loaded on Peganum harmala Seed Powdered Activated Carbon as New Adsorbent for Removal of Cr(VI) from Aqueous Solution. Ecol. Eng. 2017, 103, 180–190. [Google Scholar] [CrossRef]
  42. Bao, S.; Hou, T.; Tan, Q.; Kong, X.; Cao, H.; He, M.; Wu, G.; Xu, B. Immobilization of Zinc Oxide Nanoparticles on Graphene Sheets for Lithium Ion Storage and Electromagnetic Microwave Absorption. Mater. Chem. Phys. 2020, 245, 1–3. [Google Scholar] [CrossRef]
  43. Ren, Q.; Cao, Y.-Q.; Arulraj, D.; Liu, C.; Wu, D.; Li, W.-M.; Li, A.-D. Review—Resistive-Type Hydrogen Sensors Based on Zinc Oxide Nanostructures. J. Electrochem. Soc. 2020, 167, 067528. [Google Scholar] [CrossRef]
  44. Kim, K.B.; Kim, Y.W.; Lim, S.K.; Roh, T.H.; Bang, D.Y.; Choi, S.M.; Lim, D.S.; Kim, Y.J.; Baek, S.H.; Kim, M.K.; et al. Risk Assessment of Zinc Oxide, a Cosmetic Ingredient Used as a UV Filter of Sunscreens. J. Toxicol. Environ. Heal.-Part B Crit. Rev. 2017, 20, 155–182. [Google Scholar] [CrossRef] [PubMed]
  45. Khashan, K.S.; Sulaiman, G.M.; Hussain, S.A.; Marzoog, T.R.; Jabir, M.S. Synthesis, Characterization and Evaluation of Anti-Bacterial, Anti-Parasitic and Anti-Cancer Activities of Aluminum-Doped Zinc Oxide Nanoparticles. J. Inorg. Organomet. Polym. Mater. 2020, 30, 3677–3693. [Google Scholar] [CrossRef]
  46. Nakano, M.; Nakagawa, S.; Sato, F.; Shahiduzzaman, M.; Karakawa, M.; Taima, T.; Takahashi, K. Study on Properties of Low-Temperature-Prepared Zinc Oxide-Based Inverted Organic Solar Cells and Improvement of Their Photodurability. ACS Appl. Energy Mater. 2021, 4, 6385–6390. [Google Scholar] [CrossRef]
  47. Rafique, M.; Tahir, R.; Gillani, S.S.A.; Tahir, M.B.; Shakil, M.; Iqbal, T.; Abdellahi, M.O. Plant-Mediated Green Synthesis of Zinc Oxide Nanoparticles from Syzygium Cumini for Seed Germination and Wastewater Purification. Int. J. Environ. Anal. Chem. 2020, 102, 23–28. [Google Scholar] [CrossRef]
  48. Bandeira, M.; Chee, B.S.; Frassini, R.; Nugent, M.; Giovanela, M.; Roeschely, M.; Crespo, J. da S.; Devine, D.M. Antimicrobial Paa/Pah Electrospun Fiber Containing Green Synthesized Zinc Oxide Nanoparticles for Wound Healing. Matreials 2021, 14, 2889. [Google Scholar] [CrossRef]
  49. Sirelkhatim, A.; Mahmud, S.; Seeni, A.; Kaus, N.H.M.; Ann, L.C.; Bakhori, S.K.M.; Hasan, H.; Mohamad, D. Review on Zinc Oxide Nanoparticles: Antibacterial Activity and Toxicity Mechanism. Nano-Micro Lett. 2015, 7, 219–242. [Google Scholar] [CrossRef] [Green Version]
  50. 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]
  51. Srujana, S.; Bhagat, D. Chemical—Based Synthesis of ZnO Nanoparticles and Their Applications in Agriculture. Nanotechnol. Environ. Eng. 2022, 7, 269–275. [Google Scholar] [CrossRef]
  52. Naveed Ul Haq, A.; Nadhman, A.; Ullah, I.; Mustafa, G.; Yasinzai, M.; Khan, I. Synthesis Approaches of Zinc Oxide Nanoparticles: The Dilemma of Ecotoxicity. J. Nanomater. 2017, 1–14. [Google Scholar] [CrossRef]
  53. Jose Varghese, R.; Zikalala, N.; Sakho, E.H.M.; Oluwafemi, O.S. Green Synthesis Protocol on Metal Oxide Nanoparticles Using Plant Extracts. In Colloidal Metal Oxide Nanoparticles-Synthesis, Characterization and Applications; Thomas, S., Sani, A.T., Velayudhan, P., Eds.; Elsevier: Amsterdam, The Netherlands, 2020; pp. 67–82. ISBN 978-0-12-813357-6. [Google Scholar]
  54. Jessop, P.G.; Trakhtenberg, S.; Warner, J. The Twelve Principles of Green Chemistry. ACS Symp. Ser. 2009, 1000, 401–436. [Google Scholar] [CrossRef]
  55. Rubab, L.; Anum, A.; Al-hussain, S.A.; Irfan, A.; Ahmad, S.; Ullah, S.; Al-mutairi, A.A.; Zaki, M.E.A. Green Chemistry in Organic Synthesis: Recent Update on Green Catalytic Approaches in Synthesis of 1, 2, 4-Thiadiazoles. Catalysts 2022, 12, 1329. [Google Scholar] [CrossRef]
  56. Mourdikoudis, S.; Liz-Marzán, L.M. Oleylamine in Nanoparticle Synthesis. Chem. Mater. 2013, 25, 1465–1476. [Google Scholar] [CrossRef]
  57. Bandeira, M.; Giovanela, M.; Roesch-Ely, M.; Devine, D.M.; da Silva Crespo, J. Green Synthesis of Zinc Oxide Nanoparticles: A Review of the Synthesis Methodology and Mechanism of Formation. Sustain. Chem. Pharm. 2020, 15, 100223. [Google Scholar] [CrossRef]
  58. Swati; Verma, R.; Chauhan, A.; Shandilya, M.; Li, X.; Kumar, R.; Kulshrestha, S. Antimicrobial Potential of Ag-Doped ZnO Nanostructure Synthesized by the Green Method Using Moringa oleifera Extract. J. Environ. Chem. Eng. 2020, 8, 103730. [Google Scholar] [CrossRef]
  59. Jayaseelan, C.; Rahuman, A.A.; Kirthi, A.V.; Marimuthu, S.; Santhoshkumar, T.; Bagavan, A.; Gaurav, K.; Karthik, L.; Rao, K.V.B. Novel Microbial Route to Synthesize ZnO Nanoparticles Using Aeromonas hydrophila and Their Activity against Pathogenic Bacteria and Fungi. Spectrochim. Acta-Part A Mol. Biomol. Spectrosc. 2012, 90, 78–84. [Google Scholar] [CrossRef] [PubMed]
  60. Jain, N.; Bhargava, A.; Tarafdar, J.C.; Singh, S.K.; Panwar, J. A Biomimetic Approach towards Synthesis of Zinc Oxide Nanoparticles. Appl. Microbiol. Biotechnol. 2013, 97, 859–869. [Google Scholar] [CrossRef]
  61. Rao, M.D.; Gautam, D.P. Nanoflowers Using Chlamydomonas reinhardtii: A Green Approach. Environ. Prog. Sustain. Energy 2016, 35, 1020–1026. [Google Scholar] [CrossRef]
  62. Divya, M.; Vaseeharan, B.; Abinaya, M.; Vijayakumar, S.; Govindarajan, M.; Alharbi, N.S.; Kadaikunnan, S.; Khaled, J.M.; Benelli, G. Biopolymer Gelatin-Coated Zinc Oxide Nanoparticles Showed High Antibacterial, Antibiofilm and Anti-Angiogenic Activity. J. Photochem. Photobiol. B Biol. 2018, 178, 211–218. [Google Scholar] [CrossRef] [PubMed]
  63. Zikalala, N.; Matshetshe, K.; Parani, S.; Oluwafemi, O.S. Biosynthesis Protocols for Colloidal Metal Oxide Nanoparticles. Nano-Struct. Nano-Objects 2018, 16, 288–299. [Google Scholar] [CrossRef]
  64. Vijayaraghavan, K.; Ashokkumar, T. Plant-Mediated Biosynthesis of Metallic Nanoparticles: A Review of Literature, Factors Affecting Synthesis, Characterization Techniques and Applications. J. Environ. Chem. Eng. 2017, 5, 4866–4883. [Google Scholar] [CrossRef]
  65. Ahmed, S.; Annu; Chaudhry, S.A.; Ikram, S. A Review on Biogenic Synthesis of ZnO Nanoparticles Using Plant Extracts and Microbes: A Prospect towards Green Chemistry. J. Photochem. Photobiol. B Biol. 2017, 166, 272–284. [Google Scholar] [CrossRef]
  66. Akbar, S.; Tauseef, I.; Subhan, F.; Sultana, N.; Khan, I.; Ahmed, U.; Haleem, K.S. An Overview of the Plant-Mediated Synthesis of Zinc Oxide Nanoparticles and Their Antimicrobial Potential. Inorg. Nano-Met. Chem. 2020, 50, 257–271. [Google Scholar] [CrossRef]
  67. Prasad, A.R.; Anagha, M.; Shamsheera, K.O.; Joseph, A. Bio-Fabricated ZnO Nanoparticles: Direct Sunlight-Driven Selective Photodegradation, Antibacterial Activity, and Thermoluminescence-Emission Characteristics. New J. Chem. 2020, 44, 8273–8279. [Google Scholar] [CrossRef]
  68. Verma, P.R.; Khan, F.; Banerjee, S. Salvadora persica Root Extract-Mediated Fabrication of ZnO Nanoparticles and Characterization. Inorg. Nano-Met. Chem. 2020, 51, 427–433. [Google Scholar] [CrossRef]
  69. Fagier, M.A. Plant-Mediated Biosynthesis and Photocatalysis Activities of Zinc Oxide Nanoparticles: A Prospect towards Dyes Mineralization. J. Nanotechnol. 2021, 2021. [Google Scholar] [CrossRef]
  70. Elumalai, K.; Velmurugan, S.; Ravi, S.; Kathiravan, V.; Adaikala Raj, G. Bio-Approach: Plant Mediated Synthesis of ZnO Nanoparticles and Their Catalytic Reduction of Methylene Blue and Antimicrobial Activity. Adv. Powder Technol. 2015, 26, 1639–1651. [Google Scholar] [CrossRef]
  71. Gebre, S.H.; Sendeku, M.G. New Frontiers in the Biosynthesis of Metal Oxide Nanoparticles and Their Environmental Applications: An Overview. SN Appl. Sci. 2019, 1, 1–28. [Google Scholar] [CrossRef] [Green Version]
  72. Nasrollahzadeh, M.; Sajjadi, M.; Iravani, S.; Varma, R.S. Green-Synthesized Nanocatalysts and Nanomaterials for Water Treatment: Current Challenges and Future Perspectives. J. Hazard. Mater. 2021, 401, 123401. [Google Scholar] [CrossRef]
  73. Sharma, A.; Nagraik, R.; Sharma, S.; Sharma, G.; Pandey, S.; Azizov, S.; Chauhan, P.K.; Kumar, D. Green Synthesis of ZnO Nanoparticles Using Ficus palmata: Antioxidant, Antibacterial and Antidiabetic Studies. Results Chem. 2022, 4, 100509. [Google Scholar] [CrossRef]
  74. Bala, N.; Saha, S.; Chakraborty, M.; Maiti, M.; Das, S.; Basu, R.; Nandy, P. Green Synthesis of Zinc Oxide Nanoparticles Using Hibiscus subdariffa Leaf Extract: Effect of Temperature on Synthesis, Anti-Bacterial Activity and Anti-Diabetic Activity. RSC Adv. 2015, 5, 4993–5003. [Google Scholar] [CrossRef]
  75. Senthilkumar, N.; Nandhakumar, E.; Priya, P.; Soni, D.; Vimalan, M.; Vetha Potheher, I. Synthesis of ZnO Nanoparticles Using Leaf Extract of Tectona grandis (L.) and Their Anti-Bacterial, Anti-Arthritic, Anti-Oxidant and in Vitro Cytotoxicity Activities. New J. Chem. 2017, 41, 10347–10356. [Google Scholar]
  76. Bowles, D.; Lim, E.K.; Poppenberger, B.; Vaistij, F.E. Glycosyltransferases of Lipophilic Small Molecules. Annu. Rev. Plant Biol. 2006, 57, 567–597. [Google Scholar] [CrossRef] [PubMed]
  77. Król, A.; Railean-Plugaru, V.; Pomastowski, P.; Buszewski, B. Phytochemical Investigation of Medicago sativa L. Extract and Its Potential as a Safe Source for the Synthesis of ZnO Nanoparticles: The Proposed Mechanism of Formation and Antimicrobial Activity. Phytochem. Lett. 2019, 31, 170–180. [Google Scholar] [CrossRef]
  78. Krężel, A.; Maret, W. The Biological Inorganic Chemistry of Zinc Ions. Arch. Biochem. Biophys. 2016, 611, 3–19. [Google Scholar] [CrossRef] [Green Version]
  79. Kumar, B.; Smita, K.; Cumbal, L.; Debut, A. Green Approach for Fabrication and Applications of Zinc Oxide Nanoparticles. Bioinorg. Chem. Appl. 2014, 2014. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  80. 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]
  81. De Souza, R.F.V.; Giovani, W.F. De Synthesis, Spectral and Electrochemical Properties of Al (III) and Zn (II) Complexes with Flavonoids. Spectrochim. Acta Part A Mol. Biomol. Spectrosc. 2005, 61, 1985–1990. [Google Scholar] [CrossRef]
  82. Mirgane, N.A.; Shivankar, V.S.; Kotwal, S.B.; Wadhawa, G.C.; Sonawale, M.C. Waste Pericarp of Ananas comosus in Green Synthesis Zinc Oxide Nanoparticles and Their Application in Waste Water Treatment. Mater. Today Proc. 2020, 37, 886–889. [Google Scholar] [CrossRef]
  83. Madhumitha, G.; Fowsiya, J.; Gupta, N.; Kumar, A.; Singh, M. Green Synthesis, Characterization and Antifungal and Photocatalytic Activity of Pithecellobium dulce Peel–Mediated ZnO Nanoparticles. J. Phys. Chem. Solids 2019, 127, 43–51. [Google Scholar] [CrossRef]
  84. Mohamad Sukri, S.N.A.; Shameli, K.; Mei-Theng Wong, M.; Teow, S.Y.; Chew, J.; Ismail, N.A. Cytotoxicity and Antibacterial Activities of Plant-Mediated Synthesized Zinc Oxide (ZnO) Nanoparticles Using Punica granatum (Pomegranate) Fruit Peels Extract. J. Mol. Struct. 2019, 1189, 57–65. [Google Scholar] [CrossRef]
  85. Golmohammadi, M.; Honarmand, M.; Ghanbari, S. A Green Approach to Synthesis of ZnO Nanoparticles Using Jujube Fruit Extract and Their Application in Photocatalytic Degradation of Organic Dyes. Spectrochim. Acta-Part A Mol. Biomol. Spectrosc. 2020, 229, 117961. [Google Scholar] [CrossRef] [PubMed]
  86. Faisal, S.; Jan, H.; Shah, S.A.; Shah, S.; Khan, A.; Akbar, M.T.; Rizwan, M.; Jan, F.; Wajidullah; Akhtar, N.; et al. Green Synthesis of Zinc Oxide (ZnO) Nanoparticles Using Aqueous Fruit Extracts of Myristica fragrans: Their Characterizations and Biological and Environmental Applications. ACS Omega 2021, 6, 9709–9722. [Google Scholar] [CrossRef] [PubMed]
  87. Soto-Robles, C.A.; Luque, P.A.; Gómez-Gutiérrez, C.M.; Nava, O.; Vilchis-Nestor, A.R.; Lugo-Medina, E.; Ranjithkumar, R.; Castro-Beltrán, A. Study on the Effect of the Concentration of Hibiscus sabdariffa Extract on the Green Synthesis of ZnO Nanoparticles. Results Phys. 2019, 15, 102807. [Google Scholar] [CrossRef]
  88. Khara, G.; Padalia, H.; Moteriya, P.; Chanda, S. Peltophorum pterocarpum Flower-Mediated Synthesis, Characterization, Antimicrobial and Cytotoxic Activities of ZnO Nanoparticles. Arab. J. Sci. Eng. 2018, 43, 3393–3401. [Google Scholar] [CrossRef]
  89. Diallo, A.; Ngom, B.D.; Park, E.; Maaza, M. Green Synthesis of ZnO Nanoparticles by Aspalathus linearis: Structural & Optical Properties. J. Alloys Compd. 2015, 646, 425–430. [Google Scholar] [CrossRef]
  90. Chauhan, A.K.; Kataria, N.; Garg, V.K. Green Fabrication of ZnO Nanoparticles Using Eucalyptus Spp. Leaves Extract and Their Application in Wastewater Remediation. Chemosphere 2020, 247, 125803. [Google Scholar] [CrossRef]
  91. Nilavukkarasi, M.; Vijayakumar, S.; Prathipkumar, S. Capparis Zeylanica Mediated Bio-Synthesized ZnO Nanoparticles as Antimicrobial, Photocatalytic and Anti-Cancer Applications. Mater. Sci. Energy Technol. 2020, 3, 335–343. [Google Scholar] [CrossRef]
  92. Bala Chennaiah, M.; Kumar, K.D.; Kumar, B.S.; Tanneeru, S.R. Characterisation of Zinc Oxide Nanoparticles–Herbal Synthesised Coated with Ocimum tenuiflorum. Adv. Mater. Process. Technol. 2021, 1–12. [Google Scholar] [CrossRef]
  93. Ekennia, A.C.; Uduagwu, D.N.; Nwaji, N.N.; Oje, O.O.; Emma-Uba, C.O.; Mgbii, S.I.; Olowo, O.J.; Nwanji, O.L. Green Synthesis of Biogenic Zinc Oxide Nanoflower as Dual Agent for Photodegradation of an Organic Dye and Tyrosinase Inhibitor. J. Inorg. Organomet. Polym. Mater. 2021, 31, 886–897. [Google Scholar] [CrossRef]
  94. Tabrizi Hafez Moghaddas, S.M.; Elahi, B.; Javanbakht, V. Biosynthesis of Pure Zinc Oxide Nanoparticles Using Quince Seed Mucilage for Photocatalytic Dye Degradation. J. Alloys Compd. 2020, 821, 153519. [Google Scholar] [CrossRef]
  95. Prasad, A.R.; Garvasis, J.; Oruvil, S.K.; Joseph, A. Bio-Inspired Green Synthesis of Zinc Oxide Nanoparticles Using Abelmoschus esculentus Mucilage and Selective Degradation of Cationic Dye Pollutants. J. Phys. Chem. Solids 2019, 127, 265–274. [Google Scholar] [CrossRef]
  96. Kaliraj, L.; Ahn, J.C.; Rupa, E.J.; Abid, S.; Lu, J.; Yang, D.C. Synthesis of Panos Extract Mediated ZnO Nano-Flowers as Photocatalyst for Industrial Dye Degradation by UV Illumination. J. Photochem. Photobiol. B Biol. 2019, 199, 111588. [Google Scholar] [CrossRef] [PubMed]
  97. Matinise, N.; Fuku, X.G.; Kaviyarasu, K.; Mayedwa, N.; Maaza, M. ZnO Nanoparticles via Moringa Oleifera Green Synthesis: Physical Properties & Mechanism of Formation. Appl. Surf. Sci. 2017, 406, 339–347. [Google Scholar] [CrossRef]
  98. Thema, F.T.; Manikandan, E.; Dhlamini, M.S.; Maaza, M. Green Synthesis of ZnO Nanoparticles via Agathosma betulina Natural Extract. Mater. Lett. 2015, 161, 124–127. [Google Scholar] [CrossRef]
  99. Ngom, I.; Ndiayea, N.M.; Bakayoko, A.F.M.; Ngom, B.D.; Maaza, M. On the Use of Moringa oleifera Leaves Extract for the Biosynthesis of NiO and ZnO Nanoparticles. MRS Adv. 2020, 5, 1145–1155. [Google Scholar] [CrossRef]
  100. Rajendran, N.K.; George, B.P.; Houreld, N.N.; Abrahamse, H. Synthesis of Zinc Oxide Nanoparticles Using Rubus Fairholmianus Root Extract and Their Activity against Pathogenic Bacteria. Molecules 2021, 26, 3029. [Google Scholar] [CrossRef]
  101. Zafar, Z.; Yi, S.; Li, J.; Li, C.; Zhu, Y.; Zada, A.; Yao, W.; Liu, Z.; Yue, X. Recent Development in Defects Engineered Photocatalysts: An Overview of the Experimental and Theoretical Strategies. Energy Environ. Mater. 2022, 5, 68–114. [Google Scholar] [CrossRef]
  102. Noman, M.T.; Amor, N.; Petru, M.; Mahmood, A.; Kejzlar, P. Photocatalytic Behaviour of Zinc Oxide Nanostructures on Surface Activation of Polymeric Fibres. Polymers 2021, 13, 1227. [Google Scholar] [CrossRef] [PubMed]
  103. Shelar, S.G.; Mahajan, V.K.; Patil, S.P.; Sonawane, G.H. Effect of Doping Parameters on Photocatalytic Degradation of Methylene Blue Using Ag Doped ZnO Nanocatalyst. SN Appl. Sci. 2020, 2, 1–10. [Google Scholar] [CrossRef] [Green Version]
  104. Srivastava, A.K.; Gupta, V.K.; Lal, B.; Roy, S.; Yadav, S.K.; Gurjar, M.S.; Bag, T.K.; Pandey, N.K.; Singh, B.P. Assessment of the Level of Knowledge and Training Needs of Potato Growing Tribal Farmers of Meghalaya. Int. J. Agric. Environ. Biotechnol. 2012, 5, 483–487. [Google Scholar]
  105. Jain, D.; Shivani; Bhojiya, A.A.; 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]
  106. Panchal, P.; Paul, D.R.; Sharma, A.; Choudhary, P.; Meena, P.; Nehra, S.P. Biogenic Mediated Ag/ZnO Nanocomposites for Photocatalytic and Antibacterial Activities towards Disinfection of Water. J. Colloid Interface Sci. 2020, 563, 370–380. [Google Scholar] [CrossRef] [PubMed]
  107. Khan, M.M.; Harunsani, M.H.; Tan, A.L.; Hojamberdiev, M.; Azamay, S.; Ahmad, N. Antibacterial Activities of Zinc Oxide and Mn-Doped Zinc Oxide Synthesized Using Melastoma malabathricum (L.) Leaf Extract. Bioprocess Biosyst. Eng. 2020, 43, 1499–1508. [Google Scholar] [CrossRef]
  108. Akbar Jan, F.; Wajidullah; Ullah, R.; Ullah, N.; Salman; Usman, M. Exploring the Environmental and Potential Therapeutic Applications of Myrtus communis L. Assisted Synthesized Zinc Oxide (ZnO) and Iron Doped Zinc Oxide (Fe-ZnO) Nanoparticles. J. Saudi Chem. Soc. 2021, 25, 101278. [Google Scholar] [CrossRef]
  109. Okeke, I.S.; Agwu, K.K.; Ubachukwu, A.A.; Madiba, I.G.; Maaza, M.; Whyte, G.M.; Ezema, F.I. Impact of Particle Size and Surface Defects on Antibacterial and Photocatalytic Activities of Undoped and Mg-Doped ZnO Nanoparticles, Biosynthesized Using One-Step Simple Process. Vacuum 2021, 187, 110110. [Google Scholar] [CrossRef]
  110. Rahman, A.; Harunsani, M.H.; Tan, A.L.; Ahmad, N.; Min, B.K.; Khan, M.M. Influence of Mg and Cu Dual-Doping on Phytogenic Synthesized ZnO for Light Induced Antibacterial and Radical Scavenging Activities. Mater. Sci. Semicond. Process. 2021, 128, 105761. [Google Scholar] [CrossRef]
  111. Nguyen, T.H.A.; Le, V.T.; Doan, V.D.; Tran, A.V.; Nguyen, V.C.; Nguyen, A.T.; Vasseghian, Y. Green Synthesis of Nb-Doped ZnO Nanocomposite for Photocatalytic Degradation of Tetracycline Antibiotic under Visible Light. Mater. Lett. 2022, 308, 131129. [Google Scholar] [CrossRef]
  112. Kitching, M.; Ramani, M.; Marsili, E. Fungal Biosynthesis of Gold Nanoparticles: Mechanism and Scale Up. Microb. Biotechnol. 2015, 8, 904–917. [Google Scholar] [CrossRef] [PubMed]
  113. Salvadori, M.R.; Ando, R.A.; Oller Do Nascimento, C.A.; Corrêa, B. Intracellular Biosynthesis and Removal of Copper Nanoparticles by Dead Biomass of Yeast Isolated from the Wastewater of a Mine in the Brazilian Amazonia. PLoS ONE 2014, 9, e87968. [Google Scholar] [CrossRef]
  114. 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] [PubMed]
  115. Velusamy, P.; Kumar, G.V.; Jeyanthi, V.; Das, J.; Pachaiappan, R. Bio-Inspired Green Nanoparticles: Synthesis, Mechanism, and Antibacterial Application. Toxicol. Res. 2016, 32, 95–102. [Google Scholar] [CrossRef] [PubMed]
  116. Tripathi, R.M.; Bhadwal, A.S.; Gupta, R.K.; Singh, P.; Shrivastav, A.; Shrivastav, B.R. ZnO Nanoflowers: Novel Biogenic Synthesis and Enhanced Photocatalytic Activity. J. Photochem. Photobiol. B Biol. 2014, 141, 288–295. [Google Scholar] [CrossRef]
  117. Slavin, Y.N.; Asnis, J.; Häfeli, U.O.; Bach, H. Metal Nanoparticles: Understanding the Mechanisms behind Antibacterial Activity. J. Nanobiotechnol. 2017, 15, 1–20. [Google Scholar] [CrossRef] [PubMed]
  118. Sengani, M.; Grumezescu, A.M.; Rajeswari, V.D. Recent Trends and Methodologies in Gold Nanoparticle Synthesis – A Prospective Review on Drug Delivery Aspect. OpenNano 2017, 2, 37–46. [Google Scholar] [CrossRef]
  119. Mohd Yusof, H.; Mohamad, R.; Zaidan, U.H.; Abdul Rahman, N.A. Microbial Synthesis of Zinc Oxide Nanoparticles and Their Potential Application as an Antimicrobial Agent and a Feed Supplement in Animal Industry: A Review. J. Anim. Sci. Biotechnol. 2019, 10, 1–22. [Google Scholar] [CrossRef] [Green Version]
  120. Shedbalkar, U.; Singh, R.; Wadhwani, S.; Gaidhani, S.; Chopade, B.A. Microbial Synthesis of Gold Nanoparticles: Current Status and Future Prospects. Adv. Colloid Interface Sci. 2014, 209, 40–48. [Google Scholar] [CrossRef]
  121. Jeevanandam, J.; Chan, Y.S.; Danquah, M.K. Biosynthesis of Metal and Metal Oxide Nanoparticles. ChemBioEng Rev. 2016, 3, 55–67. [Google Scholar] [CrossRef]
  122. Molnár, Z.; Bódai, V.; Szakacs, G.; Erdélyi, B.; Fogarassy, Z.; Sáfrán, G.; Varga, T.; Kónya, Z.; Tóth-Szeles, E.; Szucs, R.; et al. Green Synthesis of Gold Nanoparticles by Thermophilic Filamentous Fungi. Sci. Rep. 2018, 8, 1–12. [Google Scholar] [CrossRef] [Green Version]
  123. Hulkoti, N.I.; Taranath, T.C. Biosynthesis of Nanoparticles Using Microbes-A Review. Colloids Surf. B Biointerfaces 2014, 121, 474–483. [Google Scholar] [CrossRef]
  124. Dhandapani, P.; Prakash, A.A.; AlSalhi, M.S.; Maruthamuthu, S.; Devanesan, S.; Rajasekar, A. Ureolytic Bacteria Mediated Synthesis of Hairy ZnO Nanostructure as Photocatalyst for Decolorization of Dyes. Mater. Chem. Phys. 2020, 243, 122619. [Google Scholar] [CrossRef]
  125. Rehman, S.; Jermy, B.R.; Akhtar, S.; Borgio, J.F.; Abdul Azeez, S.; Ravinayagam, V.; Al Jindan, R.; Alsalem, Z.H.; Buhameid, A.; Gani, A. Isolation and Characterization of a Novel Thermophile; Bacillus haynesii, Applied for the Green Synthesis of ZnO Nanoparticles. Artif. Cells Nanomed. Biotechnol. 2019, 47, 2072–2082. [Google Scholar] [CrossRef] [PubMed]
  126. Al-Kordy, H.M.H.; Sabry, S.A.; Mabrouk, M.E.M. Statistical Optimization of Experimental Parameters for Extracellular Synthesis of Zinc Oxide Nanoparticles by a Novel Haloalaliphilic alkalibacillus sp. W7. Sci. Rep. 2021, 11, 1–14. [Google Scholar] [CrossRef]
  127. 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]
  128. 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]
  129. Mohd Yusof, H.; Abdul Rahman, N.A.; Mohamad, R.; Zaidan, U.H.; Samsudin, A.A. Biosynthesis of Zinc Oxide Nanoparticles by Cell-Biomass and Supernatant of Lactobacillus plantarum TA4 and Its Antibacterial and Biocompatibility Properties. Sci. Rep. 2020, 10, 1–13. [Google Scholar] [CrossRef]
  130. Rajivgandhi, G.; Mythili Gnanamangai, B.; Heela Prabha, T.; Poornima, S.; Maruthupandy, M.; Alharbi, N.S.; Kadaikunnan, S.; Li, W.-J. Biosynthesized Zinc Oxide Nanoparticles (ZnO NPs) Using Actinomycetes Enhance the Anti-Bacterial Efficacy against K. pneumoniae. J. King Saud Univ.-Sci. 2022, 34, 101731. [Google Scholar] [CrossRef]
  131. Raliya, R.; Tarafdar, J.C. ZnO Nanoparticle Biosynthesis and Its Effect on Phosphorous-Mobilizing Enzyme Secretion and Gum Contents in Clusterbean (Cyamopsis tetragonoloba L.). Agric. Res. 2013, 2, 48–57. [Google Scholar] [CrossRef] [Green Version]
  132. Rajan, A.; Cherian, E.; Baskar, G. Biosynthesis of Zinc Oxide Nanoparticles Using Aspergillus fumigatus JCF and Its Antibacterial Activity. Int. J. Mod. Sci. Technol. 2016, 1, 52–57. [Google Scholar]
  133. Kalpana, V.N.; Kataru, B.A.S.; Sravani, N.; Vigneshwari, T.; Panneerselvam, A.; Devi Rajeswari, V. Biosynthesis of Zinc Oxide Nanoparticles Using Culture Filtrates of Aspergillus niger: Antimicrobial Textiles and Dye Degradation Studies. OpenNano 2018, 3, 48–55. [Google Scholar] [CrossRef]
  134. Mohana, S.; Sumathi, S. Synthesis of Zinc Oxide Using Agaricus Bisporus and Its In-Vitro Biological Activities. J. Environ. Chem. Eng. 2020, 8, 104192. [Google Scholar] [CrossRef]
  135. 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, 8519–8536. [Google Scholar] [CrossRef]
  136. Moghaddam, A.B.; Moniri, M.; Azizi, S.; Rahim, R.A.; Ariff, A.B.; Saad, W.Z.; Namvar, F.; Navaderi, M.; Mohamad, R. Biosynthesis of ZnO Nanoparticles by a New Pichia kudriavzevii Yeast Strain and Evaluation of Their Antimicrobial and Antioxidant Activities. Molecules 2017, 22, 872. [Google Scholar] [CrossRef] [PubMed]
  137. Kadam, V.V.; Ettiyappan, J.P.; Mohan Balakrishnan, R. Mechanistic Insight into the Endophytic Fungus Mediated Synthesis of Protein Capped ZnO Nanoparticles. Mater. Sci. Eng. B Solid-State Mater. Adv. Technol. 2019, 243, 214–221. [Google Scholar] [CrossRef]
  138. Motazedi, R.; Rahaiee, S.; Zare, M. Efficient Biogenesis of ZnO Nanoparticles Using Extracellular Extract of Saccharomyces cerevisiae: Evaluation of Photocatalytic, Cytotoxic and Other Biological Activities. Bioorg. Chem. 2020, 101, 103998. [Google Scholar] [CrossRef]
  139. Mani, V.M.; Nivetha, S.; Sabarathinam, S.; Barath, S.; Das, M.P.A.; Basha, S.; Elfasakhany, A.; Pugazhendhi, A. Multifunctionalities of Mycosynthesized Zinc Oxide Nanoparticles (ZnONPs) from Cladosporium tenuissimum FCBGr: Antimicrobial Additives for Paints Coating, Functionalized Fabrics and Biomedical Properties. Prog. Org. Coat. 2022, 163, 106650. [Google Scholar] [CrossRef]
  140. Khalafi, T.; Buazar, F.; Ghanemi, K. Phycosynthesis and Enhanced Photocatalytic Activity of Zinc Oxide Nanoparticles Toward Organosulfur Pollutants. Sci. Rep. 2019, 9, 1–10. [Google Scholar] [CrossRef] [Green Version]
  141. El-Belely, E.F.; Farag, M.M.S.; 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] [PubMed]
  142. Rajaboopathi, S.; Thambidurai, S. Enhanced Photocatalytic Activity of Ag-ZnO Nanoparticles Synthesized by Using Padina gymnospora Seaweed Extract. J. Mol. Liq. 2018, 262, 148–160. [Google Scholar] [CrossRef]
  143. Rabie, A.M.; Abukhadra, M.R.; Rady, A.M.; Ahmed, S.A.; Labena, A.; Mohamed, H.S.H.; Betiha, M.A.; Shim, J.J. Instantaneous Photocatalytic Degradation of Malachite Green Dye under Visible Light Using Novel Green Co–ZnO/Algae Composites. Res. Chem. Intermed. 2020, 46, 1955–1973. [Google Scholar] [CrossRef]
  144. Syed, A.; Ahmad, A. Extracellular Biosynthesis of Platinum Nanoparticles Using the Fungus Fusarium oxysporum. Colloids Surf. B Biointerfaces 2012, 97, 27–31. [Google Scholar] [CrossRef] [PubMed]
  145. Zielonka, A.; Klimek-Ochab, M. Fungal Synthesis of Size-Defined Nanoparticles. Adv. Nat. Sci. Nanosci. Nanotechnol. 2017, 8, 043001. [Google Scholar] [CrossRef]
  146. Azizi, S.; Ahmad, M.B.; Namvar, F.; Mohamad, R. Green Biosynthesis and Characterization of Zinc Oxide Nanoparticles Using Brown Marine Macroalga Sargassum muticum Aqueous Extract. Mater. Lett. 2014, 116, 275–277. [Google Scholar] [CrossRef]
  147. Mohamed Isa, E.D.; Shameli, K.; Ch’ng, H.J.; Che Jusoh, N.W.; Hazan, R. Photocatalytic Degradation of Selected Pharmaceuticals Using Green Fabricated Zinc Oxide Nanoparticles. Adv. Powder Technol. 2021, 32, 2398–2409. [Google Scholar] [CrossRef]
  148. El-Saied, H.A.A.; Ibrahim, A.M. Effective Fabrication and Characterization of Eco-Friendly Nano Chitosan Capped Zinc Oxide Nanoparticles for Effective Marine Fouling Inhibition. J. Environ. Chem. Eng. 2020, 8, 103949. [Google Scholar] [CrossRef]
  149. Vijayakumar, T.S.; Mahboob, S.; Bupesh, G.; Vasanth, S.; Al-Ghanim, K.A.; Al-Misned, F.; Govindarajan, M. Facile Synthesis and Biophysical Characterization of Egg Albumen-Wrapped Zinc Oxide Nanoparticles: A Potential Drug Delivery Vehicles for Anticancer Therapy. J. Drug Deliv. Sci. Technol. 2020, 60, 102015. [Google Scholar] [CrossRef]
  150. Qian, C.; Yin, J.; Zhao, J.; Li, X.; Wang, S.; Bai, Z.; Jiao, T. Facile Preparation and Highly Efficient Photodegradation Performances of Self-Assembled Artemia Eggshell-ZnO Nanocomposites for Wastewater Treatment. Colloids Surf. A Physicochem. Eng. Asp. 2021, 610, 125752. [Google Scholar] [CrossRef]
  151. Rekha, R.; Mahboob, S.; Ramya, A.K.; Kerthekeyan, S.; Govindarajan, M.; Al-Ghanim, K.A.; Al-Misned, F.; Ahmed, Z.; Vaseeharan, B. Synthesis and Bio-Physical Characterization of Crustin Capped Zinc Oxide Nanoparticles, and Their Photocatalytic, Antibacterial, Antifungal and Antibiofilm Activity. J. Clust. Sci. 2021, 32, 843–855. [Google Scholar] [CrossRef]
  152. Almehizia, A.A.; Al-Omar, M.A.; Naglah, A.M.; Bhat, M.A.; Al-Shakliah, N.S. Facile Synthesis and Characterization of ZnO Nanoparticles for Studying Their Biological Activities and Photocatalytic Degradation Properties toward Methylene Blue Dye. Alex. Eng. J. 2021, 61, 2386–2395. [Google Scholar] [CrossRef]
  153. Kamaruzaman, A.; Lah, N.A.C. Formation of ZnO Nanoparticles in the Presence of Tannic Acid. Mater. Today Proc. 2020, 41, 61–64. [Google Scholar] [CrossRef]
  154. Ranjithkumar, B.; Ramalingam, H.B.; Kumar, E.R.; Srinivas, C.; Magesh, G.; Rahale, C.S.; El-Metwaly, N.M.; Shekar, B.C. Natural Fuels (Honey and Cow Urine) Assisted Combustion Synthesis of Zinc Oxide Nanoparticles for Antimicrobial Activities. Ceram. Int. 2021, 47, 14475–14481. [Google Scholar] [CrossRef]
  155. Motshekga, S.C.; Sinha Ray, S.; Maity, A. Synthesis and Characterization of Alginate Beads Encapsulated Zinc Oxide Nanoparticles for Bacteria Disinfection in Water. J. Colloid Interface Sci. 2018, 512, 686–692. [Google Scholar] [CrossRef] [PubMed]
  156. Peychev, B.; Vasileva, P. Novel Starch-Mediated Synthesis of Au/ZnO Nanocrystals and Their Photocatalytic Properties. Heliyon 2021, 7, e07402. [Google Scholar] [CrossRef] [PubMed]
  157. Wang, Y.; Wang, H.; Li, Z.; Yang, D.; Qiu, X.; Liu, Y.; Yan, M.; Li, Q. Fabrication of Litchi-like Lignin/Zinc Oxide Composites with Enhanced Antibacterial Activity and Their Application in Polyurethane Films. J. Colloid Interface Sci. 2021, 594, 316–325. [Google Scholar] [CrossRef]
  158. Das, R.K.; Gogoi, N.; Babu, P.J.; Sharma, P.; Mahanta, C.; Bora, U. The Synthesis of Gold Nanoparticles Using Amaranthus spinosus Leaf Extract and Study of Their Optical Properties. Adv. Mater. Phys. Chem. 2012, 2, 275–281. [Google Scholar] [CrossRef]
  159. Umar, H.; Kavaz, D.; Rizaner, N. Biosynthesis of 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] [PubMed] [Green Version]
  160. Barabadi, H.; Honary, S.; Ebrahimi, P.; Mohammadi, M.A.; Alizadeh, A.; Naghibi, F. Microbial Mediated Preparation, Characterization and Optimization of Gold Nanoparticles. Braz. J. Microbiol. 2014, 45, 1493–1501. [Google Scholar] [CrossRef] [Green Version]
  161. Guan, Y.; Yu, H.Y.; Abdalkarim, S.Y.H.; Wang, C.; Tang, F.; Marek, J.; Chen, W.L.; Militky, J.; Yao, J.M. Green One-Step Synthesis of ZnO/Cellulose Nanocrystal Hybrids with Modulated Morphologies and Superfast Absorption of Cationic Dyes. Int. J. Biol. Macromol. 2019, 132, 51–62. [Google Scholar] [CrossRef]
  162. Zheng, B.; Kong, T.; Jing, X.; Odoom-Wubah, T.; Li, X.; Sun, D.; Lu, F.; Zheng, Y.; Huang, J.; Li, Q. Plant-Mediated Synthesis of Platinum Nanoparticles and Its Bioreductive Mechanism. J. Colloid Interface Sci. 2013, 396, 138–145. [Google Scholar] [CrossRef]
  163. Vanaja, M. Kinetic Study on Green Synthesis of Silver Nanoparticles Using Coleus aromaticus Leaf Extract. Adv. Appl. Sci. Res. 2013, 4, 50–55. [Google Scholar]
  164. Jamdagni, P.; Khatri, P.; Rana, J.S. Green Synthesis of Zinc Oxide Nanoparticles Using Flower Extract of Nyctanthes arbor-tristis and Their Antifungal Activity. J. King Saud Univ.-Sci. 2018, 30, 168–175. [Google Scholar] [CrossRef] [Green Version]
  165. Dhandapani, P.; Siddarth, A.S.; Kamalasekaran, S.; Maruthamuthu, S.; Rajagopal, G. Bio-Approach: Ureolytic Bacteria Mediated Synthesis of ZnO Nanocrystals on Cotton Fabric and Evaluation of Their Antibacterial Properties. Carbohydr. Polym. 2014, 103, 448–455. [Google Scholar] [CrossRef] [PubMed]
  166. Shaziman, S.; Ismailrosdi, A.S.; Mamat, M.H.; Zoolfakar, A.S. Influence of Growth Time and Temperature on the Morphology of ZnO Nanorods via Hydrothermal. IOP Conf. Ser. Mater. Sci. Eng. 2015, 99, 1–2. [Google Scholar] [CrossRef] [Green Version]
  167. 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]
  168. Zak, K.; Razali, R.; Majid, W.A.; Darround, M. Synthesis and Characterization of a Narrow Size Distribution of Zinc Oxide Nanoparticles. Int. J. Nanomed. 2011, 6, 1399–1403. [Google Scholar] [CrossRef] [Green Version]
  169. Aldeen, T.S.; Ahmed Mohamed, H.E.; Maaza, M. ZnO Nanoparticles Prepared via a Green Synthesis Approach: Physical Properties, Photocatalytic and Antibacterial Activity. J. Phys. Chem. Solids 2022, 160, 110313. [Google Scholar] [CrossRef]
  170. Gabriel Kaningini, A.; Azizi, S.; Sintwa, N.; Mokalane, K.; Cecilia Mohale, K.; Nixwell Mudau, F.; Maaza, M. Effect of Optimized Precursor Concentration, Temperature, and Doping on Optical Properties of ZnO Nanoparticles Synthesized via a Green Route Using Bush Tea (Athrixia phylicoides DC.) Leaf Extracts. ACS Omega 2022, 7, 31658–31666. [Google Scholar] [CrossRef]
  171. Bharathi, D.; Ranjithkumar, R.; Chandarshekar, B.; Bhuvaneshwari, V. Preparation of Chitosan Coated Zinc Oxide Nanocomposite for Enhanced Antibacterial and Photocatalytic Activity: As a Bionanocomposite. Int. J. Biol. Macromol. 2019, 129, 989–996. [Google Scholar] [CrossRef]
  172. Ayoub, I.; Kumar, V.; Abolhassani, R.; Sehgal, R.R.; Sharma, V.; Sehgal, R.R.; Swart, H.C.; Mishra, Y.K.; Chen, D.; Wang, Z.; et al. Enhanced Photocatalytic Performance for the BiPO4-x Nanorod Induced by Surface Oxygen Vacancy. J. Phys. Chem. C 2022, 5, 18520–18528. [Google Scholar] [CrossRef]
  173. Hafiz, M.; 1ã, M.; Khusaimi, Z.; Zahidi, M.M.; Abu Bakar, S.; Siran, Y.M.; Anuar, S.; Rejab, M.; Asis, A.J.; Tahiruddin, S.; et al. Controllable Growth of Vertically Aligned Aluminum-Doped Zinc Oxide Nanorod Arrays by Sonicated Sol-Gel Immersion Method Depending on Precursor Solution Volumes. Jpn. J. Appl. Phys. 2011, 50, 06GH04. [Google Scholar] [CrossRef]
  174. Chen, W.; Liu, Q.; Tian, S.; Zhao, X. Exposed Facet Dependent Stability of ZnO Micro/Nano Crystals as a Photocatalyst. Appl. Surf. Sci. 2019, 470, 807–816. [Google Scholar] [CrossRef]
  175. Debroye, E.; Van Loon, J.; Yuan, H.; Janssen, K.P.F.; Lou, Z.; Kim, S.; Majima, T.; Roeffaers, M.B.J. Facet-Dependent Photoreduction on Single ZnO Crystals. J. Phys. Chem. Lett. 2017, 8, 340–346. [Google Scholar] [CrossRef] [PubMed]
  176. Adhikari, S.; Sarkar, D.; Madras, G. Synthesis and Photocatalytic Performance of Quasi-Fibrous ZnO. RSC Adv. 2014, 4, 55807–55814. [Google Scholar] [CrossRef]
  177. Zhao, Y.; Cui, T.; Wu, T.; Jin, C.; Qiao, R.; Qian, Y.; Tong, G. Polymorphous ZnO Nanostructures: Zn Polar Surface-Guided Size and Shape Evolution Mechanism and Enhanced Photocatalytic Activity. ChemCatChem 2017, 9, 3180–3190. [Google Scholar] [CrossRef]
  178. Naderi, M. Surface Area: Brunauer–Emmett–Teller (BET). In Progress in Filtration and Separation; Tarleton, S., Ed.; Elsevier& Academic Press: London, UK, 2015; pp. 585–608. ISBN 9780123983077. [Google Scholar]
  179. Salvadori, M.R.; Ando, R.A.; Nascimento, C.A.O.; Corrêa, B. Extra and Intracellular Synthesis of Nickel Oxide Nanoparticles Mediated by Dead Fungal Biomass. PLoS ONE 2015, 10, e0129799. [Google Scholar] [CrossRef] [PubMed]
  180. Dimapilis, E.A.S.; Hsu, C.S.; Mendoza, R.M.O.; Lu, M.C. Zinc Oxide Nanoparticles for Water Disinfection. Sustain. Environ. Res. 2018, 28, 47–56. [Google Scholar] [CrossRef]
  181. Gnanaprakasam, A.; Sivakumar, V.M.; Thirumarimurugan, M. Influencing Parameters in the Photocatalytic Degradation of Organic Effluent via Nanometal Oxide Catalyst: A Review. Indian J. Mater. Sci. 2015, 2015, 1–16. [Google Scholar] [CrossRef]
  182. Rathnasamy, R.; Thangasamy, P.; Thangamuthu, R.; Sampath, S.; Alagan, V. Green Synthesis of ZnO Nanoparticles Using Carica papaya Leaf Extracts for Photocatalytic and Photovoltaic Applications. J. Mater. Sci. Mater. Electron. 2017, 28, 10374–10381. [Google Scholar] [CrossRef]
  183. Ahmed, W.; Chantes, P.; Jarusutthirak, C.; Danwittayakul, S. A Comparison Study of Photocatalytic Activity of TiO2 and ZnO on the Degradation of Real Batik Wastewater. Int. Conf. Biol. Environ. Food Eng. 2015, 16. [Google Scholar] [CrossRef] [Green Version]
  184. Gonçalves, R.A.; Toledo, R.P.; Joshi, N.; Berengue, O.M. Green Synthesis and Applications of Zno and TiO2 Nanostructures. Molecules 2021, 26, 2236. [Google Scholar] [CrossRef]
  185. Sin, J.C.; Lam, S.M.; Satoshi, I.; Lee, K.T.; Mohamed, A.R. Sunlight Photocatalytic Activity Enhancement and Mechanism of Novel Europium-Doped ZnO Hierarchical Micro/Nanospheres for Degradation of Phenol. Appl. Catal. B Environ. 2014, 148–149, 258–268. [Google Scholar] [CrossRef]
  186. Nava, O.J.; Soto-Robles, C.A.; Gómez-Gutiérrez, C.M.; Vilchis-Nestor, A.R.; Castro-Beltrán, A.; Olivas, A.; Luque, P.A. Fruit Peel Extract Mediated Green Synthesis of Zinc Oxide Nanoparticles. J. Mol. Struct. 2017, 1147, 1–6. [Google Scholar] [CrossRef]
  187. Vinayagam, R.; Selvaraj, R.; Arivalagan, P.; Varadavenkatesan, T. Synthesis, Characterization and Photocatalytic Dye Degradation Capability of Calliandra haematocephala-Mediated Zinc Oxide Nanoflowers. J. Photochem. Photobiol. B Biol. 2020, 203, 111760. [Google Scholar] [CrossRef] [PubMed]
  188. Mirgane, N.A.; Shivankar, V.S.; Kotwal, S.B.; Wadhawa, G.C.; Sonawale, M.C. Degradation of Dyes Using Biologically Synthesized Zinc Oxide Nanoparticles. Mater. Today Proc. 2020, 37, 849–853. [Google Scholar] [CrossRef]
  189. Nava, O.J.; Luque, P.A.; Gómez-Gutiérrez, C.M.; Vilchis-Nestor, A.R.; Castro-Beltrán, A.; Mota-González, M.L.; Olivas, A. Influence of Camellia sinensis Extract on Zinc Oxide Nanoparticle Green Synthesis. J. Mol. Struct. 2017, 1134, 121–125. [Google Scholar] [CrossRef]
  190. Vidya, C.; Manjunatha, C.; Chandraprabha, M.N.; Rajshekar, M.; Antony Raj, M.A.L. Hazard Free Green Synthesis of ZnO Nano-Photo-Catalyst Using Artocarpus heterophyllus Leaf Extract for the Degradation of Congo Red Dye in Water Treatment Applications. J. Environ. Chem. Eng. 2017, 5, 3172–3180. [Google Scholar] [CrossRef]
  191. Razanamahandry, L.C.; Sackey, J.; Furqan, C.M.; Ntwampe, S.K.O.; Fosso-Kankeu, E.; Manikandan, E.; Maaza, M. Removal of Free Cyanide by a Green Photocatalyst ZnO Nanoparticle Synthesized via Eucalyptus globulus Leaves. In Photocatalysts in Advanced Oxidation Processes for Wastewater Treatment; Elvis Fosso-Kankeu, Pandey, S., Ray, S.S., Eds.; Srinevener Publishing, Wiley: Beverley, UK, 2020; pp. 271–288. ISBN 9781119631422. [Google Scholar]
  192. Patil, S.S.; Mali, M.G.; Tamboli, M.S.; Patil, D.R.; Kulkarni, M.V.; Yoon, H.; Kim, H.; Al-Deyab, S.S.; Yoon, S.S.; Kolekar, S.S.; et al. Green Approach for Hierarchical Nanostructured Ag-ZnO and Their Photocatalytic Performance under Sunlight. Catal. Today 2016, 260, 126–134. [Google Scholar] [CrossRef]
  193. Noukelag, S.K.; Razanamahandry, L.C.; Ntwampe, S.K.O.; Arendse, C.J.; Maaza, M. Industrial Dye Removal Using Bio-Synthesized Ag-Doped ZnO Nanoparticles. Environ. Nanotechnol. Monit. Manag. 2021, 16, 100463. [Google Scholar] [CrossRef]
  194. Karthik, K.V.; Raghu, A.V.; Reddy, K.R.; Ravishankar, R.; Sangeeta, M.; Shetti, N.P.; Reddy, C.V. Green Synthesis of Cu-Doped ZnO Nanoparticles and Its Application for the Photocatalytic Degradation of Hazardous Organic Pollutants. Chemosphere 2022, 287, 132081. [Google Scholar] [CrossRef]
  195. Li, J.F.; Rupa, E.J.; Hurh, J.; Huo, Y.; Chen, L.; Han, Y.; Ahn, J.C.; Park, J.K.; Lee, H.A.; Mathiyalagan, R.; et al. Cordyceps militaris Fungus Mediated Zinc Oxide Nanoparticles for the Photocatalytic Degradation of Methylene Blue Dye. Optik 2019, 183, 691–697. [Google Scholar] [CrossRef]
  196. Kumar, N.; Bhadwal, A.S.; Garg, M.; Sharma, R.; Singh, S.; Mizaikoff, B. Photocatalytic and Antibacterial Biomimetic ZnO Nanoparticles. Anal. Methods 2017, 9, 4776–4782. [Google Scholar] [CrossRef]
  197. Deyab, M.; Elkatony, T.; Ward, F. Qualitative and Quantitative Analysis of Phytochemical Studies on Brown Seaweed, Dictyota dichotoma. Int. J. Eng. Dev. Res. 2016, 4, 2321–9939. [Google Scholar]
  198. Preethi, J.; Farzana, M.H.; Meenakshi, S. Photo-Reduction of Cr(VI) Using Chitosan Supported Zinc Oxide Materials. Int. J. Biol. Macromol. 2017, 104, 1783–1793. [Google Scholar] [CrossRef] [PubMed]
  199. Li, X.; Zhang, L.; Wang, Z.; Wu, S.; Ma, J. Cellulose Controlled Zinc Oxide Nanoparticles with Adjustable Morphology and Their Photocatalytic Performances. Carbohydr. Polym. 2021, 259, 117752. [Google Scholar] [CrossRef] [PubMed]
  200. Rashtbari, Y.; Hazrati, S.; Azari, A.; Afshin, S.; Fazlzadeh, M.; Vosoughi, M. A Novel, Eco-Friendly and Green Synthesis of PPAC-ZnO and PPAC-NZVI Nanocomposite Using Pomegranate Peel: Cephalexin Adsorption Experiments, Mechanisms, Isotherms and Kinetics. Adv. Powder Technol. 2020, 31, 1612–1623. [Google Scholar] [CrossRef]
  201. Yokwana, K.; Kuvarega, A.T.; Mhlanga, S.D.; Nxumalo, E.N. Mechanistic Aspects for the Removal of Congo Red Dye from Aqueous Media through Adsorption over N-Doped Graphene Oxide Nanoadsorbents Prepared from Graphite Flakes and Powders. Phys. Chem. Earth 2018, 107, 58–70. [Google Scholar] [CrossRef]
  202. Nagar, N.; Devra, V. A Kinetic Study on the Degradation and Biodegradability of Silver Nanoparticles Catalyzed Methyl Orange and Textile Effluents. Heliyon 2019, 5, e01356. [Google Scholar] [CrossRef] [Green Version]
  203. Mahdavi, S.; Jalali, M.; Afkhami, A. Removal of Heavy Metals from Aqueous Solutions Using Fe3O4, ZnO, and CuO Nanoparticles. Nanotechnol. Sustain. Dev. 2014, 846, 171–188. [Google Scholar] [CrossRef]
  204. Kim, K.M.; Choi, M.H.; Lee, J.K.; Jeong, J.; Kim, Y.R.; Kim, M.K.; Paek, S.M.; Oh, J.M. Physicochemical Properties of Surface Charge-Modified ZnO Nanoparticles with Different Particle Sizes. Int. J. Nanomed. 2014, 9, 41–56. [Google Scholar] [CrossRef] [Green Version]
  205. Nayak, A.; Sahoo, J.K.; Sahoo, S.K.; Sahu, D. Removal of Congo Red Dye from Aqueous Solution Using Zinc Oxide Nanoparticles Synthesised from Ocimum sanctum (Tulsi Leaf): A Green Approach. Int. J. Environ. Anal. Chem. 2020, 1–22. [Google Scholar] [CrossRef]
  206. Muthukumaran, P.; Babu, P.S.; Shyamalagowri, S.; Kamaraj, M.; Manikandan, A.; Aravind, J. Nanotechnological Approaches as a Promising Way for Heavy Metal Mitigation in an Aqueous System. J. Basic Microbiol. 2021, 62, 376–394. [Google Scholar] [CrossRef] [PubMed]
  207. Le, A.T.; Pung, S.Y.; Sreekantan, S.; Matsuda, A.; Huynh, D.P. Mechanisms of Removal of Heavy Metal Ions by ZnO Particles. Heliyon 2019, 5, e01440. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  208. Joshi, N.; Singh, A. Adsorptive Performances and Characterisations of Biologically Synthesised Zinc Oxide Based Nanosorbent (ZOBN). Groundw. Sustain. Dev. 2020, 10, 100325. [Google Scholar] [CrossRef]
  209. Tshangana, C.; Chabalala, M.; Muleja, A.; Nxumalo, E.; Mamba, B. Shape-Dependant Photocatalytic and Antimicrobial Activity of ZnO Nanostructures When Conjugated to Graphene Quantum Dots. J. Environ. Chem. Eng. 2020, 8, 103930. [Google Scholar] [CrossRef]
  210. Sun, H.; Brocato, J.; Costa, M. Oral Chromium Exposure and Toxicity. Curr. Environ. Health Rep. 2015, 2, 295–303. [Google Scholar] [CrossRef]
  211. Khan, T.A.; Nazir, M.; Ali, I.; Kumar, A. Removal of Chromium(VI) from Aqueous Solution Using Guar Gum–Nano Zinc Oxide Biocomposite Adsorbent. Arab. J. Chem. 2017, 10, S2388–S2398. [Google Scholar] [CrossRef]
  212. Kamaraj, M.; Srinivasan, N.R.; Assefa, G.; Adugna, A.T.; Kebede, M. Facile Development of Sunlit ZnO Nanoparticles-Activated Carbon Hybrid from Pernicious Weed as an Operative Nano-Adsorbent for Removal of Methylene Blue and Chromium from Aqueous Solution: Extended Application in Tannery Industrial Wastewater. Environ. Technol. Innov. 2020, 17, 100540. [Google Scholar] [CrossRef]
  213. Wani, A.L.; Ara, A.; Usmani, J.A. Lead Toxicity: A Review. Interdiscip. Toxicol. 2015, 8, 55–64. [Google Scholar] [CrossRef] [Green Version]
  214. Gordon, B.; Callan, P.; Vickers, C. WHO Guidelines for Drinking-Water Quality. WHO Chron. 2008, 38, 564. [Google Scholar] [CrossRef]
  215. Somu, P.; Paul, S. Casein Based Biogenic-Synthesized Zinc Oxide Nanoparticles Simultaneously Decontaminate Heavy Metals, Dyes, and Pathogenic Microbes: A Rational Strategy for Wastewater Treatment. J. Chem. Technol. Biotechnol. 2018, 93, 2962–2976. [Google Scholar] [CrossRef]
  216. Joshi, N.; Malik, S.; Gururani, P. Utilization of Polypyrrole/ZnO Nanocomposite in the Adsorptive Removal of Cu2+, Pb2+ and Cd2+ Ions from Wastewater. Lett. Appl. NanoBioScience 2020, 10, 2339–2351. [Google Scholar] [CrossRef]
  217. Nel, L.; Strydom, N.A.; Bouwman, H. Preliminary Assessment of Contaminants in the Sediment and Organisms of the Swartkops Estuary, South Africa. Mar. Pollut. Bull. 2015, 101, 878–885. [Google Scholar] [CrossRef] [PubMed]
  218. Debnath, P.; Mondal, N.K. Effective Removal of Congo Red Dye from Aqueous Solution Using Biosynthesized Zinc Oxide Nanoparticles. Environ. Nanotechnol. Monit. Manag. 2020, 14, 100320. [Google Scholar] [CrossRef]
  219. Arab, C.; El Kurdi, R.; Patra, D. Efficient Removal of Congo Red Using Curcumin Conjugated Zinc Oxide Nanoparticles as New Adsorbent Complex. Chemosphere 2021, 276, 1–3. [Google Scholar] [CrossRef]
  220. Jose, L.M.; Raj, R.S.A.; Sajan, D.; Aravind, A. Adsorption and Photocatalytic Activity of Biosynthesised ZnO Nanoparticles Using Aloe vera Leaf Extract. Nano Express 2021, 2, 010039. [Google Scholar] [CrossRef]
  221. Mittal, H.; Morajkar, P.P.; Al Alili, A.; Alhassan, S.M. In-Situ Synthesis of ZnO Nanoparticles Using Gum Arabic Based Hydrogels as a Self-Template for Effective Malachite Green Dye Adsorption. J. Polym. Environ. 2020, 28, 1637–1653. [Google Scholar] [CrossRef]
  222. Kahsay, M.H. Synthesis and Characterization of ZnO Nanoparticles Using Aqueous Extract of Becium grandiflorum for Antimicrobial Activity and Adsorption of Methylene Blue. Appl. Water Sci. 2021, 11, 1–12. [Google Scholar] [CrossRef]
  223. Oyewo, O.A.; Adeniyi, A.; Sithole, B.B.; Onyango, M.S. Sawdust-Based Cellulose Nanocrystals Incorporated with ZnO Nanoparticles as Efficient Adsorption Media in the Removal of Methylene Blue Dye. ACS Omega 2020, 5, 18798–18807. [Google Scholar] [CrossRef]
  224. Akpomie, K.G.; Conradie, J. Synthesis, Characterization, and Regeneration of an Inorganic–Organic Nanocomposite (ZnO@biomass) and Its Application in the Capture of Cationic Dye. Sci. Rep. 2020, 10, 1–12. [Google Scholar] [CrossRef]
  225. Singh, J.; Vishwakarma, K.; Ramawat, N.; Rai, P.; Singh, V.K.; Mishra, R.K.; Kumar, V.; Tripathi, D.K.; Sharma, S. Nanomaterials and Microbes’ Interactions: A Contemporary Overview. 3 Biotech 2019, 9, 1–14. [Google Scholar] [CrossRef]
  226. Soren, S.; Kumar, S.; Mishra, S.; Jena, P.K.; Verma, S.K.; Parhi, P. Evaluation of Antibacterial and Antioxidant Potential of the Zinc Oxide Nanoparticles Synthesized by Aqueous and Polyol Method. Microb. Pathog. 2018, 119, 145–151. [Google Scholar] [CrossRef] [PubMed]
  227. Padmavathy, N.; Vijayaraghavan, R. Enhanced Bioactivity of ZnO Nanoparticles—An Antimicrobial Study. Sci. Technol. Adv. Mater. 2008, 9, 035004. [Google Scholar] [CrossRef] [PubMed]
  228. Stoimenov, P.K.; Klinger, R.L.; Marchin, G.L.; Klabunde, K.J. Metal Oxide Nanoparticles as Bactericidal Agents. Langmuir 2002, 18, 6679–6686. [Google Scholar] [CrossRef]
  229. Mahendra, C.; Murali, M.; Manasa, G.; Ponnamma, P.; Abhilash, M.R.; Lakshmeesha, T.R.; Satish, A.; Amruthesh, K.N.; Sudarshana, M.S. Antibacterial and Antimitotic Potential of Bio-Fabricated Zinc Oxide Nanoparticles of Cochlospermum religiosum (L.). Microb. Pathog. 2017, 110, 620–629. [Google Scholar] [CrossRef]
  230. Agarwal, H.; Menon, S.; Shanmugam, V.K. Functionalization of Zinc Oxide Nanoparticles Using Mucuna pruriens and Its Antibacterial Activity. Surf. Interfaces 2020, 19, 100521. [Google Scholar] [CrossRef]
  231. Theophil Anand, G.; Renuka, D.; Ramesh, R.; Anandaraj, L.; John Sundaram, S.; Ramalingam, G.; Magdalane, C.M.; Bashir, A.K.H.; Maaza, M.; Kaviyarasu, K. Green Synthesis of ZnO Nanoparticle Using Prunus dulcis (Almond Gum) for Antimicrobial and Supercapacitor Applications. Surf. Interfaces 2019, 17, 100376. [Google Scholar] [CrossRef]
  232. Kokabi, M.; Yousefzadi, M.; Nejad Ebrahimi, S.; Zarei, M. Green Synthesis of Zinc Oxide Nanoparticles Using Seaweed Aqueous Extract and Evaluation of Antibacterial and Ecotoxicological Activity. J. Persian Gulf (Mar. Sci.) 2017, 8, 61–72. [Google Scholar] [CrossRef]
  233. Jamil, Z.; Naqvi, S.T.Q.; Rasul, S.; Hussain, S.B.; Fatima, N.; Qadir, M.I.; Muhammad, S.A. Antibacterial Activity and Characterization of Zinc Oxide Nanoparticles Synthesized by Microalgae. Pak. J. Pharm. Sci. 2020, 33, 2497–2504. [Google Scholar] [CrossRef]
  234. Subramanian, H.; Krishnan, M.; Mahalingam, A. Photocatalytic Dye Degradation and Photoexcited Anti-Microbial Activities of Green Zinc Oxide Nanoparticles Synthesized via Sargassum muticum Extracts. RSC Adv. 2022, 12, 985–997. [Google Scholar] [CrossRef]
  235. Gharpure, S.; Yadwade, R.; Ankamwar, B. Non-Antimicrobial and Non-Anticancer Properties of ZnO Nanoparticles Biosynthesized Using Different Plant Parts of Bixa orellana. ACS Omega 2022, 7, 1914–1933. [Google Scholar] [CrossRef]
  236. Mocanu, A.; Isopencu, G.; Busuioc, C.; Popa, O.M.; Dietrich, P.; Socaciu-Siebert, L. Bacterial Cellulose Films with ZnO Nanoparticles and Propolis Extracts: Synergistic Antimicrobial Effect. Sci. Rep. 2019, 9, 1–10. [Google Scholar] [CrossRef]
Figure 1. ZnO stable polymorphs (Reprinted with permission from Ref. [32]).
Figure 1. ZnO stable polymorphs (Reprinted with permission from Ref. [32]).
Catalysts 12 01442 g001
Figure 2. ZnO morphologies: (a) nanorods (Reprinted with permission from Ref. [33], © 2020, American Institute of Chemical Engineers) (b) spherical (Reprinted with permission from Ref. [34], © 2020, The Society for Applied Microbiology) (c) nanosheets [24], (d) hexagonal [35], (e) tripods (Reprinted with permission from Ref. [36], © 2015, American Chemical Society) (f) nanowires (Reprinted from Ref. [27], © 2020, American Chemical Society) (g) nanotubes, (Reprinted with permission from Ref. [28], © IOP Publishing Ltd) (h) nanostars [29], (i) nanoflowers [37], and (j) nanocubes (Reprinted with permission from Ref. [31],© 2016, Elsevier Inc).
Figure 2. ZnO morphologies: (a) nanorods (Reprinted with permission from Ref. [33], © 2020, American Institute of Chemical Engineers) (b) spherical (Reprinted with permission from Ref. [34], © 2020, The Society for Applied Microbiology) (c) nanosheets [24], (d) hexagonal [35], (e) tripods (Reprinted with permission from Ref. [36], © 2015, American Chemical Society) (f) nanowires (Reprinted from Ref. [27], © 2020, American Chemical Society) (g) nanotubes, (Reprinted with permission from Ref. [28], © IOP Publishing Ltd) (h) nanostars [29], (i) nanoflowers [37], and (j) nanocubes (Reprinted with permission from Ref. [31],© 2016, Elsevier Inc).
Catalysts 12 01442 g002
Figure 3. Schematic representation for the green synthesis of ZnO nanoparticles.
Figure 3. Schematic representation for the green synthesis of ZnO nanoparticles.
Catalysts 12 01442 g003
Figure 4. A proposed mechanism of ZnO formation via the phytosynthesis route. (Reprinted with permission from Ref. [77], © 2019, Elsevier Ltd.).
Figure 4. A proposed mechanism of ZnO formation via the phytosynthesis route. (Reprinted with permission from Ref. [77], © 2019, Elsevier Ltd.).
Catalysts 12 01442 g004
Figure 5. Schematic diagram of the synthesis of Mg/Cu dual-doped ZnO. (Reprinted with permission from Ref. [109], © 2021, Elsevier Ltd.).
Figure 5. Schematic diagram of the synthesis of Mg/Cu dual-doped ZnO. (Reprinted with permission from Ref. [109], © 2021, Elsevier Ltd.).
Catalysts 12 01442 g005
Figure 6. Schematic representation of the (a) microbial intracellular synthesis and (b) microbial extracellular synthesis of ZnO NPs [119].
Figure 6. Schematic representation of the (a) microbial intracellular synthesis and (b) microbial extracellular synthesis of ZnO NPs [119].
Catalysts 12 01442 g006
Figure 7. Antimicrobial activity of Aspergillus fumigatus synthesized ZnO NPs [132] working against (A) Klebsiella pneumonia, (B) P. aeruginosa, (C) E. coli, (D) S. aureus, and (E) B. subtilis at different concentrations.
Figure 7. Antimicrobial activity of Aspergillus fumigatus synthesized ZnO NPs [132] working against (A) Klebsiella pneumonia, (B) P. aeruginosa, (C) E. coli, (D) S. aureus, and (E) B. subtilis at different concentrations.
Catalysts 12 01442 g007
Figure 8. The probable method for the algal production of ZnO NPs, using aqueous Chlorella extract, is depicted schematically [140].
Figure 8. The probable method for the algal production of ZnO NPs, using aqueous Chlorella extract, is depicted schematically [140].
Catalysts 12 01442 g008
Figure 9. XRD spectra for ZnO (a) synthesized using Chlorella microalgae extract that was dried at 50 °C [140]; (b) synthesized using cyanobacteria extract and dried at 60 °C [29]; and (c) the possible growth directions for the ZnO Wurtzite crystals ( lattice parameters a = 0.325 nm and c = 0.521 nm) [102].
Figure 9. XRD spectra for ZnO (a) synthesized using Chlorella microalgae extract that was dried at 50 °C [140]; (b) synthesized using cyanobacteria extract and dried at 60 °C [29]; and (c) the possible growth directions for the ZnO Wurtzite crystals ( lattice parameters a = 0.325 nm and c = 0.521 nm) [102].
Catalysts 12 01442 g009
Figure 11. (a) Graph showing the amount of remaining MB according to irradiation time; (b) chemical kinetics for the photocatalytic degradation of MB using ZnO, Ag-ZnO, and P25 TiO2 catalysts. Figure adapted with permission from Ref. [192], © 2015, Elsevier B.V.
Figure 11. (a) Graph showing the amount of remaining MB according to irradiation time; (b) chemical kinetics for the photocatalytic degradation of MB using ZnO, Ag-ZnO, and P25 TiO2 catalysts. Figure adapted with permission from Ref. [192], © 2015, Elsevier B.V.
Catalysts 12 01442 g011
Figure 12. Mechanisms for HM ion adsorption with ZnO NPs: (a) adsorption (reprinted with permission from Ref. [207], © 2019 Elsevier Ltd); (b) schematic representation of the adsorption mechanism of Congo red onto ZnO-T nanoparticles (reprinted with permission from Ref. [205], © Taylor & Francis Ltd.).
Figure 12. Mechanisms for HM ion adsorption with ZnO NPs: (a) adsorption (reprinted with permission from Ref. [207], © 2019 Elsevier Ltd); (b) schematic representation of the adsorption mechanism of Congo red onto ZnO-T nanoparticles (reprinted with permission from Ref. [205], © Taylor & Francis Ltd.).
Catalysts 12 01442 g012
Figure 13. A schematic representation of the ZnO NPs’ antimicrobial mechanism against bacteria [49].
Figure 13. A schematic representation of the ZnO NPs’ antimicrobial mechanism against bacteria [49].
Catalysts 12 01442 g013
Table 2. Phytosynthesis of the doped ZnO nanoparticles.
Table 2. Phytosynthesis of the doped ZnO nanoparticles.
Serial No.Plant NameDopantExtraction Solvent, Temp and TimePrecursorsBiosynthesis Time (h) and Temp (°C)Calcination
Temp (°C)/Time (h)
ShapeSize (nm)Ref
1Calotropis gigantea (leaves)AgRefluxingZn(NO3)2·6H2O
AgNO3
3/RT400/3RodsNone[33]
2Moringa oleifera (seeds)AgAqueous powder, RTZn(OOCCH3)2
AgNO3
50 min/RT-Flower-like54.1[58]
3Ocimum tenuiflorum (seeds)AgWater, boiledZn(NO3)2·6H2O
AgNO3
2/RT450/3Hexagonal and spherical50–60[106]
4Melastoma malabathricum (leaves)MnWater, 60 °CZn(NO3)2·6H2O
Mn(CH3CO2)2·4H2O
–/60400/2Spherical222[107]
5Myrtus communis L.FeWater, 40 °C for 1 hZn(NO3)2·6H2O
Fe(NO)3·9H2O
3/60400/3Pseudo spherical17[108]
6Piper guineense (leaves)MgWater, 60 °C for 20 minZn(NO3)2
Mg(NO3)2
0.25/80350/2Spherical364[109]
7Ziziphus mauritiana Lam. (leaves)Mg and CuWater, RT for 1 hZn(NO3)2·6H2O
Cu(NO3)2·3H2O
Mg(NO3)2
–/60400/2Hexagonal disc shaped100–1000[110]
8Vernonia amygdalinaNbNot availableNot mentioned1/60400/1Spherical1000[111]
Table 3. Microbial synthesis of ZnO nanoparticles.
Table 3. Microbial synthesis of ZnO nanoparticles.
Serial No.Microbe NameExtraction MethodSaltBiosynthesis
Time (h)/Temp (°C)
Calcination
Temp (°C)
Time (h)
ShapeSize (nm)Ref
Bacteria
1Cyanobacterium Nostoc sp.Water, 60 °C for 15 minZn(CH3COO)2·2H2O2/RTStar-shaped60[29]
2Aeromonas hydrophilaCell culture grew for 24 h and was diluted with waterZnO24/30 [59]
3Serratia nematodiphila.Inoculated in Luria-Bertani (LB) brothZnSO40.17/80
Then stored for 24 h
Near spherical15–30[105]
4Bacillus subtillisCell culture in ammonium carbonateZn(CH3COO)23/80250/3Spheres with hair-like appendages10–15 and
1.2–1.3 um aggregates
[124]
5Bacillus haynesiiCell culture grown on agarZnSO424/55700/5Spherical20–200[125]
6Alkalibacillus sp.Cell-free supernatant collected by centrifugationZnSO4·7H2O48/35Quasi-spherical1–30[126]
7Pseudomonas putidaCell culture solutionZn (NO3)224/37400/2Spherical agglomerates25–45[127]
8Halomonas elongataCentrifugationZnCl2TaguchiMultiform shapes18.11[128]
9Lactobacillus plantarumCell biomass collected by centrifugationZn(NO3)2·6H2O,24/37Flower-like152.8–613.5[129]
10Streptomyces spp.Broth dilution then centrifugationZnSO40.25/40400/8Needle-like12–35[130]
Fungi
11Aspergillus aeneusCell-free filtrateZn (NO3)248/28Spherical100–140[60]
12Aspergillus fumigatusPlating the inoculum on Martin Rose-Bengal agar mediumZn (NO3)272/28Oblate spherical1.2–6.8[131]
13Aspergillus fumigatusCell biomassZnSO472/32Spherical60–80[132]
14Aspergillus niger.Culture filtrateZn(NO3)248/32Near spherical53–69[133]
15Agaricus bisporusPurchasedZn(NO3)2·6H2O2/90400/4Spheroids32[134]
16Xylaria acutaCultured then centrifugedZn(NO3)2·6H2O–/400700/2Hexagonal34–55[135]
17Pichia kudriavzeviiMycelia separated from the cultureZn(CH3COO)2·2H2O36/35Hexagonal10–61 (TEM, XRD)[136]
18Cochliobolus geniculatusMycelial-free filtrateZn(CH3COO)272/28Quasi-spherical2–6 (TEM)[137]
19Saccharomyces cerevisiaeCell-free filtrateZn(CH3COO)2·2H2O24/30Spherical20–30[138]
20Cladosporium tenuissimum FCBGrExtracellular cultureZn(NO3)236/RTBouquet-like57[139]
Algae
21ChlorellaSoaking methodZn(CH3COO)2·2H2O1/58Hexagonal20[140]
22Arthrospira platensisBiomass mixed with supernatantZn(CH3COO)2·2H2O24/30Spherical30–55[141]
23Padina gymnosporaMillipore water stirred for 2 h at 100 °CZn(NO3)2
AgNO3
2/60Spherical20–40[142]
24Sargassum speciesBall-milledZn (NO3)2·6H2O
Co(NO3)2·6H2O
2/RTNearly spherical5.4–6.8[143]
Table 4. ZnO nanoparticles prepared from biological derivatives.
Table 4. ZnO nanoparticles prepared from biological derivatives.
Serial No.Bio-ProductPrecursorsBiosynthesis
Temp (°C)/Duration
Thermal Treatment
Temp (°C)/Time (h)
ShapeSize (nm)Ref
1AlginatePurchased ZnO Irregular spheres20–90[155]
2StarchZn(NO3)2·6H2O
HAuCl4·3H2O
Stirred at 90–100 °COven dried at 80 °C
Calcined at 600 °C for 4 h
Spherical18.9[156]
3LigninZn(NO3)2·6H2OOven at 120 °C for 4 hOven dried at 120 °C for 4 hRods5–10[157]
4PullulanZn(NO3)2·6H2O,Stirred at RT for 5 hOven dried at 60 °C for 24 h
Calcined at 400 °C for 1 h
Spherical and hexagonal58.13[130]
5ChitosanZn(NO3)2·6H2OStirred at 80 °C for 2 hOven dried at 80 °C for 24 hRods50–70[148]
6Egg albumenZn(CH3COO)2Hydrothermal, at 85 °C for 3 hSpherical and diagonal platelet20–60[149]
7Artemia eggshellZn(CH3COO)2Hydrothermal, 180 °C for 20 hEllipsoidal50[150]
8CrustinZn(CH3COO)2Co-precipitation, magnetically stirred for 2 hVacuum dried at 30 °C 50[151]
9Amino acidsZn(NO3)2·6H2OStirred at 22 °C for 30 min prior to increasing the temperature to 150 °CCalcined at 550 °C for
3 h 20 min
Sheet-like particles22.46–40.29[152]
10Tannic acidZn powderStirred at 70 °C for 1 h and 30 minSpherical26–34[153]
11Honey and cow urineZn(NO3)2·6H2OCombustion method, further stirred at 100 °CHexagonal, leaf-like structure (cow urine)
Spherical (honey)
27
44
[154]
Table 7. Summary of the reported antimicrobial activities of ZnO nanoparticles that were synthesized from various biological sources.
Table 7. Summary of the reported antimicrobial activities of ZnO nanoparticles that were synthesized from various biological sources.
SourceName/PartZnO MorphologyMethodConcentration of ZnO NP and MicrobeMIC (μg/mL)Test Organism and EfficiencyRef.
Antibacterial
PlantRubus fairholmianus rootSphericalAgar well diffusionN/A
1 × 105 CFU/mL bacteria
157.22S. aureus[100]
Ficus palmata.SphericalAgar well diffusionN/A/
N/A
390.6S. aureus (19.6 mm)
B. subtilus (21.3 mm)
E coli (18.6 mm)
S. typhi (20.6 mm)
K. pneumoniae (21.6 mm)
[73]
Cochlospermum religiosum (L.)HexagonalDisc diffusion50 μL of (100 μg)
N/A
N/A/
100 μL of 1.5 × 108 CFU/mL bacteria
4.8 to 625B. subtilis
S. aureus
E. coli
P. aeruginosa
[229]
Vitex trifolia L.SphericalDisc diffusion200 µg/mL
N/A
6.25–50ZOI=
S. aureus (22.5 mm)
B. subtilis (21.5 mm)
P. aeruginosa (19.8 mm)
P. mirabilis (19.3 mm)
E. coli (20.1 mm)
[70]
Mucuna
pruriens
Nano-flower-likeBroth dilution assay100 µg/mL/
N/A
20B. subtilis[230]
Phoenix roebeleniiSphericalAgar well diffusion0.4 mg/mL/
N/A
-S. typhi (21.4 mm)
E. coli (15.8 mm)
S. pneumoniae (15.1 mm)
S. aureus (15 mm)
[169]
Prunus dulcisNear sphericalWell diffusion methodN/A/
Standard inoculum
25S. aureus (18 mm)
E. coli (32 mm)
S. paratyphi (25 mm)
K. pneumoniae (0 nm)
P. mirabilius (0 nm)
[231]
AlgaeHypnea musciformisSphericalSerial dilution32 mg/mL S. aureus
E. coli
[232]
Phormidium, Viridiplanta, and Cosmarium sp.Varying shapesWell diffusion60 μL ofE. coli (20 mm)
S. aureus (14.5 mm)
K. pneumoniae (11 mm), C. freundii (20 mm)
M. luteus (12 mm)
[233]
Sargassum muticum Agar well diffusion1.5 × 108 CFU/mL200 mg/mLP. stutzeri
A. baumannii (18 mm)
B. filamentous
Bacillus flexus (12 mm)
[234]
FungiAspergillus fumigatus Well diffusion1000 μg/mLK. pneumoniae (30 mm)
P. aeruginosa (27 mm)
E. coli (25 mm)
S. aureus (30 mm)
B. subtilis (20 mm)
[132]
Pichia kudriavzeviiHexagonalDisc diffusion100 μg/mLB. subtilis (19 mm
S. epidermidis (9 mm)
S. aurous (10 mm)
E. coli (10 mm)
S. marcescens (mm)
[136]
Saccharomyces cerevisiaeSphericalColony forming units100 μg/mL
108 CFU/mL E. coli
E. coli[196]
Biological derivativesLigninLitchi-likeOptical density200 μg/mL100 lg/mLInactivation of
E. coli (97.54%)
S. aureus (99.55%)
[157]
CrustinSphericalAgar well diffusion50 μg/mLThe inhibition zone for
S. mutans (9.5 mm)
L. fusiformis (4.5 mm) Proteus vulgaris 6.1 mm, A. hydrophila (7.2 mm)
[151]
Honey and cow urineHexagonal and leaf-likeDilution and plate count1 × 108 CFU/mLE. coli
B. subtilis
[154]
ChitosanRodsDisc diffusion40 μg/mL
105 CFU/mL bacteria
E. coli (25.5 mm)
K.pneumoniae (24.5 mm)
S. aureus (22.5 mm)
B. subtilis (21 mm)
[171]
Antifungal
PlantPithecellobium dulce (peel)SphericalBroth dilution1000 mg/L ZnO
10 mL of nutrient
potato dextrose agar/1 mL of fungal stains.
Growth inhibition
A. flavus (63.57%)
A. niger (57.17)
[83]
Capparis zeylanicaSphericalWell diffusion100 μg/mLZOI=
C. albicans (30 mm)
A. niger (12 mm)
[91]
Bixa orellanaSpherical
Almond-shaped
Well diffusion10 mg/mL
104–105 CFU/mL (100 μL fungal spores)
Penicillium sp.
Aspergillus flavus
Fusarium oxysporum,
Rhizoctonia solani
[235]
GelatinSphericalLight microscopy50 μg/mLC. albicans[62]
Bacterial cellulose with propolisIrregularDisc diffusion0.4987/143 mL0.438 mg/mLC. albicans[236]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Zikalala, N.E.; Azizi, S.; Zikalala, S.A.; Kamika, I.; Maaza, M.; Zinatizadeh, A.A.; Mokrani, T.; Kaviyarasu, K. An Evaluation of the Biocatalyst for the Synthesis and Application of Zinc Oxide Nanoparticles for Water Remediation—A Review. Catalysts 2022, 12, 1442. https://doi.org/10.3390/catal12111442

AMA Style

Zikalala NE, Azizi S, Zikalala SA, Kamika I, Maaza M, Zinatizadeh AA, Mokrani T, Kaviyarasu K. An Evaluation of the Biocatalyst for the Synthesis and Application of Zinc Oxide Nanoparticles for Water Remediation—A Review. Catalysts. 2022; 12(11):1442. https://doi.org/10.3390/catal12111442

Chicago/Turabian Style

Zikalala, Nkosingiphile E., Shohreh Azizi, Sithembela A. Zikalala, Ilunga Kamika, Malik Maaza, Ali Akbar Zinatizadeh, Touhami Mokrani, and Kasinathan Kaviyarasu. 2022. "An Evaluation of the Biocatalyst for the Synthesis and Application of Zinc Oxide Nanoparticles for Water Remediation—A Review" Catalysts 12, no. 11: 1442. https://doi.org/10.3390/catal12111442

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