Next Article in Journal
Skin Pigmentation Types, Causes and Treatment—A Review
Next Article in Special Issue
Wild Asparagus Shoots Constitute a Healthy Source of Bioactive Compounds
Previous Article in Journal
Preclinical EIS Study of the Inflammatory Response Evolution of Pure Titanium Implant in Hank’s Biological Solution
Previous Article in Special Issue
Identification of Polyphenolic Compounds Responsible for Antioxidant, Anti-Candida Activities and Nutritional Properties in Different Pistachio (Pistacia vera L.) Hull Cultivars
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Systematic Review

Biogenic Synthesis of Copper Nanoparticles: A Systematic Review of Their Features and Main Applications

by
Cristina M. Luque-Jacobo
1,
Andrea L. Cespedes-Loayza
1,
Talia S. Echegaray-Ugarte
1,
Jacqueline L. Cruz-Loayza
1,
Isemar Cruz
1,
Júlio Cesar de Carvalho
2 and
Luis Daniel Goyzueta-Mamani
1,3,*
1
Sustainable Innovative Biomaterials Department, Le Qara Research Center, Arequipa 04000, Peru
2
Bioprocess Engineering and Biotechnology Department, Federal University of Paraná—Polytechnic Center, Curitiba 81531-980, Brazil
3
Vicerrectorado de Investigación, Universidad Católica de Santa María, Urb. San José s/n-Umacollo, Arequipa 04000, Peru
*
Author to whom correspondence should be addressed.
Molecules 2023, 28(12), 4838; https://doi.org/10.3390/molecules28124838
Submission received: 26 May 2023 / Revised: 14 June 2023 / Accepted: 15 June 2023 / Published: 18 June 2023
(This article belongs to the Special Issue Biological Activity of Phenolics and Polyphenols in Nature Products)

Abstract

:
Nanotechnology is an innovative field of study that has made significant progress due to its potential versatility and wide range of applications, precisely because of the development of metal nanoparticles such as copper. Nanoparticles are bodies composed of a nanometric cluster of atoms (1–100 nm). Biogenic alternatives have replaced their chemical synthesis due to their environmental friendliness, dependability, sustainability, and low energy demand. This ecofriendly option has medical, pharmaceutical, food, and agricultural applications. When compared to their chemical counterparts, using biological agents, such as micro-organisms and plant extracts, as reducing and stabilizing agents has shown viability and acceptance. Therefore, it is a feasible alternative for rapid synthesis and scaling-up processes. Several research articles on the biogenic synthesis of copper nanoparticles have been published over the past decade. Still, none provided an organized, comprehensive overview of their properties and potential applications. Thus, this systematic review aims to assess research articles published over the past decade regarding the antioxidant, antitumor, antimicrobial, dye removal, and catalytic activities of biogenically synthesized copper nanoparticles using the scientific methodology of big data analytics. Plant extract and micro-organisms (bacteria and fungi) are addressed as biological agents. We intend to assist the scientific community in comprehending and locating helpful information for future research or application development.

Graphical Abstract

1. Introduction

During the past two decades, extensive research has focused on synthesizing and utilizing nanoparticles, which play an essential role in multiple areas, including feed and pharmaceuticals, due to their unique properties [1]. Copper nanoparticles (CuNPs) have exhibited extraordinary properties due to their versatile activity, such as antitumor, antimicrobial, antioxidant, dye removal, and catalytic degradation. However, their traditional synthesis (chemical method) has issues, such as prolonged technical work and toxic raw materials, e.g., hydrazine, N-dimethylformamide, and sodium borohydride [2,3], which are potentially carcinogenic compounds [4]. In addition, the traditional chemical synthesis may present environmental risks, such as dispersion and exposure, to bioactive residues, with soil and water ecotoxicity.
According to Santhroshkumar et al. (2018) [1], green synthesis is most effective for reducing and stabilizing metal ions. Biological synthesis offers numerous benefits and methods, such as minimizing time consumption while providing non-toxicity and cost-effectiveness. In accordance with these statements, micro-organisms are capable of intracellular and cellular biosynthesis [5]. Most micro-organisms secrete extracellular enzymes with functional groups that have an affinity for metals, whereas some phytochemicals can perform cellular reduction, providing a natural coating that improves the NP biogenic stability. The cellular wall of fungi is composed of chitin and glucan, which can carry out the complexation of heavy metals that form metallic nanoparticles [6,7]. Despite the advantages of micro-organism synthesis, such as ease of handling, high growth capacity, and low environmental toxicity, the industrialization of nanoparticles is affected by the possibility of culture contamination, lengthy procedures, and a lack of control over their size [8].
On the other hand, biological synthesis utilizing natural plant extracts, such as leaves, fruits, stems, roots, flowers, seeds, or substances produced by plants, such as latex, have been used as reducing, stabilizing, and topping agents. The production of nanoparticles occurs through the interaction of the metal salt and organic compounds present in the extract, including proteins, amino acids, organic acids, vitamins, and secondary metabolites (bioactive compounds), such as flavonoids, alkaloids, polyphenols, terpenoids, heterocyclic compounds, and polysaccharides, which can act as binders [9]. In addition, these organic compounds can help control and form homogenous shapes, such as spherical, linear, and cubic, with an average size of 10 to 100 nm without the addition of synthetic polymers and surfactants [10]. Therefore, plant extracts can provide a more cost-effective method for nanoparticle synthesis than micro-organism synthesis [11], as evidenced by CuNPs.
CuNPs biologically synthesized using plant extracts are in high demand due to their wide variety of industrial uses and applications due to their innovative and tunable properties, such as high surface area, excellent conductivity, chemical reactivity, stability, and oxidation. As previously discussed, these nanoparticles can be utilized as antimicrobial and antifungal agents, where copper oxide NPs are more effective against Gram-positive and harmful bacteria such as Escherichia coli, Staphylococcus aureus, Klebsiella pneumoniae, Salmonella typhi, and Bacillus subtilis [12,13]. On the other hand, the antioxidant activity is associated with various bioreductive groups on the surface of the CuNPs [14]. Concerning antitumoral issues, CuNPs have been studied for their anticancer properties, which can be attained via multiple routes (reactive oxygen species-ROS generation, apoptosis, and autophagy, among others) that demonstrate the CuO nanoparticles’ anticancer ability toward cancer cell types; additionally, the type of plant used also affects anticancer response [15]. CuNPs synthesized from plant extracts are an environmentally friendly alternative for obtaining photocatalysts. Dye removal via catalytic degradation is another critical application of these nanoparticles, where various nanomaterials, such as transition metal oxides (i.e., CuO and Cu2O) are used. The precise reaction mechanism by which biogenic CuNPs and their oxides remove and degrade dyes is unknown. Nonetheless, numerous dyes can be converted into colorless or less pigmented reduced derivates because of dye degradation.
The lack of an organized and comprehensive overview of the properties and potential applications of CuNPs demonstrates the need for the current study. However, a fragmented understanding of the synthesis of CuNPs and their primary applications leads to a variety of conclusions. Consequently, this study aims to assess the research articles published over the past decade on the different applications of biogenically synthesized CuNPs, including antioxidant, antitumor, antimicrobial, dye-scavenging, and catalytic activities.

2. Results and Discussions

As shown in Figure 1, the bibliographic search for the diagnosis of AD resulted in 4098 articles. After removing the duplicates, unrelated topics, and other parameters, a total of 192 publications were selected for complete analysis. The studies examined were related to antioxidant capacity (20.31%), antitumoral (21.88%), antibacterial (27.08%), and catalytic effects (30.73%). The minimum number of occurrences of keywords was set at 30 keywords, and co-occurrences were generated 29 times (Figure 1).

2.1. The Antioxidant Effect of CuNPs

The antioxidant properties of CuNPs synthesized from plant extracts have been demonstrated. Figure 2 reveals that a determining factor was the salt concentration, which, despite being variable, ranged from 1 to 100 mM in most cases (Figure 2C). According to the research, the antioxidant capacity depends on the concentration due to the same behavior observed in metallic salts, such as copper acetate, copper chloride, and copper nitrate. However, copper sulfate exceeds these values [16].
Furthermore, the percentage of the antioxidant activity of 16 plants was screened by analyzing the antioxidant effects of CuNPs in each of the papers reviewed: this quantity is due to the complete available information. The antioxidant activity of Koelreuteria paniculata seed extract (14.54%) was lower than that of Solanum nigrum leaf extract (90%) (Figure 2A). Despite these values, the half maximal inhibitory concentration (IC50) of Tinospora cordifolia extracts was higher (566 µg/mL) than those of Solanum nigrum leaf extracts and Borreria hispida (60 µg/mL and 0.6 µg/mL, respectively) (Figure 2B).
Table 1 presents an overview of different metallic salts and concentrations, with various plants and the component extract used in CuNPs. As a result, the part of the plant used for extract preparation, such as leaves, fruits, stems, roots, flowers, and seeds, is significant because of the relationship between the phytochemical characteristics and the total antioxidant capacity necessary for the reduction to synthesize CuNPs. Because phenolic components are considered the main contributors to plant extracts’ total non-enzymatic antioxidant capacity, a decrease in phenolic compounds most likely resulted in reduced radical scavenging capacity [17]. In some cases, such as with tree bark, a direct relationship between antioxidant activity and total polyphenolic content was observed [18]. It is important to note that the phytochemicals present in the extract can be used as reducing and stabilizing agents in the overall synthesis [14].
In contrast, the antioxidant effect underlies the inhibition of chain reactions, such as the breakdown of peroxides, the binding of transition metal ion catalysts, radical scavenging activity, and the inhibition of continued hydrogen abstraction [14]. The antioxidants present in plants can act as stabilizing agents during synthesis, preventing nanoparticle aggregation or clustering. These plant components with antioxidant activity can reduce oxidative stress during synthesis. Oxidative stress can lead to the formation of ROS, which can be detrimental to the synthesis and stability of nanoparticles [14].

2.2. The Antitumoral Effect of CuNPs

2.2.1. Anticancer Activity of CuNPs

In recent decades, radiotherapy, chemotherapy, and surgery have been the options with which to treat cancer. However, these standard methods have cost and usage restrictions. Thus, a natural, cheap, non-toxic, and side-effect-free treatment and prevention option is urgently needed. The Food and Drug Administration of the United States (FDA) recognizes metallic nanoparticles (i.e., iron, gold, zinc, titanium, and silver) as safe therapeutic compounds. Thus, CuNPs (374 USD/lb) have attracted considerable interest from researchers because they are cheaper than gold (1973 USD/t oz), silver (24.17 USD/t oz), and platinum (997 USD/t oz) [54]. Therefore, approaches involving CuNPs will be highly cost-effective [16].
CuNPs are promising cancer diagnosis and evaluation agents. Due to their unique properties, including their high surface-to-volume ratio, diffusion, efficiency of synthesis, and optical properties, they are effective against many cancer cell lines. These crucial factors are essential for extending the drug’s half-life and delivery application [55].
Depending on the source of CuNPs and the type of cell lines, CuNPs can act via various cytotoxic mechanisms, primarily ROS production, apoptosis, autophagy, and DNA damage. CuNPs interact actively with intracellular protein functional groups, nitrogen bases, and phosphate groups in DNA, causing cytotoxicity in tumor cell lines. Nanoparticles with anticancer properties are known for their potential ability to inhibit abnormally expressed signaling proteins, such as Akt and Ras, cytokine-based therapies, DNA- or protein-based vaccines against specific tumor markers, and tyrosine kinase inhibitors with a consistent antitumor effect [56].
In vitro research of various human cell lines, including neuronal cells, cardiac microvascular endothelial cells, kidney cells, liver cells, and lung epithelial cells, demonstrated that oxidative stress mediates the cytotoxicity of CuNPs. Thus, the excessive use and disposal of CuNPs increase their potential toxicity to the environment and human health. Therefore, the biocompatibility of synthesized CuNPs must be determined [57]. The high concentration of free radicals in normal cells causes numerous mutations in their DNA and RNA, accelerating the proliferation and growth of abnormal or cancerous cells.
Some reports have suggested that CuNPs could induce apoptosis in cancerous cells via ROS generation by modulating the uptake of P53 and Bax/Bcl-2. Previous research indicates that the mechanism of action involves the destruction of ROS generated during cancer cell proliferation and transported as radicals or free radicals [58]. Regarding the ROS species (·OH, ·O2, and H2O2), they play a crucial role in the death of eukaryotic cells induced by biogenic CuNPs. The highly reactive ·OH is a significant oxidant that influences oxidative DNA damage, including single- and double-strand breaks, mutation discovery, and the production of oxidized nucleotides [59].
The anticancer efficacy of CuNPs seems to depend on their size, morphology, specific surface area, increase in oxygen vacancies, reactant molecule diffusion ability, and release of Cu2+. When particle size decreases, particle surface area increases dramatically. It increases the potential number of ROS groups on the particle surface, which could substantially impact adverse biological effects [60]. In addition, small NPs offer a larger surface area to produce ROS, such as hydrogen peroxide, superanion radicals, and hydroxyl radicals, in cancer cells [57]. The smaller size of CuNPs may result in widespread tissue distribution, deeper penetration within specific tissues, improved cellular uptake, and enhanced cytotoxicity against cancer cells [61]. Research shows particles smaller than 50 nm exhibit more activity in the different cancer cell lines [28]. NPs with dimensions less than 200 nm exhibit efficient extravasation into leaky tumor vasculature and accumulation in tumor tissues due to their increased permeability and retention effect [62]. In vivo studies on the biodistribution and toxicity of CuNPs have revealed that smaller particle sizes exhibit greater transvascular and interstitial transport. In tumors, 50 nm NPs have demonstrated significantly greater permeability than 125 nm NPs [63]. Therefore, an efficient drug carrier must be small enough to leave the bloodstream, enter the vessels, and reach the tumor site [64].

2.2.2. Cytotoxic Effect of CuNPs on Cancer Cell Lines

The cytotoxic effects and biocompatibility of CuNPs depend on their concentration and synthesis routes. The synthesis of CuNPs has been reported mainly by physical methods, like ball milling, chemical methods, such as the sol-gel method, and biological methods by different plants or animal extracts. Among these routes, the biological routes, often called ‘green synthesis’, have proven to be one of the most biocompatible and ecofriendly methods for CuNP synthesis due to using eco-compatible reagents for the synthesis process [65]. Both the synthesis route and the inherent nature of CuNPs are important considerations for their effectiveness. The choice of synthesis route affects their specific characteristics, while their inherent nature determines their functional properties and behavior in specific applications.
Table 2 shows different plants that have been used as biological sources for the synthesis of CuNPs that have a cytotoxic effect on cancer cell lines.
CuNPs became of great interest due to their cytotoxicity for multiple types of cancer in a dose-dependent manner without affecting healthy cells, compared to chemically synthesized NPs, which cause cell death in both benign and cancerous cells.
(a)
CuNPs against breast cancer
The effects of CuNPs derived from Prunus nepalensis on MCF7 were studied by analyzing the expression of oncogenes (Ras, Myc) and tumor suppressor genes (P14/P19, P53, P21, and Caspase 3). The results demonstrated that CuNPs increased the expression of the genes involved in apoptosis in a dose-dependent manner. Furthermore, CuNPs induced apoptosis in MCF-7 cells by downregulating oncogenes and upregulating tumor suppressor genes [56]. Additionally, XRD studies on CuNPs synthesized from G. Sylvestre leaf extract showed a smaller crystal size with a higher surface area, which increased their anticancer activity. The cytotoxicity results showed that green CuNPs were more effective against MCF-7 cells than chemically synthesized CuNPs [60].
The cytotoxicity of CuNPs derived from Echinophora platyloba was observed in Raji and MCF-7 cells. The interaction of CuNPs with circulating tumor DNA (ct-DNA) showed an unusual binding mode that combines the characteristics of groove binding and intercalation, suggesting that the cytotoxicity of CuNPs may result from their interaction with DNA [66].
Additionally, the treatment of MDA-MB-231 cells with CuNPs resulted in distinct morphological changes, such as shrinkage, detachment, membrane blebbing, and distortion [61].
On the other hand, it uses an endophytic bacterium. A low concentration of green synthesized CuNPs (2–28 nm) has antiproliferative effects on breast cancer (T47D) cell lines [67].
(b)
CuNPs against cervical cancer
CuNPs were synthesized by using the aqueous leaf extracts of Azadirachta indica, Hibiscus rosa-sinensis, Murraya koenigii, Moringa oleifera, and Tamarindus indica and were tested against cervical cancer cells, revealing changes such as caspase activation, plasma membrane blebbing, binucleation, cytoplasmic vacuolation, cell shrinkage, nuclear fragmentation, chromatin condensation, chromosomal apoptotic bodies, and DNA fragmentation. The IC50 ranged between 20.32 to 26.73 µg/mL, with an average size of 12 nm [19]. Another study used black beans to produce NPs that caused cell death by activating apoptotic pathways initiated by intracellular ROS [68].
Similarly, CuNPs 40–45 nm in size were synthesized using Houttuynia cordata extract. The fluorescent staining analysis revealed the inhibition of cell proliferation and promotion of apoptotic cell death in HeLa cells targeting the PI3K/Akt signaling pathways at 5 and 7.5 µg/mL doses with an IC50 of 5 µg/mL [69].
(c)
Copper/copper oxide NPs against lung cancer
The cytotoxic effects of CuNPs were examined in A549 cells, where they induced apoptotic pathways and caused low cell viability with fragmented nuclei and loss of membrane integrity. Aqueous Ficus religiosa leaf extract was used to synthesize copper NPs, showing that, at higher concentrations (500 g/mL), cell viability was reduced by up to 6%. The presence of bioactive molecules in F. religiosa leaf extract may explain the improved cytotoxicity; in conclusion, copper oxide nanoparticles activate the apoptosis pathway via the generation of reactive oxygen species, and mitochondrial depolarization causes cell death via nuclear fragmentation [70].
Similarly, as previously stated, copper oxide NPs synthesized by the aqueous leaf extracts of A. indica, H. rosa-sinensis, M. koenigii, M. oleifera, and T. indica showed a cytotoxic effect on lung cancer cells. The induction of apoptosis was accompanied by membrane blebbing, cell shrinkage, caspase activation in the cytoplasm, and nuclear fragmentation [19].
(d)
Copper/copper oxide NPs against ovarian cancer
Green CuNPs synthesized from Camellia sinensis aqueous leaf extract showed high cell death and anti-human ovarian cancer properties against CAOV-3, SW-626, and SK-OV-3 cell lines. Even at high doses, the healthy cells were unharmed [53]. In another study, small CuNPs were obtained (6–15 nm) using Cressa leaf extract, inhibiting SKOV3 human ovarian carcinoma cell line growth with an IC 20 value of µg/mL [71].
Olea europaea leaf extracts had the highest inhibition yield (94%) at 50 g/L. CuNPs targeted AMJ-13 and SKOV-3 cell lines and induced cell shape changes, clumping, and cell communication inhibition [72].
(e)
CuNPs against other cancer cells lines
CuNPs synthesized from plant extracts showed promising results against other cancer cells. For instance, pumpkin seed extract was used to synthesize CuNPs against colorectal cancer cells (HCT-116) [73], in which the results showed significant apoptotic induction after treating the cells with 25 g/mL CuNP. CuNPs made from Allium noeanum extract against three endometrial cancer lines (Ishikawa, HEC-1-A, HEC-1-B, and KLE) showed that the percentage of cell viability decreases when increasing CuNP concentration [28]. In the case of melanoma (B16F10)-synthesized CuNPs from Quisqualis indica floral extract, the results indicate that CuNPs induced cytotoxicity via apoptosis involving LDH release, ROS generation, and GSH depletion in a dose-dependent manner [74].
In contrast, when CuNPs were synthesized using Vibrio sp. to research cytotoxicity activity in esophageal cancer cells (KYSE30), the viability of KYSE30 cells was significantly reduced over time [IC50 = 37.52 mg/L (24 h), IC50 = 18.26 mg/L (48 h), and IC50 = 13.96 mg/L (72 h)] [75]. In another study using the fresh biomass of the cyanobacteria Cylindrospermum stagnale, CuNPs were found to significantly increase the concentration in a time-dependent manner. At 24 h, increasing the particle concentration from 25 to 100 μg/mL increased cell viability inhibition from 45 to 66% [76].
As shown in Figure 3, the IC50 antitumor effect on the major types of cancer cell lines is demonstrated using biological sources, mainly plant extracts.
Additionally, in human bone marrow mesenchymal stem cells, studies on the effect of CuNPs have demonstrated a certain suppression of proliferation development due to the activation of apoptotic pathways [77]. Nonetheless, numerous other studies report a positive effect when applied to biomaterials, primarily due to osteogenic properties for the regeneration of bone tissue and cartilage, where an increase in porosity, mechanical strength, and cross-linking is observed in scaffolds. These results and those of a number of other authors are contradictory, indicating that the topic must be explored and studied further, as it appears that determining the appropriate forms and concentration of copper is essential for their safe and effective application [78,79,80].
Table 2. The antitumoral effect of biogenically synthetized CuNPs, salt, concentration, and biological source.
Table 2. The antitumoral effect of biogenically synthetized CuNPs, salt, concentration, and biological source.
PlantPartMetallic ConcentrationCell LineSize (nm)AssayIC50
(µg·mL−1)
Reference
Breast cancer
Copper sulfate
Syzygium alternifoliumStem bark5 mM(MDA-MB-231)5–13MTT50[81]
Olea europaeaLeaves2 mM(AMJ-13)20–50MTT, AE-EB1.47 [72]
Justicia glaucaLeaves0.1 M(MCF-7)19.72MTT28.72[82]
Dovyalis caffraLeaves1 mM(MCF-7)30–50MTT 4.04[58]
Punica granatumPeel1.0 M(MCF7)6MTT7.1[83]
Phoenix dactyliferaPits1.0 M(MCF7)20MTT45.7[83]
Prunus nepalensisFruit1 mM (MCF-7)35–50MTT158.5[56]
Cystoseira myricaAlgae1 mM (MCF-7)21MTT-[84]
Delonix regiaLeaves5 mM (MCF-7)69–108MTT3.77[85]
Acalypha indicaLeaves-(MCF-7)26–30MTT56.16[86]
Copper acetate
Cucurbita spp.Seed3 mM(MDA-MB-231)20MTT, AO-EB, ROS, MMP20[61]
Hibiscus rosa-sinensiLeaves0.2 M (MCF-7)12MTT, Hoechst 3325822.45 [19]
Murraya koenigiiLeaves0.2 M (MCF-7)12MTT, Hoechst 3325825.32[19]
Moringa oleiferaLeaves0.2 M(MCF-7)12MTT, Hoechst 3325826.71[19]
TamarindusindicaLeaves0.2 M (MCF-7)12MTT, Hoechst 3325819.77[19]
Brevibacillus brevisBiomass100 ul(T47D)2–28MTT122.3[67]
Camellia sinensisLeaves (MCF-7)22MTT50[87]
Azadirachta indicaLeaves0.2 M (MCF-7)12MTT, Hoechst 3325825.55[19]
Copper nitrate
Salacia reticulataLeaves1 mM(MCF-7)42MTT0.42[88]
Echinophora platyloba-0.1 M (MCF-7)10MTT, AO/BE21.44[66]
Cervical cancer
Copper sulfate
Houttuynia cordataPlant3 mM (HeLa)40–45MTT, AO/EtBr5[69]
Black beansBeans10 mM (HeLa)26SRB, ROS, MMD, Clonogenic survival-[68]
Brassica oleracea var acephalaLeaves1 mM(HeLa)60–100MTT119.0[89]
Carica papayaLeaves5 mM (HeLa)77MTT139.[90]
Mucuna pruriens utilisSeed2.5 gr(HeLa)28MTT22.48[91]
Copper acetate
Azadirachta indicaLeaves0.2 M (HeLa)12MTT, Hoechst 3325826.73[19]
Hibiscus rosa-sinensiLeaves0.2 M (HeLa)12MTT, Hoechst 3325821.63[19]
Murraya koenigiiLeaves0.2 M (HeLa)12MTT, Hoechst 3325823.22[19]
Moringa oleiferaLeaves0.2 M (HeLa)12MTT, Hoechst 3325830.08[19]
TamarindusindicaLeaves0.2 M (HeLa)12MTT, Hoechst 3325820.32[19]
Epithelioma
Copper acetate
Moringa oleiferaLeaves0.2 M (Hep-2)12MTT, Hoechst 3325829.58[19]
TamarindusindicaLeaves0.2 M (Hep-2)12MTT, Hoechst 3325821.66[19]
Azadirachta indicaLeaves0.2 M (Hep-2)12MTT, Hoechst 3325828.59[19]
Hibiscus rosa-sinensiLeaves0.2 M (Hep-2)12MTT, Hoechst 3325822.59[19]
Murraya koenigiiLeaves0.2 M (Hep-2)12MTT, Hoechst 3325825.59[19]
Hepatocellular carcinoma
Copper acetate
Eclipta prostrataLeaves3 mM(HepG2)23–57MTT-[20]
Curcuma longaRoot6 gr(HepG2)27MTT64.10[92]
Azadirachta indicaLeaves2 gr (HepG2)15–16MTT-[93]
Cylindrospermum stagnale (cyanobacteria)Biomass1 mM(HepG2) 12MTT-[76]
Copper sulfate
Cystoseira myricaBrown alga1 mM(HepG2)21MTT-[84]
Momordica cochinchinensisLeaves0.01 M(HepG2)56MTT, ROS75 [94]
Lung carcinoma
Copper acetate
Andrographis paniculataLeaves0.5 M(A549)23MTT14.76 [59]
TamarindusindicaLeaves0.2 M(A549)12MTT, Hoechst 3325818.11[19]
Murraya koenigiiLeaves0.2 M(A549)12MTT, Hoechst 3325825.05[19]
Moringa oleiferaLeaves0.2 M(A549)12MTT, Hoechst 3325834.37[19]
Azadirachta indicaLeaves0.2 M(A549)12MTT, Hoechst 3325826.03[19]
Hibiscus rosa-sinensiLeaves0.2 M(A549)12MTT, Hoechst 3325820.15[19]
Copper sulfate
Ficus religiosaLeaf5 mM (A549)577MTT, AO/BE200[70]
Delonix regiaLeaves5 mM (A549)69–108MTT4.70[85]
llex paraguariensistLeaves1 mM (A549)26–40MTT36.89[95]
Copper nitrate
Trichoderma asperellumCell-freeextract5 mM (A549)110WST-140.625[96]
Cinnamomum zelanicumLeaves0.3 M(NCI-H2126II, NCI-H1437.III, NCI-H1573.IV, NCI-H661)9–69MTT250, 348, 301, and 261[97]
Mussaenda frondosaLeaf, stem and callus-(A549)2–10MTT, Dual AO/EB 85.66–458.35 [98]
Calendula officinalisLeaves0.3 M(LC-2/ad, PC-14, HLC-1)19–39 MTT 328, 297 and 514 [55]
Colorectal cancer
Carica papayaLeavesCopper sulfate
5 mM
(HT-29)77MTT93.[90]
Ormocarpum cochinchinenseLeavesCopper chloride
0.003 M
(HCT-116)1–2MTT40[99]
Cucurbita maximaSeedCopper acetate
3 mM
(HCT-116)20MTT, AO/BE25[73]
Ovarian cancer
Olea europaeaLeavesCopper sulfate
2 mM
(SKOV-3)20–50MTT, AO-EB2.27 [72]
Camellia sinensisLeavesCopper chloride
1.7 gr
(CAOV-3, SW-626, and SK-OV-3)10–20 MTT263, 208 and 315[53]
Cressa spp.LeavesCopper sulfate(SKOV3)6–15 MTT20[71]
Other types of cancer
Brassica oleracea-Copper sulfate
5 mM
Prostate cancer
(PC-3)
4 --[100]
Echinophora platyloba-Copper nitrate
0.1 M
Raji Burkitt’s Lymphoma10MTT, AO/BE10.79[66]
Allium noeanumLeavesCopper nitrate
1 mM
Endometrial Cancer (Ishikawa, HEC-1-A, HEC-1-B,and KLE)10–12MTT357, 356, 331 and 411[28]
Vibrio spp.Bacteria Copper nitrateEsophageal Cancer (KYSE30)8MTT18.26 [75]
Rhus punjabensis--Leukemia
(HL-60)
31–36SRB1.82[101]
Citrus aurantifoliaEnzymeCopper sulfate
1 mM
Melanoma
(SK MEL 28)
4 MTT 56.6[102]
Quisqualis indicaFloralCopper acetate
5.0 mM
Melanoma
(B16F10)
39 MTT, LDH102 [74]
Arbustus unedoLeavesCopper sulfate
0.01 mM
Nasopharynx cancer (KB)30MTT-[103]
Aerva javanicaLeavesCopper chloride
4 mM
Neuroblastoma (Neuro2A)15–23 MTT -[57]
Rhus punjabensis--Prostate adenocarcinoma
(PC-3)
31–36SRB19.25[101]
MTT: 3-[4,5-dimethylthiazol-2-yl]-2,5 diphenyl tetrazolium bromide assay, LDH: Cytosolic lactate dehydrogenase assay, SRB: Sulforhodamine B assay, AO/BE: Acridine orange/ethidium bromide staining, WST-1: Water-soluble tetrazolium salt assay, ROS: total reactive oxygen species (ROS) assay, MMD: mitochondrial membrane depolarization assay, MMP: mitochondrial membrane potential, AO/EtBr: acridine orange/ethidium bromide assay.

2.3. The Antibacterial Effect of CuNPs

The antimicrobial effect of nanoparticles on bacteria is due to the effect caused at the cell wall level, primarily composed of the polymers of peptidoglycans, sugars, and amino acids, which, due to their porosity, facilitate the passage of nanoparticles [104]. This effect is also highly dependent on the type of bacteria: Gram-negative cell walls comprise a single layer of peptidoglycan, whereas Gram-positive cell walls comprise multiple layers [26]. Gram-negative bacteria are less resistant due to greater negative surface charge [105].
Gram-positive micro-organisms have walls composed of amines and carboxyl groups, which, when combined with aryls in the presence of copper, result in an amination reaction, and increased membrane permeability [106,107].
The inhibitory or bactericidal effect of CuNPs results precisely from the inhibition of cell membrane enzymes caused by the attraction between NPs and the membrane, thereby promoting the oxidation of NPs that are electrostatically attracted to membrane-based plasma reductases [108]. These ions enter through the lipid layers and move towards the cytosol, causing the production of oxygen species as O2, which leads to the formation of H2O2, which is responsible for the oxidation of proteins and lipids [109,110].
According to the bioinformatic results of Ul-Hamid et al., 2022 [111], the CUNP binds to the Ile14, Thr12, GLN95, PHE92, Tyr98, Thr46, and Thr121 residues of S. aureus dihydrofolate redrawn tRNA synthetase and dihydropteroate synthetase, consequently inhibiting the activity of these enzymes.
Particularly when nanoparticles are small and spherical, their size and shape could have potential as inhibitors [112]. Nanoparticles enhance antimicrobial activity and membrane performance by increasing the surface area [113]. After interacting with compounds, such as sulfur and phosphorus, the nanoparticles appear to have introduced reactive hydroxyl radicals capable of causing irreversible damage, oxidizing the proteins, and causing damage at the RNA and DNA levels, thereby altering and destroying the helical structure [114].
The phytochemical composition of the source (such as phenolic compounds and other antioxidants) that allows for the synthesis and stability of the nanoparticles will significantly impact these characteristics and effects [115]. CuNPs combined with plant extracts can further enhance the antimicrobial effect due to terpenoids (present in essential oils), phenolic compounds, tannins, flavonoids, and alkaloids that can cause ion transport disruptions and alter the activity of ion transport [116,117].
The destruction of proton efflux bombs results in the release of toxic metal ions, which affects the permeability of pathogen membranes and the respiratory system [118].
According to research by Selvan et al., 2018 [118], CuNPs have an anti-larvicidal effect on the Anopheles subpartus due to their accumulation in the alimentary and respiratory channels, which causes a rupture in the layers of tissues.
Alternatively, when compared to bacteria, fungi can be less sensitive to CuNPs due to the nature of their cell walls. Fungi have cell walls composed of polysaccharides, such as chitin (N-acetylglucosamine) and lipids, which provide stiffness and resistance to nanoparticles [119].
Table 3 includes columns for salt concentration, shape, and biological source, allowing for a comprehensive analysis of the antibacterial effect in different micro-organisms.

2.4. The Catalytic Effect of CuNPs

The catalytic effect depends on chemical production, elimination, and industrial catalytic process efficiency, selectivity, and yield. Selectivity reduces waste and impurities, making products safer and greener [170]. Copper oxide nanoparticles are reactive metal oxide semiconductors. Their high surface area facilitates catalytic, antimicrobial, and antifouling effects [171]. They are relevant in two areas: a reduction in components, characterized by the reduction or elimination of dyes and colorants, and the synthesis and catalysis of elements. Table 4 summarizes the screened data.
When a dye solution containing nanoparticles is exposed to sunlight, a free electron and a hole are generated from the nanoparticles; the electron interacts with oxygen to form superoxide free radicals, whereas the hole interacts with water to form hydroxyl ions. The dye is decolored by the formed superoxide free radicals and hydroxyl ions [172].
Each reaction or product has its pathway in component synthesis and catalysis. Quantum effects and a large surface-to-volume ratio give metal nanoparticles fascinating UV–visible, catalytic, and antibacterial properties [173]. The catalytic activity might be affected by nanoparticle size [42], shape, and exposed crystal planes [32]. CuNPs with a high surface area are popular due to their stability, cost, toxicity, manufacturability, and potential as catalysts [174].

2.4.1. Reducing and Capping Agents in the Synthesis of CuNPs

The reducing agent is an important component of nanoparticle synthesis. All research is included due to the ecofriendly approach, where the preference of agents derived from micro-organisms or plant species is critical. Thus, this review covers some of the most important nanoparticle-synthesizing plants. Most of the papers in this section use agricultural waste like calli [98], rhizomes [175,176], fruit hulls [177], aerial parts [178], beans [179], leaves, peels, flowers, juice, and peels [180]. There are also some byproducts, such as gum [181], and some species from another kingdom, such as the algae Cystoseira trinodis [42] and the bacteria Escherichia sp. SINT7 [182]. We can also highlight the application of plants like Plukenetia volubilis [183] and Moringa oleifera [184], which are widely cultivated in Peru and can be used for CuNP synthesis.

2.4.2. Factors That Affect the Synthesis and Catalysis of CuNPs

(a)
Particle size: Most of the papers reviewed have synthesized nanoparticles with a maximum size of 100 nm, noting that decreasing particle size will increase catalytic activity [185] and that the solvent concentration affects the particles’ size and shape [186]. Furthermore, it has been discovered that a larger extract volume is required to produce nanoparticles with a narrower size range [155];
(b)
Temperature: The temperature of the reaction is an essential factor to consider when synthesizing nanoparticles. Some of the papers reviewed emphasized the importance of a high temperature for synthesis;
(c)
pH: pH is essential in nanoparticle synthesis, ranging from 9 [155,187] to 10 [188] and to 12 [172,189].
Table 4. The catalytic application of biogenically synthetized CuNPs, salt, concentration, and biological source.
Table 4. The catalytic application of biogenically synthetized CuNPs, salt, concentration, and biological source.
Scientific NameSourceConcentrationSize
(nm)
Catalytic Application/ResultsReference
Compound Synthesis, Degradation/Yield (%)Compound Removal/Yield (%)
Copper nitrate
Carica papayaFruit peels0.004 M28.0Palm oil effluents degradation: 66%.ND[190]
Sacha inchiLeaves0.01 M8–32NDMB degradation: 78.90%.[155]
Camelia sinensis and Prunus africanaLeaves0.0005 M6–8NDMB degradation: 83–85%[191]
Aglaia elaeagnoideaFlowers0.001 M20–45Reduction of 4-NP to 4-AP: ~99%.MB degradation: yield ~99%.
CR degradation: yield ~99%.
[171]
Achyranthes aspera and Crotalaria verrucosaLeaves0.003 M10–20NDRhB degradation: 91–96%.[192]
Cordia sebestenaFlowers0.1 M20–35Synthesis of pyrimidinones: 98%.
DHPM derivatives synthesis: 98% (case VIII).
BTB dye degradation: ~99%[131]
Solanum nigrumLeaves0.1 M25NDMB degradation: 97%.[26]
Moringa oleiferaLeaves0.1 M28NDPararosaniline dye degradation: 96.4%.[184]
Passiflora edulisLeaves0.1 M60–65NDMB degradation: 93%.[131]
Citrofortunella microcarpaLeaves1 M54–68NDRhodamine degradation: 98.31%.[193]
P. emblicaLeaves26.7 M80NDAs (V) removal: 98.9%[194]
Musaenda frondosaCallus-2.2NDMB degradation: 88–97%[98]
Rauvolfia serpentinaLeaves-10–20NDTB dye degradation: ~99%[134]
Copper acetate
Syzygium jambos (L.) AlstonLeaves0.001 M7.62Ipso-Hydroxylation of aryl boronic acids:100% (case IV,XIV).ND[195]
I. tinctoriaLeaves0.001 M10–30Synthesis of 2-(benzo[d]thiazol-2-ylthio) benzoic acid: 65–80% (case II).ND[196]
Thymbra spicataLeaves0.001 M10–20Reaction of aniline with 4-chlorobromobenzene: 96%ND[197]
Stachys LavandulifoliaFlowers0.001 M20–35Thyoldibenzene synthesis: 96%.
Benzenethiol synthesis: 98%.
ND[198]
Salvia hispanicaLeaves0.1 M30Cycloaddition of alkyl halides: 93% for CuO NPs, 99% for Cu2O NPsMB degradation: ~99%
[43]
Ocimum tenuiflorumLeaves0.1 M6–18NDMO degradation: 96.4%.[87]
Mimosa pudicaLeaves0.1 M8Reduction of p-nitrophenol: ~99%ND[199]
Quercus infectoriaFruit0.2 M26NDBV 3 dye degradation: 86%[177]
Citrus limonJuice0.3 M5–20NDCr (VI) adsorption: 98.3 ± 0.6%[200]
Lantana camaraFlowers0.4 M13–28Aza Michael reaction: 80%.
Enamine production: 90% (case VI).
ND[189]
Andrographis paniculata-0.5 M23NDMB degradation: 98%.[59]
Psidium guajavaLeaves2 M2–6NDNB degradation: 97%.
RY160 degradation: 80%.
[201]
Copper chloride
Gundelia tournefortiiLeaves0.003 MNDHydration of cyanamides: 89%.
Reduction of 4-nitrophenol: ~99%
ND[202]
Tamarix galicaLeaves0.003 MVarious sizesN-arylation of triazoles: 92% (case XII)ND[174]
Anthemis nobilisFlowers0.003 M38.6A3 coupling reaction with piperidine: 89%ND[203]
Thymus vulgaris L.Leaves0.003 M30N-arylation of indoles: 97% (case V)ND[204]
Jatropha curcasLeaves0.003 M10NDMB catalysis: ~99%.[205]
Ageratum houstonianumLeaves0.003 M200NDCR degradation: ~99%[206]
Euphorbia esula L.Leaves0.005 M20–1104 nitrophenol reduction: ~99%.ND[207]
Ginkgo biloba L.Leaves0.005 M15–20Reaction of benzyl azide with phenyl acetylene: 98% (case X).ND[208]
Otostegia persicaLeaves0.005 MNDSynthesis of 1,2,3-triazoles: 93%ND[209]
Plantago asiaticaLeaves0.005 M7–35Aldehyde cyanation: 93% (case IX).ND[210]
Brassica oleraceae, Pisum sativum and Solanum tuberosumPeels0.01 M22–31NDMB degradation: 79–96%[211]
Coffea ArabicaBeans0.1 M5–8NDAB 10B reduction: ~99%,
MB and XO reduction: ~99%,
[179]
Camellia sinensisLeaves0.2 M10–20pyrano[2,3-d] pyrimidines synthesis: 90–98% (case XIII).ND[53]
Punica granatumSeeds1 M40–80NDMB degradation: 87.1%[212]
T. CordifoliaLeaves 90NDDirect green, Eosin and
Safranine and reactive dye: 90%
[213]
Copper sulfate
Convolvulus percicusLeaves0.001 M15–30Arylation for C−N and C−O coupling reactions: 92%.ND[157]
Odina wodierGum0.001 M60–100NDAcid blue degradation: 96%.[185]
Alchornea laxifloraLeaves0.001 M3.2Oxidative Desulphurization: 63.92% (case XI).ND[214]
Cystoseira trinodis (algae)Biomass0.001 M7–10NDMB degradation: 98%.[42]
Euphorbia maculataLeaves0.001 M18.02NDMB degradation: 96%,
CR degradation: 85%,
RhB degradation: 89%
[178]
Rosmarinus officinalisLeaves0.001 MNDNDMB degradation: 97.4%.[215]
Myrtus communisLeaves0.001 M554-nitrophenol (4-NP) reduction: ~99%.ND[188]
Duranta erectaFruits0.005 M70NDMO reduction: 96%,
CR reduction: 90.35%
[216]
C. epigaeusRhizomes0.005 M65–80NDMB degradation: 90%.[98]
Manilkara zapotaLeaves0.005 M18.79NDMV, MG, CBB degradation:
92.2%, 94.9%, 78.8%.
[172]
A. muricataLeaves0.005 M30–40NDRR120 and MO degradation:
90%, 95%
[217]
Escherichia sp. SINT7Biomass0.005 M30NDCR, MG, DB-1, RB-5 degradation: 97.0%, 90.5%, 88.4% and 83.6%[182]
Bergenia ciliataRhizomes0.005 M50NDMB and MR degradation:
92%, 85%.
[176]
Coccinia grandisFruits0.01 M40–50Reducing para-nitrophenol (PNP) to para-amino phenol (PAP): ~97%.ND[218]
Aloe veraLeaves0.01 M24–61NDMB degradation: ~99%.[186]
Coccinia grandisFlowers0.01 M40–50Reduction of 4-nitroaniline into amino compound: 97.9%.ND[219]
Pterolobium hexapetalumLeaves0.01 M10–50NDRB 5 degradation: 98%.[156]
Triticum aestivumSeeds0.01 M22Nitrophenol reduction: ~99%.ND[220]
Diospyros montanaLeaves0.01 M5.9–21.8NDMB degradation. ~99%.[221]
Cardiospermum halicacabumLeaves0.01 M14.9NDMB degradation: 93.6%.[222]
Mentha piperita,
Citrus sinensis
Leaves0.01 M150ND% Cd (II), Ni (II) and Pb (II) removal: 18%, 52.5%, 84%[223]
Impatiens balsaminaLeaves0.02 M5–10NDMB degradation: ~85%,
CR degradation: ~80%
[224]
Z. spina-christiFruit0.02 M9NDCV removal: 93.7%[225]
Citrus grandisPeel0.04 M22–27NDMR reduction: 96%.[187]
Elsholtzia blandaLeaves0.05 M32–49NDDegradation of CR dye: 74–94%[226]
Solanum lycopersicumLeaves0.1 M20–40NDCrystal 271 violet degradation: ~97%.[146]
Salvia hispanicaLeaves0.1 M35Cycloaddition of benzyl chloride: 99%MB degradation: ~99%[43]
Centella asiaticaLeaves0.1 M20–30NDMR and MO reduction: 98%,
PR reduction:99.62%
[227]
Aegle marmelosPeel0.1 M20NDMB degradation: ~99%.[180]
Clitoria TernateaFlowers0.1 M18NDCristal violet (CV), direct red DR degradation: 88.3%, 65%[228]
Citrus aurantifoliaLeaves0.3 M55NDRhB dye degradation: 91%[155]
Musa balbisianaPeel1 M50–904-nitrophenol to 4-aminophenol conversion: 96% (case I).ND[229]
Euphorbia maculataLeaves1 M18C-S cross-coupling reaction: 89% (case VII)ND[230]
Aloe veraLeaves-80–120NDCR reduction: 70–75%[231]
MB: methylene blue, CR: congo red, MB: methylene blue, MV: methyl violet, MG: malachite green, CBB: coomassie brilliant blue, NB: Nile blue, RY160: reactive yellow 160, MO: methyl orange, CV: crystal violet, DR: direct red, acid Blue 120, BTB: bromothymol blue, crystal 271 Violet, RB5: reactive black 5 dye, RhB: rhodamine B dye, NO: naphthol orange, MR: methyl red, RR: reactive red 120, DR: direct green, EY: eosin yellowish, TB: trypan blue.
In the process of the degradation and reduction of both dyes and organic compounds using CuNPs, nanoparticle and sample’ concentration, pH, time, and light are the parameters that must be considered in these processes. The best yields can be obtained using a higher nanoparticle concentration, a lower sample concentration [200], and a smaller nanoparticle size [185]. Alkaline pH improves the degradation or reduction process, and time in many studies is directly proportional to the light factor and the use of reducing agents.
(d)
Reaction time: The reaction time was screened in all synthesis and dye removal papers. Regarding synthesis, most reviewed articles carried out a reaction time of 1 to 100 min. Reaction times of up to 200 to 500 min were also observed (Figure 4A). Regarding dye removal applications, most articles reported a time range of 1 to 200 min, and reaction times of up to 200 to 800 min were also observed. In both applications, most studies were successful in under 120 min (Figure 4B);
(e)
Salt concentration frequency: The frequency of salt concentration was also screened through all the papers reviewed for catalytic application. The four main metallic salts synthesizing CuNPs were copper sulfate, copper acetate, copper nitrate, and copper chloride (Figure 4C). Copper sulfate was the most used salt, followed by copper chloride, copper nitrate, and copper acetate. The most frequently used concentration for copper sulfate was 0.01 M, copper acetate was 0.001 M, copper chloride was 0.003 M, and copper nitrate was 0.1 M. The concentration range for the most used salts, copper sulfate, and copper chloride ranged from 1 mM to 10 mM (Figure 4D).

2.4.3. Mechanism of Compound Reduction Using CuNPs as a Catalyst

The reduction process consists of the following steps. The initial step involved the adsorption of reactants to the surface of nanoparticles. The adsorption of BH4- onto synthesized CuNPs transfers surface hydrogen to the nanoparticles. The second step is desorption, which produces a product on the nanoparticles’ surface. Immediately after the final product (amino compounds) desorption, the metal surface is made available for the catalytic cycle [219]. The pH of the solution primarily determines the removal of heavy metals from contaminated water [223]. CuNPs are an excellent alternative to adsorption for removing Cr (VI) toxicity from water [200]. Organic compounds require a reducing agent, such as NaBH4 or CuNPs, for a faster reduction [199]. One study found that using NaBH4 as a reducing agent under alkaline conditions reduced 97% of para-nitrophenol in 14 min [218].
Arsenic (V) was reduced by 98% in 50 min at pH 8 [194]; it was concluded that if no reducing agents were used, the treatment time would be longer. Lead is another heavy metal with a significant reduction percentage, with a value of 89% at pH 6. Mahmoud’s study discovered relatively low levels of cadmium and nickel, which could be attributed to their use of neutral pH rather than alkaline conditions. CuNPs are an excellent nanoabsorbent for purifying heavy metal-contaminated water, and their regeneration and reuse should be researched further [223].
Case I. Reduction of aromatic compounds: There are four steps in this reaction: hydrogen absorption, aromatic nitro compound absorption to metal surfaces, electron transfer from BH4- to aromatic nitro compounds, and aromatic amino compound desorption. Both azobenzene intermediates and hydroxyl amine reduction pathways most likely accomplish nitroarene reduction. The NPs (on their active surface) are reduced by the hydrogen liberated during sodium borohydride decomposition [229].
Case II. Production of benzoic acids: In the presence of NPs, the reaction occurs between 2-mercaptobenzothiazole and 2-iodobenzoic acid. Later, under optimized conditions, 3-iodo-5-nitro-1-tosyl-1H-indole (1 equiv.) and 2-fluoro-4-iodopyridine (1 equiv.) were reacted to give optimal (65–68%) yields for the coupling products 3a-c. 2-(benzo[d]thiazol-2-ylthio) benzoic acid was the product [196].
Case III. Synthesis of tetrazole derivatives: One specific example denotes preparing 5-(4-Nitrophenyl)-1H-tetrazole. The disappearance of reactants indicates a cycloaddition reaction, which is treated with muriatic acid and extracted with ethyl acetate. Finally, the light-yellow 5-(4-Nitrophenyl)-1H-tetrazole was obtained [232].
Case IV. Reduction of 4-nitrophenol: Another example of the catalytic properties of nanoparticles is the reduction of 4-NP to 4-aminophenol (4-AP). The result demonstrates that the natrolite zeolite/Cu NPs are required to reduce 4-NP to 4-AP [233].
Case IV. Homocoupling reactions: The base-free homocoupling reaction was also studied for various boronic acids. The reaction conditions are compatible with aryl functional groups such as aldehyde, ether, methyl, and nitro. Surprisingly, the bromo and chloro groups do not change. Other boronic acid surrogates, such as phenylboronic acid neopentylglycol ester, were also found to be suitable homocoupling reaction substrates [234].
Case V. N-arylation of compounds: It should be noted that the reaction of indole with iodobenzene was considered as a model reaction to determine the best reaction conditions for the catalytic N-arylation process [170].
Case VI. Aza-Michael reaction: The Aza-Michael reaction is ideal for creating the carbon–nitrogen (C–N) bond and other types of molecular bonds [189].
Case VII. Heterocoupling reactions: In organic synthesis, the Ullmann C−C homocoupling, C−N, and C−O hetero coupling reactions are required to prepare the bi-aryl, di-aryl amine, and di-aryl ether structures [235,236]. It should be noted that the C-S coupling reaction is also being investigated for thioether synthesis in the absence of a ligand [230].
Case VIII. Biginelli reaction: The catalytic efficacy of Cu/CuO/Cu2O nanoparticles in the Biginelli reaction to synthesize the biologically active compound 3,4-dihydropyrimidinone (DHPM). DHPM was prepared by loading various catalysts for the condensation of benzaldehyde with urea and ethylacetoacetate [131].
Case IX. Cyanation of aldehydes: This reaction occurs in the presence of non-toxic K4Fe(CN)6 as the cyanide source. The results confirmed the high stability and absence of impurities in reducing Cu II ions to CuO [210].
Case X. Huisgen [3 + 2] cycloaddition: To optimize the reaction conditions, the Huisgen [3+2] cycloaddition of azides and alkynes under ligand-free conditions with benzyl azide and phenyl acetylene [208].
Case XI. Oxidation of oils: As an oxidant, H2O2 can convert model oils (DBT in n-heptane) into the corresponding sulfone or sulfoxide. Peracetic acid and a CuNP oxide intermediate, hydroox-ocuprate, are formed by the nucleophilic attack of hydrogen peroxide on acetic acid and CuNPs, respectively [214].
Case XII. Synthesis of 1,2,3-triazoles: In synthesizing 1,2,3-triazoles, the reaction of benzyl chloride, sodium azide, and phenyl acetylene is critical [237].
Case XIII. Synthesis of pyrimidines: Another application of NPs as catalysts is the synthesis of pyrano[2,3-d] pyrimidines. Aromatic aldehydes, methylene compounds, barbituric acid, and Cu2O NPs were combined with solvents [53].
Case XIV. Ipso-hydroxylation of arylboronic acids: A typical reaction involves the ipso-hydroxylation of aryl and hetero-aryl boronic acids by phenyl-boronic acid, CuNPs, and H2O2. The reaction products are typically extracted with diethyl ether after the reaction [238].

2.4.4. Mechanism of Degradation of Toxic Organic Dyes by Biogenically Synthesized CuNPs

Figure 5 is an example of the degradation mechanism of toxic organic dyes and a compilation of previously studied articles. In this instance, three dyes were considered: methylene blue, rhodamine, and congo red [216]. In most of the articles, the three dyes had similar concentrations of 10 mg/L plus 10 mg of CuNPs. This mixture was stirred constantly for 30 min (Figure 5A). After, the solution was exposed to UV light and stirred continuously for 60 min (Figure 5B). Following these procedures, the sample was analyzed using a UV-visible spectrophotometer (Figure 5C), determining the percentage of dye degradation. The synthesis of methylene blue was also analyzed (Figure 5D). CuNPs degrade most dyes more efficiently than silver nanoparticles, particularly crystal violet [177]. Several studies have examined the significance of stirring the dye solution with the CuNPs in the dark for a few minutes to achieve the equilibrium of the adsorption and desorption of the dye with the nanoparticle surface before exposing them to sunlight or UV light, which degrades them rapidly [187].
As mentioned above, CuNPs generate electron–hole pairs via photon absorption, which is the basis for explaining the preceding technique. Electrons created in the valence band migrate to the hole in the conduction band. These valence band holes combine with hydroxyl ions to produce hydroxyl radicals (•OH). Superoxide radicals form when conduction band electrons react with dissolved oxygen. Superoxide radicals react with water to increase the concentration of •OH [239]. Due to their high oxidizing potential, •OH radicals degrade organic pollutants effectively. Bonds are broken, rings are opened, and oxidation occurs in degradation. Due to the substitution of •OH, polyhydroxylated products are easily removed from the benzene ring under radical attack [240] and break down into less hazardous fragments, such as NO3, SO42−, CO2, and H2O [227].
Based on the presented data, it is strongly recommended to prioritize and intensify further research on the effect of plant extracts on the synthesis of CuNPs. Synthesis utilizing plant extracts is a sustainable and environmentally friendly method with promising applications. However, additional research is required to fully explore the potential of plant extracts since specific data on the proper effect or synergic interaction are explicitly reported. Plant extracts are a rich source of different bioactive compounds, and their synergistic interactions with CuNPs can enhance the activities under study. In addition, by examining the effect of plant extracts, specific antibacterial or antitumor properties can be identified, allowing for targeted applications against antibiotic-resistant strains and cancer cell lines. More research in this area will unlock the potential of plant extract-mediated synthesis, resulting in the development of effective, sustainable, and economically viable solutions for addressing bacterial infections and cancer cell line resistance.

3. Materials and Methods

3.1. Search Strategy

We systematically reviewed research articles published in English, excluding non-English publications, case reports, books, letters, and patents. We analyzed the research on CuNP biogenic or green synthesis and their primary applications.
We formulate the problem based on the premise that there is no systematic review of the principal applications of biogenically synthesized CuNPs.
A bibliographic search was conducted from 2012 to date, using the PubMed, Web of Science (WOS) Core Collection, and SCOPUS databases (September 2014 to January 2023). The results were deduplicated and uploaded to EndNote (Clarivate Analytics) and Rayyan Software. The preferred reporting items for systematic reviews and meta-analyses (PRISMA) recommendations were followed. The protocol for this systematic review was registered on the International Platform of Registered Systematic Review and Meta-analysis Protocols (INPLASY) (INPLASY202350109) and is available in full at inplasy.com (https://inplasy.com/inplasy-2023–5-0109/, accessed on 15 June 2023). The systematic review has been elaborated according to the PRISMA 2020 checklist (Table S5).

3.2. Search Criteria

We carried a screening set for the biogenic or green synthesis of CuNPs in the databases mentioned above, using the keywords in the title and abstract: (nanoparticle*) AND (biogenic) OR (green AND synthesis) AND (copper) OR (cu) OR (cu2o) OR (cuo).

3.3. Study Selection, Data Extraction, and Quality Assessment

Four authors selected the studies and reviewed the titles and abstracts of all published articles using the Rayyan software, based on the selected criteria and keywords. To compile research articles on biogenic synthesis, the patents, clinical trials, reviews, duplicates, and in vivo test-related studies and articles were excluded.
CMLJ, ACL, TEU, JCDC, and LDGM carried out the review process and management by extracting the authors’ and content’s raw data in accordance with a standard procedure.
After the initial database data collection, we screened and categorized the major applications using four screening sets. The first set contained antitumor terms, the second set contained antioxidant terms, the third set contained antibacterial terms, and the fourth set contained catalytic effect and dye removal terms.
The quality and risk of bias were evaluated based on the Cochrane Handbook for Systematic Interventions recommendations (http://www.handbook.cochrane.org, accessed on 12 December 2023). Independently, the authors evaluated the possibility of bias during the review process. At any stage of the reviewing process, disagreements among the authors were discussed and resolved.
The keywords for inclusion in the Rayyan Software were
  • For the antitumoral analysis, the following keywords in the title and abstract were used in the research fields: Green synthesis, Biogenic, Cancer, Anticarcinogenic, Antitumoral, and Cytotoxic; they were also used as keywords to collect data that might be under one of these terms;
  • For the antioxidant analysis, the following keywords in the title and abstract were used in the research fields: Green synthesis, Biogenic, Antioxidant, and Oxidative, and these were also used to collect data that might fall under one of these terms;
  • For the antibacterial analysis, the following keywords in the title and abstract were used in the research fields: Nanoparticles, Copper, Antibacterial, Green synthesis, antimicrobial, and biogenic, which were also used as keywords to collect data that might fall under one of these terms;
  • For the catalytic effect and dye removal analysis, the following keywords in the title and abstract were used in the research fields: Green synthesis, Catalytic, Photocatalytic, and Biogenic, which were also used as keywords to collect data that might fall under one of these terms.
A MeaSurement Tool to Assess Systematic Reviews (AMSTAR) was used to assess and evaluate the validity, quality, and reliability of the systematic review, as the content of each article was evaluated by following a set of criteria. The 11-item eligibility instrument was used to determine these scores, with each item in each article receiving a score of 1 or 0: Can’t Answer (CA) and Not Applicable (NA). The lowest score was 5, and the highest score was 8. (Table 5). The quality of research is proportional to the score: a score between 8 and 11: good quality; a score between 4 and 7: moderate quality; a score between 0 and 3: lower quality [38].

3.4. Data Analysis

During the systematic review, we collected qualitative data. Using the free VOSviewer server software (Version 1.6.15; https://www.vosviewer.com, accessed on 23 January 2023), we compiled, analyzed, and plotted the publication datasets, observing the trends and the most studied fields and applications on the biogenic synthesis of CuNPs. We established a threshold of 30 keyword co-occurrences. These co-occurrences were words that appeared in the analysis more than 29 times. The strategy process is summarized in Figure 6. It was unnecessary to conduct a meta-analysis because the current study did not reveal any discrepant findings for CuO.

4. Conclusions

The biogenic synthesis of CuNPs is a promising method for producing functional and stable nanoparticles with numerous potential applications. Biogenic synthesis has several advantages over conventional chemical synthesis methods, including low toxicity, cost-effectiveness, and mild reaction conditions. In addition, these nanoparticles have antioxidant, antibacterial, and antitumor properties, making them useful for healthcare and medical applications.
Biologically synthesized CuNPs have demonstrated catalytic potential in a variety of chemical reactions due to their large surface area and unique surface properties. As effective catalysts, they can enhance reaction rates and selectivity, and their production can support environmentally friendly and sustainable chemical industry processes. Due to their antioxidant and antitumor properties, the biogenic synthesis of CuNPs has shown great promise in medicine as antimicrobial agents and for cancer treatment and prevention.
The potential applications of biogenic CuNPs, which include antibacterial, antitumor, antioxidant, and catalytic properties, make them an exciting study area. However, there are still challenges that need to be overcome for the further advancement and utilization of CuNPs, such as optimizing the synthesis parameters, tailoring the properties for specific applications, ensuring biodegradability, achieving scalability, understanding the mechanisms of action, addressing safety concerns, and integrating CuNPs into emerging technologies. By overcoming these challenges through research, the full potential of CuNPs will be unlocked and expand their applications in diverse fields.
The integration of CuNPs with emerging technologies, such as nanomedicine and nanoelectronics, can create new diagnostics, therapeutics, and material applications. Collaboration between researchers, regulatory bodies, and industry stakeholders will be essential in establishing safety guidelines, standardizing protocols, and translating the potential of biogenic CuNPs into real-world applications, which will ultimately benefit sectors such as healthcare, catalysis, energy, and environmental remediation.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/molecules28124838/s1, Table S1. AMSTAR table for the antioxidant effect analysis., Table S2. AMSTAR table for the antitumoral effect analysis., Table S3. AMSTAR table for the antibacterial effect analysis., Table S4. AMSTAR table for the catalytic/dye removal effect analysis., Table S5. PRISMA 2020 checklist. References [241,242] can be found in the Supplementary Materials.

Author Contributions

Revision and conceptualization, L.D.G.-M., A.L.C.-L., C.M.L.-J., J.C.d.C. and T.S.E.-U.; methodology, L.D.G.-M. and A.L.C.-L., C.M.L.-J.; formal analysis, L.D.G.-M., A.L.C.-L., C.M.L.-J. and T.S.E.-U.; writing—original draft preparation, L.D.G.-M., A.L.C.-L., C.M.L.-J. and T.S.E.-U.; Funding acquisition, J.L.C.-L. and I.C.; visualization, J.L.C.-L.; resources, I.C. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the program PROCIENCIA from CONCYTEC grant number 108-2021-FONDECYT Esquema E041-2021-02.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

We would like to express our sincere gratitude to Milagros Monroy-Negreiros and Helar Gómez-Chalco for their valuable contributions and insights into this work.

Conflicts of Interest

The authors declare no conflict of interest.

Sample Availability

Not applicable.

References

  1. Santhoshkumar, J.; Agarwal, H.; Menon, S.; Rajeshkumar, S.; Venkat Kumar, S. A Biological Synthesis of Copper Nanoparticles and Its Potential Applications. In Green Synthesis, Characterization and Applications of Nanoparticles; School of Biosciences and Technology, Vellore Institute of Technology: Vellore, India, 2018; pp. 199–221. [Google Scholar]
  2. Nourbakhsh, S.; Habibi, S.; Rahimzadeh, M. Copper Nano-Particles for Antibacterial Properties of Wrinkle Resistant Cotton Fabric. In Proceedings of the Materials Today: Proceedings; Elsevier Ltd.: Amsterdam, The Netherlands, 2017; Volume 4, pp. 7032–7037. [Google Scholar]
  3. Tiwari, M.; Jain, P.; Chandrashekhar Hariharapura, R.; Narayanan, K.; Bhat, K.U.; Udupa, N.; Rao, J.V. Biosynthesis of Copper Nanoparticles Using Copper-Resistant Bacillus Cereus, a Soil Isolate. Process. Biochem. 2016, 51, 1348–1356. [Google Scholar] [CrossRef]
  4. Abdelbasir, S.M.; McCourt, K.M.; Lee, C.M.; Vanegas, D.C. Waste-Derived Nanoparticles: Synthesis Approaches, Environmental Applications, and Sustainability Considerations. Front. Chem. 2020, 8, 782. [Google Scholar] [CrossRef] [PubMed]
  5. Frejo, M.; Díaz, M.; Lobo, M.; García, J.; Capó, M. Nanotoxicología Ambiental: Retos Actuales. Med. Balear. 2011, 26, 36–46. [Google Scholar]
  6. 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] [PubMed]
  7. Kalimuthu, K.; Suresh Babu, R.; Venkataraman, D.; Bilal, M.; Gurunathan, S. Biosynthesis of Silver Nanocrystals by Bacillus Licheniformis. Colloids Surf. B Biointerfaces 2008, 65, 150–153. [Google Scholar] [CrossRef] [PubMed]
  8. Ahmad, S.; Munir, S.; Zeb, N.; Ullah, A.; Khan, B.; Ali, J.; Bilal, M.; Omer, M.; Alamzeb, M.; Salman, S.M.; et al. Green Nanotechnology: A Review on Green Synthesis of Silver Nanoparticles—An Ecofriendly Approach. Int. J. Nanomed. 2019, 14, 5087–5107. [Google Scholar] [CrossRef] [Green Version]
  9. Jadoun, S.; Arif, R.; Jangid, N.K.; Meena, R.K. Green Synthesis of Nanoparticles Using Plant Extracts: A Review. Environ. Chem. Lett. 2021, 19, 355–374. [Google Scholar] [CrossRef]
  10. Rajesh, K.M.; Ajitha, B.; Reddy, Y.A.K.; Suneetha, Y.; Reddy, P.S. Assisted Green Synthesis of Copper Nanoparticles Using Syzygium Aromaticum Bud Extract: Physical, Optical and Antimicrobial Properties. Optik 2018, 154, 593–600. [Google Scholar] [CrossRef]
  11. Patel, B.H.; Channiwala, M.Z.; Chaudhari, S.B.; Mandot, A.A. Biosynthesis of Copper Nanoparticles; Its Characterization and Efficacy against Human Pathogenic Bacterium. J. Environ. Chem. Eng. 2016, 4, 2163–2169. [Google Scholar] [CrossRef]
  12. Chen, C.W.; Hsu, C.Y.; Lai, S.M.; Syu, W.J.; Wang, T.Y.; Lai, P.S. Metal Nanobullets for Multidrug Resistant Bacteria and Biofilms. Adv. Drug. Deliv. Rev. 2014, 78, 88–104. [Google Scholar] [CrossRef]
  13. Zazo, H.; Colino, C.I.; Lanao, J.M. Current Applications of Nanoparticles in Infectious Diseases. J. Control. Release 2016, 224, 86–102. [Google Scholar] [CrossRef]
  14. Rajeshkumar, S.; Menon, S.; Venkat Kumar, S.; Tambuwala, M.M.; Bakshi, H.A.; Mehta, M.; Satija, S.; Gupta, G.; Chellappan, D.K.; Thangavelu, L.; et al. Antibacterial and Antioxidant Potential of Biosynthesized Copper Nanoparticles Mediated through Cissus Arnotiana Plant Extract. J. Photochem. Photobiol. B 2019, 197, 111531. [Google Scholar] [CrossRef] [PubMed]
  15. Vincent, J.; Lau, K.S.; Evyan, Y.C.Y.; Chin, S.X.; Sillanpää, M.; Chia, C.H. Biogenic Synthesis of Copper-Based Nanomaterials Using Plant Extracts and Their Applications: Current and Future Directions. Nanomaterials 2022, 12, 3312. [Google Scholar] [CrossRef]
  16. Rajeshkumar, S.; Rinitha, G. Nanostructural Characterization of Antimicrobial and Antioxidant Copper Nanoparticles Synthesized Using Novel Persea Americana Seeds. OpenNano 2018, 3, 18–27. [Google Scholar] [CrossRef]
  17. Mahjouri, S.; Movafeghi, A.; Divband, B.; Kosari-Nasab, M. Toxicity Impacts of Chemically and Biologically Synthesized CuO Nanoparticles on Cell Suspension Cultures of Nicotiana Tabacum. Plant. Cell. Tissue Organ. Cult. 2018, 135, 223–234. [Google Scholar] [CrossRef]
  18. Alavi, M.; Karimi, N.; Valadbeigi, T. Antibacterial, Antibiofilm, Antiquorum Sensing, Antimotility, and Antioxidant Activities of Green Fabricated Ag, Cu, TiO2, ZnO, and Fe3O4 NPs via Protoparmeliopsis Muralis Lichen Aqueous Extract against Multi-Drug-Resistant Bacteria. ACS Biomater. Sci. Eng. 2019, 5, 4228–4243. [Google Scholar] [CrossRef]
  19. Rehana, D.; Mahendiran, D.; Kumar, R.S.; Rahiman, A.K. Evaluation of Antioxidant and Anticancer Activity of Copper Oxide Nanoparticles Synthesized Using Medicinally Important Plant Extracts. Biomed. Pharmacother. 2017, 89, 1067–1077. [Google Scholar] [CrossRef] [PubMed]
  20. Chung, I.; Rahuman, A.A.; Marimuthu, S.; Kirthi, A.V.; Anbarasan, K.; Padmini, P.; Rajakumar, G. Green Synthesis of Copper Nanoparticles Using Eclipta Prostrata Leaves Extract and Their Antioxidant and Cytotoxic Activities. Exp. Ther. Med. 2017, 14, 18–24. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  21. Ali, J.S.; Mannan, A.; Nasrullah, M.; Ishtiaq, H.; Naz, S.; Zia, M. Antimicrobial, Antioxidative, and Cytotoxic Properties of Monotheca Buxifolia Assisted Synthesized Metal and Metal Oxide Nanoparticles. Inorg. Nano-Met. Chem. 2020, 50, 770–782. [Google Scholar] [CrossRef]
  22. Venugopalan, R.; Pitchai, S.; Devarayan, K.; Swaminathan, V.C. Biogenic Synthesis of Copper Nanoparticles Using Borreria Hispida (Linn.) Extract and Its Antioxidant Activity. Mater. Today Proc. 2020, 33, 4023–4025. [Google Scholar] [CrossRef]
  23. Thakar, M.A.; Saurabh Jha, S.; Phasinam, K.; Manne, R.; Qureshi, Y.; Hari Babu, V. V X Ray Diffraction (XRD) Analysis and Evaluation of Antioxidant Activity of Copper Oxide Nanoparticles Synthesized from Leaf Extract of Cissus Vitiginea. Mater. Today Proc. 2021, 51, 319–324. [Google Scholar] [CrossRef]
  24. Hasheminya, S.-M.; Dehghannya, J. Green Synthesis and Characterization of Copper Nanoparticles Using Eryngium Caucasicum Trautv Aqueous Extracts and Its Antioxidant and Antimicrobial Properties. Part. Sci. Technol. 2020, 38, 1019–1026. [Google Scholar] [CrossRef]
  25. Asemani, M.; Anarjan, N. Green Synthesis of Copper Oxide Nanoparticles Using Juglans Regia Leaf Extract and Assessment of Their Physico-Chemical and Biological Properties. Green. Process. Synth. 2019, 8, 557–567. [Google Scholar] [CrossRef]
  26. Muthuvel, A.; Jothibas, M.; Manoharan, C. Synthesis of Copper Oxide Nanoparticles by Chemical and Biogenic Methods: Photocatalytic Degradation and in Vitro Antioxidant Activity. Nanotechnol. Environ. Eng. 2020, 5, 14. [Google Scholar] [CrossRef]
  27. Dobrucka, R. Antioxidant and Catalytic Activity of Biosynthesized CuO Nanoparticles Using Extract of Galeopsidis Herba. J. Inorg. Organomet. Polym. Mater. 2018, 28, 812–819. [Google Scholar] [CrossRef] [Green Version]
  28. Chinnathambi, A.; Awad Alahmadi, T.; Ali Alharbi, S. Biogenesis of Copper Nanoparticles (Cu-NPs) Using Leaf Extract of Allium Noeanum, Antioxidant and in-Vitro Cytotoxicity. Artif. Cells Nanomed. Biotechnol. 2021, 49, 500–510. [Google Scholar] [CrossRef]
  29. Ijaz, F.; Shahid, S.; Khan, S.A.; Ahmad, W.; Zaman, S. Green Synthesis of Copper Oxide Nanoparticles Using Abutilon Indicum Leaf Extract: Antimicrobial, Antioxidant and Photocatalytic Dye Degradation Activities. Trop. J. Pharm. Res. 2017, 16, 743–753. [Google Scholar] [CrossRef] [Green Version]
  30. Velsankar, K.; Aswin Kumara, R.M.; Preethi, R.; Muthulakshmi, V.; Sudhahar, S. Green Synthesis of CuO Nanoparticles via Allium Sativum Extract and Its Characterizations on Antimicrobial, Antioxidant, Antilarvicidal Activities. J. Environ. Chem. Eng. 2020, 8, 104123. [Google Scholar] [CrossRef]
  31. Velsankar, K.; Vinothini, V.; Sudhahar, S.; Kumar, M.K.; Mohandas, S. Green Synthesis of CuO Nanoparticles via Plectranthus Amboinicus Leaves Extract with Its Characterization on Structural, Morphological, and Biological Properties. Appl. Nanosci. 2020, 10, 3953–3971. [Google Scholar] [CrossRef]
  32. Udayabhanu; Nethravathi, P.C.; Pavan Kumar, M.A.; Suresh, D.; Lingaraju, K.; Rajanaika, H.; Nagabhushana, H.; Sharma, S.C. Tinospora Cordifolia Mediated Facile Green Synthesis of Cupric Oxide Nanoparticles and Their Photocatalytic, Antioxidant and Antibacterial Properties. Mater. Sci. Semicond. Process. 2015, 33, 81–88. [Google Scholar] [CrossRef]
  33. Abbasifar, A.; Shahrabadi, F.; ValizadehKaji, B. Effects of Green Synthesized Zinc and Copper Nano-Fertilizers on the Morphological and Biochemical Attributes of Basil Plant. J. Plant. Nutr. 2020, 43, 1104–1118. [Google Scholar] [CrossRef]
  34. Rajeshkumar, S.; Nandhini, N.T.; Manjunath, K.; Sivaperumal, P.; Krishna Prasad, G.; Alotaibi, S.S.; Roopan, S.M. Environment Friendly Synthesis Copper Oxide Nanoparticles and Its Antioxidant, Antibacterial Activities Using Seaweed (Sargassum Longifolium) Extract. J. Mol. Struct. 2021, 1242, 130724. [Google Scholar] [CrossRef]
  35. Erci, F.; Cakir-Koc, R.; Yontem, M.; Torlak, E. Synthesis of Biologically Active Copper Oxide Nanoparticles as Promising Novel Antibacterial-Antibiofilm Agents. Prep. Biochem. Biotechnol. 2020, 50, 538–548. [Google Scholar] [CrossRef]
  36. Sepasgozar, S.M.E.; Mohseni, S.; Feizyzadeh, B.; Morsali, A. Green Synthesis of Zinc Oxide and Copper Oxide Nanoparticles Using Achillea Nobilis Extract and Evaluating Their Antioxidant and Antibacterial Properties. Bull. Mater. Sci. 2021, 44, 129. [Google Scholar] [CrossRef]
  37. Dashtizadeh, Z.; Jookar Kashi, F.; Ashrafi, M. Phytosynthesis of Copper Nanoparticles Using Prunus mahaleb L. and Its Biological Activity. Mater. Today Commun. 2021, 27, 102456. [Google Scholar] [CrossRef]
  38. Wu, S.; Rajeshkumar, S.; Madasamy, M.; Mahendran, V. Green Synthesis of Copper Nanoparticles Using Cissus Vitiginea and Its Antioxidant and Antibacterial Activity against Urinary Tract Infection Pathogens. Artif. Cells Nanomed. Biotechnol. 2020, 48, 1153–1158. [Google Scholar] [CrossRef]
  39. Rani, H.; Singh, S.P.; Yadav, T.P.; Khan, M.S.; Ansari, M.I.; Singh, A.K. In-Vitro Catalytic, Antimicrobial and Antioxidant Activities of Bioengineered Copper Quantum Dots Using Mangifera indica (L.) Leaf Extract. Mater. Chem. Phys. 2020, 239, 122052. [Google Scholar] [CrossRef]
  40. Zangeneh, M.M.; Ghaneialvar, H.; Akbaribazm, M.; Ghanimatdan, M.; Abbasi, N.; Goorani, S.; Pirabbasi, E.; Zangeneh, A. Novel Synthesis of Falcaria Vulgaris Leaf Extract Conjugated Copper Nanoparticles with Potent Cytotoxicity, Antioxidant, Antifungal, Antibacterial, and Cutaneous Wound Healing Activities under in Vitro and in Vivo Condition. J. Photochem. Photobiol. B 2019, 197, 111556. [Google Scholar] [CrossRef] [PubMed]
  41. Jadhav, M.S.; Kulkarni, S.; Raikar, P.; Barretto, D.A.; Vootla, S.K.; Raikar, U.S. Green Biosynthesis of CuO & Ag-CuO Nanoparticles from Malus Domestica Leaf Extract and Evaluation of Antibacterial, Antioxidant and DNA Cleavage Activities. New. J. Chem. 2018, 42, 204–213. [Google Scholar] [CrossRef]
  42. Gu, H.; Chen, X.; Chen, F.; Zhou, X.; Parsaee, Z. Ultrasound-Assisted Biosynthesis of CuO-NPs Using Brown Alga Cystoseira Trinodis: Characterization, Photocatalytic AOP, DPPH Scavenging and Antibacterial Investigations. Ultrason. Sonochem 2018, 41, 109–119. [Google Scholar] [CrossRef]
  43. Ghadiri, A.M.; Rabiee, N.; Bagherzadeh, M.; Kiani, M.; Fatahi, Y.; Di Bartolomeo, A.; Dinarvand, R.; Webster, T.J. Green Synthesis of CuO- And Cu2O-NPs in Assistance with High-Gravity- And Flowering of Nanobiotechnology. Nanotechnology 2020, 31, 425101. [Google Scholar] [CrossRef]
  44. Pandit, R.; Gaikwad, S.; Rai, M. Biogenic Fabrication of CuNPs, Cu Bioconjugates and in Vitro Assessment of Antimicrobial and Antioxidant Activity. IET Nanobiotechnol. 2017, 11, 568–575. [Google Scholar] [CrossRef] [PubMed]
  45. Hayat, K.; Ali, S.; Ullah, S.; Fu, Y.; Hussain, M. Green Synthesized Silver and Copper Nanoparticles Induced Changes in Biomass Parameters, Secondary Metabolites Production, and Antioxidant Activity in Callus Cultures of Artemisia absinthium L. Green. Process. Synth. 2021, 10, 61–72. [Google Scholar] [CrossRef]
  46. Alavi, M.; Karimi, N. Characterization, Antibacterial, Total Antioxidant, Scavenging, Reducing Power and Ion Chelating Activities of Green Synthesized Silver, Copper and Titanium Dioxide Nanoparticles Using Artemisia Haussknechtii Leaf Extract. Artif. Cells Nanomed. Biotechnol. 2018, 46, 2066–2081. [Google Scholar] [CrossRef] [Green Version]
  47. Hassan, S.E.-D.; Fouda, A.; Radwan, A.A.; Salem, S.S.; Barghoth, M.G.; Awad, M.A.; Abdo, A.M.; El-Gamal, M.S. Endophytic Actinomycetes Streptomyces Spp Mediated Biosynthesis of Copper Oxide Nanoparticles as a Promising Tool for Biotechnological Applications. J. Biol. Inorg. Chem. 2019, 24, 377–393. [Google Scholar] [CrossRef] [PubMed]
  48. Bulut Kocabas, B.; Attar, A.; Peksel, A.; Altikatoglu Yapaoz, M. Phytosynthesis of CuONPs via Laurus Nobilis: Determination of Antioxidant Content, Antibacterial Activity, and Dye Decolorization Potential. Biotechnol. Appl. Biochem. 2021, 68, 889–895. [Google Scholar] [CrossRef] [PubMed]
  49. Lung, I.; Opriş, O.; Soran, M.-L.; Culicov, O.; Ciorîță, A.; Stegarescu, A.; Zinicovscaia, I.; Yushin, N.; Vergel, K.; Kacso, I.; et al. The Impact Assessment of Cuo Nanoparticles on the Composition and Ultrastructure of Triticum aestivum L. Int. J. Environ. Res. Public. Health 2021, 18, 6739. [Google Scholar] [CrossRef] [PubMed]
  50. El-Batal, A.I.; Al-Hazmi, N.E.; Mosallam, F.M.; El-Sayyad, G.S. Biogenic Synthesis of Copper Nanoparticles by Natural Polysaccharides and Pleurotus Ostreatus Fermented Fenugreek Using Gamma Rays with Antioxidant and Antimicrobial Potential towards Some Wound Pathogens. Microb. Pathog. 2018, 118, 159–169. [Google Scholar] [CrossRef]
  51. Kalia, A.; Kaur, M.; Shami, A.; Jawandha, S.K.; Alghuthaymi, M.A.; Thakur, A.; Abd-Elsalam, K.A. Nettle-Leaf Extract Derived Zno/Cuo Nanoparticle-Biopolymer-Based Antioxidant and Antimicrobial Nanocomposite Packaging Films and Their Impact on Extending the Post-Harvest Shelf Life of Guava Fruit. Biomolecules 2021, 11, 224. [Google Scholar] [CrossRef]
  52. Jasrotia, T.; Chaudhary, S.; Kaushik, A.; Kumar, R.; Chaudhary, G.R. Green Chemistry-Assisted Synthesis of Biocompatible Ag, Cu, and Fe2O3 Nanoparticles. Mater. Today Chem. 2020, 15, 100214. [Google Scholar] [CrossRef]
  53. Dou, L.; Zhang, X.; Zangeneh, M.M.; Zhang, Y. Efficient Biogenesis of Cu2O Nanoparticles Using Extract of Camellia Sinensis Leaf: Evaluation of Catalytic, Cytotoxicity, Antioxidant, and Anti-Human Ovarian Cancer Properties. Bioorg Chem. 2021, 106, 104468. [Google Scholar] [CrossRef]
  54. Bloomberg Precious and Industial Metals. Available online: https://www.bloomberg.com/markets/commodities/futures/metals (accessed on 13 June 2023).
  55. Gu, J.; Aidy, A.; Goorani, S. Anti-Human Lung Adenocarcinoma, Cytotoxicity, and Antioxidant Potentials of Copper Nanoparticles Green-Synthesized by Calendula Officinalis. J. Exp. Nanosci. 2022, 17, 285–296. [Google Scholar] [CrossRef]
  56. Biresaw, S.S.; Taneja, P. Copper Nanoparticles Green Synthesis and Characterization as Anticancer Potential in Breast Cancer Cells (MCF7) Derived from Prunus Nepalensis Phytochemicals. In Proceedings of the Materials Today: Proceedings; Elsevier Ltd.: Amsterdam, The Netherlands, 2020; Volume 49, pp. 3501–3509. [Google Scholar]
  57. Amin, F.; Fozia; Khattak, B.; Alotaibi, A.; Qasim, M.; Ahmad, I.; Ullah, R.; Bourhia, M.; Gul, A.; Zahoor, S.; et al. Green Synthesis of Copper Oxide Nanoparticles Using Aerva Javanica Leaf Extract and Their Characterization and Investigation of in Vitro Antimicrobial Potential and Cytotoxic Activities. Evid. Based Complement. Altern. Med. 2021, 2021, 5589703. [Google Scholar] [CrossRef] [PubMed]
  58. Adeyemi, J.O.; Onwudiwe, D.C.; Oyedeji, A.O. Biogenic Synthesis of CuO, ZnO, and CuO–ZnO Nanoparticles Using Leaf Extracts of Dovyalis Caffra and Their Biological Properties. Molecules 2022, 27, 3206. [Google Scholar] [CrossRef]
  59. Kannan, K.; Radhika, D.; Vijayalakshmi, S.; Sadasivuni, K.K.; Ojiaku, A.A.; Verma, U. Facile Fabrication of CuO Nanoparticles via Microwave-Assisted Method: Photocatalytic, Antimicrobial and Anticancer Enhancing Performance. Int. J. Environ. Anal. Chem. 2022, 102, 1095–1108. [Google Scholar] [CrossRef]
  60. Jafar Ahamed, A.; Vijaya Kumar, P.; Loganathan, K.; Karthikeyan, C.; Haja Hameed, A.S. Synthesis, Characterization and Cytotoxicity Studies of CuO Nanoparticles by Using Gymnema Sylvestre Leaf Extracts. J. Indian. Chem. Soc. 2016, 93, 655–660. [Google Scholar]
  61. Zughaibi, T.A.; Mirza, A.A.; Suhail, M.; Jabir, N.R.; Zaidi, S.K.; Wasi, S.; Zawawi, A.; Tabrez, S. Evaluation of Anticancer Potential of Biogenic Copper Oxide Nanoparticles (CuO NPs) against Breast Cancer. J. Nanomater. 2022, 2022, 5326355. [Google Scholar] [CrossRef]
  62. Jiang, W.; Kim, B.Y.S.; Rutka, J.T.; Chan, W.C.W. Nanoparticle-Mediated Cellular Response Is Size-Dependent. Nat. Nanotechnol. 2008, 3, 145–150. [Google Scholar] [CrossRef]
  63. Mohammadyari, A.; Razavipour, S.T.; Mohammadbeigi, M.; Negahdary, M.; Ajdary, M. Explore In-Vivo Toxicity Assessment of Copper Oxide Nanoparticle in Wistar Rats. J. Biol. Today’s World 2014, 3, 124–128. [Google Scholar]
  64. Decuzzi, P.; Godin, B.; Tanaka, T.; Lee, S.-Y.; Chiappini, C.; Liu, X.; Ferrari, M. Size and Shape Effects in the Biodistribution of Intravascularly Injected Particles. J. Control. Release 2010, 141, 320–327. [Google Scholar] [CrossRef] [PubMed]
  65. Kumari, P.; Panda, P.K.; Jha, E.; Pramanik, N.; Nisha, K.; Kumari, K.; Soni, N.; Mallick, M.A.; Verma, S.K. Molecular Insight to in Vitro Biocompatibility of Phytofabricated Copper Oxide Nanoparticles with Human Embryonic Kidney Cells. Nanomedicine 2018, 13, 2415–2433. [Google Scholar] [CrossRef] [PubMed]
  66. Shahabadi, N.; Zendehcheshm, S.; Khademi, F. Green Synthesis, in Vitro Cytotoxicity, Antioxidant Activity and Interaction Studies of CuO Nanoparticles with DNA, Serum Albumin, Hemoglobin and Lysozyme. ChemistrySelect 2022, 7, 2365–6549. [Google Scholar] [CrossRef]
  67. Fouda, A.; Hassan, S.E.D.; Eid, A.M.; Awad, M.A.; Althumayri, K.; Badr, N.F.; Hamza, M.F. Endophytic Bacterial Strain, Brevibacillus Brevis-Mediated Green Synthesis of Copper Oxide Nanoparticles, Characterization, Antifungal, in Vitro Cytotoxicity, and Larvicidal Activity. Green. Process. Synth. 2022, 11, 931–950. [Google Scholar] [CrossRef]
  68. Nagajyothi, P.C.; Muthuraman, P.; Sreekanth, T.V.M.; Kim, D.H.; Shim, J. Green Synthesis: In-Vitro Anticancer Activity of Copper Oxide Nanoparticles against Human Cervical Carcinoma Cells. Arab. J. Chem. 2017, 10, 215–225. [Google Scholar] [CrossRef] [Green Version]
  69. Chen, H.; Feng, X.; Gao, L.; Mickymaray, S.; Paramasivam, A.; Abdulaziz Alfaiz, F.; Almasmoum, H.A.; Ghaith, M.M.; Almaimani, R.A.; Aziz Ibrahim, I.A. Inhibiting the PI3K/AKT/MTOR Signalling Pathway with Copper Oxide Nanoparticles from Houttuynia Cordata Plant: Attenuating the Proliferation of Cervical Cancer Cells. Artif. Cells Nanomed. Biotechnol. 2021, 49, 240–249. [Google Scholar] [CrossRef] [PubMed]
  70. Sankar, R.; Maheswari, R.; Karthik, S.; Shivashangari, K.S.; Ravikumar, V. Anticancer Activity of Ficus Religiosa Engineered Copper Oxide Nanoparticles. Mater. Sci. Eng. C 2014, 44, 234–239. [Google Scholar] [CrossRef] [PubMed]
  71. Hui, H.; Esmaeili, E.; Tayebee, R.; He, Q.; Abbaspour, S.; Akram, M.; Jalili, Z.; Mahdizadeh, N.; Ahmadi, A. Biosynthesis, Characterization, and Application of Cu2O Nanoparticles Originated from Cressa Leaf Extract as an Efficient Green Catalyst in the Synthesis of Some Chromenes. J. Iran. Chem. Soc. 2022, 19, 1261–1270. [Google Scholar] [CrossRef]
  72. Sulaiman, G.M.; Tawfeeq, A.T.; Jaaffer, M.D. Biogenic Synthesis of Copper Oxide Nanoparticles Using Olea Europaea Leaf Extract and Evaluation of Their Toxicity Activities: An in Vivo and in Vitro Study. Biotechnol. Prog. 2018, 34, 218–230. [Google Scholar] [CrossRef] [PubMed]
  73. Tabrez, S.; Khan, A.U.; Mirza, A.A.; Suhail, M.; Jabir, N.R.; Zughaibi, T.A.; Alam, M. Biosynthesis of Copper Oxide Nanoparticles and Its Therapeutic Efficacy against Colon Cancer. Nanotechnol. Rev. 2022, 11, 1322–1331. [Google Scholar] [CrossRef]
  74. Mukhopadhyay, R.; Kazi, J.; Debnath, M.C. Synthesis and Characterization of Copper Nanoparticles Stabilized with Quisqualis Indica Extract: Evaluation of Its Cytotoxicity and Apoptosis in B16F10 Melanoma Cells. Biomed. Pharmacother. 2018, 97, 1373–1385. [Google Scholar] [CrossRef] [PubMed]
  75. Nakhaeepour, Z.; Mashreghi, M.; Matin, M.M.; NakhaeiPour, A.; Housaindokht, M.R. Multifunctional CuO Nanoparticles with Cytotoxic Effects on KYSE30 Esophageal Cancer Cells, Antimicrobial and Heavy Metal Sensing Activities. Life Sci. 2019, 234, 116758. [Google Scholar] [CrossRef] [PubMed]
  76. Sonbol, H.; AlYahya, S.; Ameen, F.; Alsamhary, K.; Alwakeel, S.; Al-Otaibi, S.; Korany, S. Bioinspired Synthesize of CuO Nanoparticles Using Cylindrospermum Stagnale for Antibacterial, Anticancer and Larvicidal Applications. Appl. Nanosci. 2021, 13, 917–927. [Google Scholar] [CrossRef]
  77. Raghav, P.K.; Mann, Z.; Ahlawat, S.; Mohanty, S. Mesenchymal Stem Cell-Based Nanoparticles and Scaffolds in Regenerative Medicine. Eur. J. Pharm. 2022, 918, 174657. [Google Scholar] [CrossRef]
  78. Shen, Q.; Qi, Y.; Kong, Y.; Bao, H.; Wang, Y.; Dong, A.; Wu, H.; Xu, Y. Advances in Copper-Based Biomaterials with Antibacterial and Osteogenic Properties for Bone Tissue Engineering. Front. Bioeng. Biotechnol. 2022, 9, 1526. [Google Scholar] [CrossRef] [PubMed]
  79. Teno, J.; Corral, A.; Gorrasi, G.; Sorrentino, A.; Benito, J.G. Fibrous Nanocomposites Based on EVA40 Filled with Cu Nanoparticles and Their Potential Antibacterial Action. Mater. Today Commun. 2019, 20, 100581. [Google Scholar] [CrossRef]
  80. Wang, Y.; Zhang, W.; Yao, Q. Copper-Based Biomaterials for Bone and Cartilage Tissue Engineering. J. Orthop. Transl. 2021, 29, 60–71. [Google Scholar] [CrossRef] [PubMed]
  81. Yugandhar, P.; Vasavi, T.; Uma Maheswari Devi, P.; Savithramma, N. Bioinspired Green Synthesis of Copper Oxide Nanoparticles from Syzygium Alternifolium (Wt.) Walp: Characterization and Evaluation of Its Synergistic Antimicrobial and Anticancer Activity. Appl. Nanosci. 2017, 7, 417–427. [Google Scholar] [CrossRef] [Green Version]
  82. Agila, A.; Vimala, J.R.; Bharathy, M.; Dayana Jeyaleela, G.; Sheela, S.M. Anti-Oxidant and Anti-Cancer Activities of Biogenic Synthesized Copper Oxide Nanoparticles. Biomed. Biotechnol. Res. J. (BBRJ) 2022, 6, 341. [Google Scholar] [CrossRef]
  83. Mahmoud, N.M.R.; Mohamed, H.I.; Ahmed, S.B.; Akhtar, S. Efficient Biosynthesis of CuO Nanoparticles with Potential Cytotoxic Activity. Chem. Pap. 2020, 74, 2825–2835. [Google Scholar] [CrossRef]
  84. Mohamed, R.M.; Fawzy, E.M.; Shehab, R.A.; Ali, D.M.; Salah, R.A.; Din, E.; Abd, H.M.; Fatah, E. Green Biosynthesis, Structural Characterization and Anticancer Activity of Copper Oxide Nanoparticles from the Brown Alga Cystoseira Myrica. Egypt. J. Aquatic Biol. Fisheries 2021, 25, 341–358. [Google Scholar] [CrossRef]
  85. Siddiquee, M.A.; ud din Parray, M.; Kamli, M.R.; Malik, M.A.; Mehdi, S.H.; Imtiyaz, K.; Rizvi, M.M.A.; Rajor, H.K.; Patel, R. Biogenic Synthesis, in-Vitro Cytotoxicity, Esterase Activity and Interaction Studies of Copper Oxide Nanoparticles with Lysozyme. J. Mater. Res. Technol. 2021, 13, 2066–2077. [Google Scholar] [CrossRef]
  86. Sivaraj, R.; Rahman, P.K.S.M.; Rajiv, P.; Narendhran, S.; Venckatesh, R. Biosynthesis and Characterization of Acalypha Indica Mediated Copper Oxide Nanoparticles and Evaluation of Its Antimicrobial and Anticancer Activity. Spectrochim. Acta A Mol. Biomol. Spectrosc. 2014, 129, 255–258. [Google Scholar] [CrossRef] [PubMed]
  87. Sharma, S.; Kumar, K.; Thakur, N.; Chauhan, S.; Chauhan, M.S. Eco-Friendly Ocimum Tenuiflorum Green Route Synthesis of CuO Nanoparticles: Characterizations on Photocatalytic and Antibacterial Activities. J. Environ. Chem. Eng. 2021, 9, 105395. [Google Scholar] [CrossRef]
  88. Sabeena, G.; Rajaduraipandian, S.; Pushpalakshmi, E.; Alhadlaq, H.A.; Mohan, R.; Annadurai, G.; Ahamed, M. Green and Chemical Synthesis of CuO Nanoparticles: A Comparative Study for Several in Vitro Bioactivities and in Vivo Toxicity in Zebrafish Embryos. J. King Saud. Univ. Sci. 2022, 34, 102092. [Google Scholar] [CrossRef]
  89. Sundaram, C.S.; Sivakumar, J.; Suresh Kumar, S.; Ramesh, P.L.N.; Zin, T.; Mahadeva Rao, U.S. Antibacterial and Anticancer Potential of Brassica Oleracea Var Acephala Using Biosynthesised Copper Nanoparticles. Med. J. Malays. 2020, 75, 677–684. [Google Scholar]
  90. Dulta, K.; Ağçeli, G.K.; Chauhan, P.; Chauhan, P.K. Biogenic Production and Characterization of CuO Nanoparticles by Carica Papaya Leaves and Its Biocompatibility Applications. J. Inorg. Organomet. Polym. Mater. 2021, 31, 1846–1857. [Google Scholar] [CrossRef]
  91. Gamedze, N.P.; Mthiyane, D.M.N.; Babalola, O.O.; Singh, M.; Onwudiwe, D.C. Physico-Chemical Characteristics and Cytotoxicity Evaluation of CuO and TiO2 Nanoparticles Biosynthesized Using Extracts of Mucuna Pruriens Utilis Seeds. Heliyon 2022, 8, e10187. [Google Scholar] [CrossRef]
  92. Faisal, S.; Al-Radadi, N.S.; Jan, H.; Abdullah; Shah, S.A.; Shah, S.; Rizwan, M.; Afsheen, Z.; Hussain, Z.; Uddin, M.N.; et al. Curcuma Longa Mediated Synthesis of Copper Oxide, Nickel Oxide and Cu-Ni Bimetallic Hybrid Nanoparticles: Characterization and Evaluation for Antimicrobial, Anti-Parasitic and Cytotoxic Potentials. Coatings 2021, 11, 849. [Google Scholar] [CrossRef]
  93. Fakhar-e-Alam, M.; Shafiq, Z.; Mahmood, A.; Atif, M.; Anwar, H.; Hanif, A.; Yaqub, N.; Farooq, W.A.; Fatehmulla, A.; Ahmad, S.; et al. Assessment of Green and Chemically Synthesized Copper Oxide Nanoparticles against Hepatocellular Carcinoma. J. King Saud. Univ. Sci. 2021, 33, 101669. [Google Scholar] [CrossRef]
  94. Liu, Y.; Zeng, Z.; Jiang, O.; Li, Y.X.; Xu, Q.; Jiang, L.J.; Yu, J.; Xu, D. Green Synthesis of CuO NPs, Characterization and Their Toxicity Potential against HepG2 Cells. Mater. Res. Express 2021, 8, 15011. [Google Scholar] [CrossRef]
  95. Majid, A.; Faraj, H.R. Green Synthesis of Copper Nanoparticles Using Aqueous Extract of Yerba Mate (Llex Paraguarients St. Hill) and Its Anticancer Activity. Int. J. Nanosci. Nanotechnol. 2022, 18, 99–108. [Google Scholar]
  96. Saravanakumar, K.; Shanmugam, S.; Varukattu, N.B.; MubarakAli, D.; Kathiresan, K.; Wang, M.H. Biosynthesis and Characterization of Copper Oxide Nanoparticles from Indigenous Fungi and Its Effect of Photothermolysis on Human Lung Carcinoma. J. Photochem. Photobiol. B 2019, 190, 103–109. [Google Scholar] [CrossRef] [PubMed]
  97. Liu, H.; Wang, G.; Liu, J.; Nan, K.; Zhang, J.; Guo, L.; Liu, Y. Green Synthesis of Copper Nanoparticles Using Cinnamomum Zelanicum Extract and Its Applications as a Highly Efficient Antioxidant and Anti-Human Lung Carcinoma. J. Exp. Nanosci. 2021, 16, 411–423. [Google Scholar] [CrossRef]
  98. Manasa, D.J.; Chandrashekar, K.R.; Madhu Kumar, D.J.; Niranjana, M.; Navada, K.M. Mussaenda frondosa L. Mediated Facile Green Synthesis of Copper Oxide Nanoparticles—Characterization, Photocatalytic and Their Biological Investigations. Arab. J. Chem. 2021, 14, 103184. [Google Scholar] [CrossRef]
  99. Gnanavel, V.; Palanichamy, V.; Roopan, S.M. Biosynthesis and Characterization of Copper Oxide Nanoparticles and Its Anticancer Activity on Human Colon Cancer Cell Lines (HCT-116). J. Photochem. Photobiol. B 2017, 171, 133–138. [Google Scholar] [CrossRef] [PubMed]
  100. Prasad, P.R.; Kanchi, S.; Naidoo, E.B. In-Vitro Evaluation of Copper Nanoparticles Cytotoxicity on Prostate Cancer Cell Lines and Their Antioxidant, Sensing and Catalytic Activity: One-Pot Green Approach. J. Photochem. Photobiol. B 2016, 161, 375–382. [Google Scholar] [CrossRef] [PubMed]
  101. Naz, S.; Tabassum, S.; Freitas Fernandes, N.; Mujahid, M.; Zia, M.; Carcache de Blanco, E.J. Anticancer and Antibacterial Potential of Rhus Punjabensis and CuO Nanoparticles. Nat. Prod. Res. 2020, 34, 720–725. [Google Scholar] [CrossRef]
  102. Vincent, J.; Lekha, N.C. Bio-Engineered Copper Oxide Nanoparticles Using Citrus Aurantifolia Enzyme Extract and Its Anticancer Activity. J. Clust. Sci. 2022, 33, 45–53. [Google Scholar] [CrossRef]
  103. Yu, Y.; Fei, Z.; Cui, J.; Miao, B.; Lu, Y.; Wu, J. Biosynthesis of Copper Oxide Nanoparticles and Their in Vitro Cytotoxicity towards Nasopharynx Cancer (KB Cells) Cell Lines. Int. J. Pharmacol. 2018, 14, 609–614. [Google Scholar] [CrossRef] [Green Version]
  104. Yugandhar, P.; Savithramma, N. Biosynthesis, Characterization and Antimicrobial Studies of Green Synthesized Silver Nanoparticles from Fruit Extract of Syzygium Alternifolium (Wt.) Walp. an Endemic, Endangered Medicinal Tree Taxon. Appl. Nanosci. 2016, 6, 223–233. [Google Scholar] [CrossRef] [Green Version]
  105. 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]
  106. Baiker, A.; Monti, D.; Fan, Y.S. Deactivation of Copper, Nickel, and Cobalt Catalysts by Interaction with Aliphatic Amines. J. Catal. 1984, 88, 81–88. [Google Scholar] [CrossRef]
  107. Guo, Z.; Chen, G.; Liu, L.; Zeng, G.; Huang, Z.; Chen, A.; Hu, L. Activity Variation of Phanerochaete Chrysosporium under Nanosilver Exposure by Controlling of Different Sulfide Sources. Sci. Rep. 2016, 6, 20813. [Google Scholar] [CrossRef] [Green Version]
  108. Bogdanović, U.; Lazić, V.; Vodnik, V.; Budimir, M.; Marković, Z.; Dimitrijević, S. Copper Nanoparticles with High Antimicrobial Activity. Mater. Lett. 2014, 128, 75–78. [Google Scholar] [CrossRef]
  109. Meghana, S.; Kabra, P.; Chakraborty, S.; Padmavathy, N. Understanding the Pathway of Antibacterial Activity of Copper Oxide Nanoparticles. RSC Adv. 2015, 5, 12293–12299. [Google Scholar] [CrossRef]
  110. Raja, S.; Ramesh, V.; Thivaharan, V. Green Biosynthesis of Silver Nanoparticles Using Calliandra Haematocephala Leaf Extract, Their Antibacterial Activity and Hydrogen Peroxide Sensing Capability. Arab. J. Chem. 2017, 10, 253–261. [Google Scholar] [CrossRef] [Green Version]
  111. Ul-Hamid, A.; Dafalla, H.; Hakeem, A.S.; Haider, A.; Ikram, M. In-Vitro Catalytic and Antibacterial Potential of Green Synthesized CuO Nanoparticles against Prevalent Multiple Drug Resistant Bovine Mastitogen Staphylococcus Aureus. Int. J. Mol. Sci. 2022, 23, 2335. [Google Scholar] [CrossRef] [PubMed]
  112. Gold, K.; Slay, B.; Knackstedt, M.; Gaharwar, A.K. Antimicrobial Activity of Metal and Metal-oxide Based Nanoparticles. Adv. Ther. 2018, 1, 1700033. [Google Scholar] [CrossRef]
  113. Ghosh, S.; Kaushik, R.; Nagalakshmi, K.; Hoti, S.L.; Menezes, G.A.; Harish, B.N.; Vasan, H.N. Antimicrobial Activity of Highly Stable Silver Nanoparticles Embedded in Agar–Agar Matrix as a Thin Film. Carbohydr. Res. 2010, 345, 2220–2227. [Google Scholar] [CrossRef]
  114. Li, Y.; Yang, D.; Cui, J. Graphene Oxide Loaded with Copper Oxide Nanoparticles as an Antibacterial Agent against Pseudomonas Syringae Pv. Tomato. RSC Adv. 2017, 7, 38853–38860. [Google Scholar] [CrossRef] [Green Version]
  115. Silva, C.J.; Barbosa, L.C.A.; Demuner, A.J.; Montanari, R.M.; Pinheiro, A.L.; Dias, I.; Andrade, N.J. Chemical Composition and Antibacterial Activities from the Essential Oils of Myrtaceae Species Planted in Brazil. Quim. Nova 2010, 33, 104–108. [Google Scholar] [CrossRef]
  116. Arunkumar, B.; Jeyakumar, S.J.; Jothibas, M. A Sol-Gel Approach to the Synthesis of CuO Nanoparticles Using Lantana Camara Leaf Extract and Their Photo Catalytic Activity. Optik 2019, 183, 698–705. [Google Scholar] [CrossRef]
  117. Deans, S.G.; Ritchie, G. Antibacterial Properties of Plant Essential Oils. Int. J. Food Microbiol. 1987, 5, 165–180. [Google Scholar] [CrossRef]
  118. Suresh, S.; Vennila, S.; Anita Lett, J.; Fatimah, I.; Mohammad, F.; Al-Lohedan, H.A.; Alshahateet, S.F.; Motalib Hossain, M.A.; Rafie Johan, M. Star Fruit Extract-Mediated Green Synthesis of Metal Oxide Nanoparticles. Inorg. Nano-Met. Chem. 2022, 52, 173–180. [Google Scholar] [CrossRef]
  119. Nieto-Maldonado, A.; Bustos-Guadarrama, S.; Espinoza-Gomez, H.; Flores-López, L.Z.; Ramirez-Acosta, K.; Alonso-Nuñez, G.; Cadena-Nava, R.D. Green Synthesis of Copper Nanoparticles Using Different Plant Extracts and Their Antibacterial Activity. J. Environ. Chem. Eng. 2022, 10, 107130. [Google Scholar] [CrossRef]
  120. Vinu, D.; Govindaraju, K.; Vasantharaja, R.; Amreen Nisa, S.; Kannan, M.; Vijai Anand, K. Biogenic Zinc Oxide, Copper Oxide and Selenium Nanoparticles: Preparation, Characterization and Their Anti-Bacterial Activity against Vibrio Parahaemolyticus. J. Nanostructure Chem. 2021, 11, 271–286. [Google Scholar] [CrossRef]
  121. Shehabeldine, A.M.; Amin, B.H.; Hagras, F.A.; Ramadan, A.A.; Kamel, M.R.; Ahmed, M.A.; Atia, K.H.; Salem, S.S. Potential Antimicrobial and Antibiofilm Properties of Copper Oxide Nanoparticles: Time-Kill Kinetic Essay and Ultrastructure of Pathogenic Bacterial Cells. Appl. Biochem. Biotechnol. 2022, 195, 467–485. [Google Scholar] [CrossRef]
  122. Sharma, S.; Kumar, K. Aloe-Vera Leaf Extract as a Green Agent for the Synthesis of CuO Nanoparticles Inactivating Bacterial Pathogens and Dye. J. Dispers. Sci. Technol. 2020, 42, 1950–1962. [Google Scholar] [CrossRef]
  123. Emima Jeronsia, J.; Allwin Joseph, L.; Annie Vinosha, P.; Jerline Mary, A.; Jerome Das, S. Camellia Sinensis Leaf Extract Mediated Synthesis of Copper Oxide Nanostructures for Potential Biomedical Applications. Mater. Today Proc. 2019, 8, 214–222. [Google Scholar] [CrossRef]
  124. Sackey, J.; Razanamahandry, L.C.; Ntwampe, S.K.O.; Mlungisi, N.; Fall, A.; Kaonga, C.; Nuru, Z.Y. Biosynthesis of CuO Nanoparticles Using Mimosa Hamata Extracts. Mater. Today Proc. 2019, 36, 540–548. [Google Scholar] [CrossRef]
  125. Nagore, P.; Ghotekar, S.; Mane, K.; Ghoti, A.; Bilal, M.; Roy, A. Structural Properties and Antimicrobial Activities of Polyalthia Longifolia Leaf Extract-Mediated CuO Nanoparticles. Bionanoscience 2021, 11, 579–589. [Google Scholar] [CrossRef]
  126. Arya, A.; Gupta, K.; Chundawat, T.S.; Vaya, D. Biogenic Synthesis of Copper and Silver Nanoparticles Using Green Alga Botryococcus Braunii and Its Antimicrobial Activity. Bioinorg. Chem. Appl. 2018, 2018, 7879403. [Google Scholar] [CrossRef] [Green Version]
  127. Nadeem, A.; Sumbal; Ali, J.S.; Latif, M.; Rizvi, Z.F.; Naz, S.; Mannan, A.; Zia, M. Green Synthesis and Characterization of Fe, Cu and Mg Oxide Nanoparticles Using Clematis Orientalis Leaf Extract: Salt Concentration Modulates Physiological and Biological Properties. Mater. Chem. Phys. 2021, 271, 124900. [Google Scholar] [CrossRef]
  128. Shanmugapriya, J.; Reshma, C.A.; Srinidhi, V.; Harithpriya, K.; Ramkumar, K.M.; Umpathy, D.; Gunasekaran, K.; Subashini, R. Green Synthesis of Copper Nanoparticles Using Withania Somnifera and Its Antioxidant and Antibacterial Activity. J. Nanomater. 2022, 2022, 7967294. [Google Scholar] [CrossRef]
  129. Razavi, R.; Molaei, R.; Moradi, M.; Tajik, H.; Ezati, P.; Shafipour Yordshahi, A. Biosynthesis of Metallic Nanoparticles Using Mulberry Fruit (Morus alba L.) Extract for the Preparation of Antimicrobial Nanocellulose Film. Appl. Nanosci. 2020, 10, 465–476. [Google Scholar] [CrossRef]
  130. Raveesha, H.R.; Bharath, H.L.; Vasudha, D.R.; Sushma, B.K.; Pratibha, S.; Dhananjaya, N. Antibacterial and Antiproliferation Activity of Green Synthesized Nanoparticles from Rhizome Extract of Alpinia galangal (L.) Wild. Inorg. Chem. Commun. 2021, 132, 108854. [Google Scholar] [CrossRef]
  131. Prakash, S.; Elavarasan, N.; Venkatesan, A.; Subashini, K.; Sowndharya, M.; Sujatha, V. Green Synthesis of Copper Oxide Nanoparticles and Its Effective Applications in Biginelli Reaction, BTB Photodegradation and Antibacterial Activity. Adv. Powder Technol. 2018, 29, 3315–3326. [Google Scholar] [CrossRef]
  132. Naseer, M.; Ramadan, R.; Xing, J.; Samak, N.A. Facile Green Synthesis of Copper Oxide Nanoparticles for the Eradication of Multidrug Resistant Klebsiella Pneumonia and Helicobacter Pylori Biofilms. Int. Biodeterior. Biodegrad. 2021, 159, 105201. [Google Scholar] [CrossRef]
  133. Naika, H.R.; Lingaraju, K.; Manjunath, K.; Kumar, D.; Nagaraju, G.; Suresh, D.; Nagabhushana, H. Green Synthesis of CuO Nanoparticles Using Gloriosa superba L. Extract and Their Antibacterial Activity. J. Taibah Univ. Sci. 2015, 9, 7–12. [Google Scholar] [CrossRef] [Green Version]
  134. Lingaraju, K.; Raja Naika, H.; Manjunath, K.; Nagaraju, G.; Suresh, D.; Nagabhushana, H. Rauvolfia Serpentina-Mediated Green Synthesis of CuO Nanoparticles and Its Multidisciplinary Studies. Acta Metall. Sin. (Engl. Lett.) 2015, 28, 1134–1140. [Google Scholar] [CrossRef]
  135. Kumar, P.P.N.V.; Shameem, U.; Kollu, P.; Kalyani, R.L.; Pammi, S.V.N. Green Synthesis of Copper Oxide Nanoparticles Using Aloe Vera Leaf Extract and Its Antibacterial Activity Against Fish Bacterial Pathogens. Bionanoscience 2015, 5, 135–139. [Google Scholar] [CrossRef]
  136. Kalaiyan, G.; Suresh, S.; Thambidurai, S.; Prabu, K.M.; Kumar, S.K.; Pugazhenthiran, N.; Kandasamy, M. Green Synthesis of Hierarchical Copper Oxide Microleaf Bundles Using Hibiscus Cannabinus Leaf Extract for Antibacterial Application. J. Mol. Struct. 2020, 1217, 128379. [Google Scholar] [CrossRef]
  137. Velsankar, K.; Suganya, S.; Muthumari, P.; Mohandoss, S.; Sudhahar, S. Ecofriendly Green Synthesis, Characterization and Biomedical Applications of CuO Nanoparticles Synthesized Using Leaf Extract of Capsicum Frutescens. J. Environ. Chem. Eng. 2021, 9, 106299. [Google Scholar] [CrossRef]
  138. Aziz, W.J.; Abid, M.A.; Hussein, E.H. Biosynthesis of CuO Nanoparticles and Synergistic Antibacterial Activity Using Mint Leaf Extract. Mater. Technol. 2020, 35, 447–451. [Google Scholar] [CrossRef]
  139. Angeline Mary, A.P.; Thaminum Ansari, A.; Subramanian, R. Sugarcane Juice Mediated Synthesis of Copper Oxide Nanoparticles, Characterization and Their Antibacterial Activity. J. King Saud. Univ. Sci. 2019, 31, 1103–1114. [Google Scholar] [CrossRef]
  140. Andualem, W.W.; Sabir, F.K.; Mohammed, E.T.; Belay, H.H.; Gonfa, B.A. Synthesis of Copper Oxide Nanoparticles Using Plant Leaf Extract of Catha Edulis and Its Antibacterial Activity. J. Nanotechnol. 2020, 2020, 2932434. [Google Scholar] [CrossRef]
  141. Alishah, H.; Pourseyedi, S.; Ebrahimipour, S.Y.; Mahani, S.E.; Rafiei, N. Green Synthesis of Starch-Mediated CuO Nanoparticles: Preparation, Characterization, Antimicrobial Activities and in Vitro MTT Assay against MCF-7 Cell Line. Rend. Lincei 2017, 28, 65–71. [Google Scholar] [CrossRef]
  142. Das, P.; Ghosh, S.; Ghosh, R.; Dam, S.; Baskey, M. Madhuca Longifolia Plant Mediated Green Synthesis of Cupric Oxide Nanoparticles: A Promising Environmentally Sustainable Material for Waste Water Treatment and Efficient Antibacterial Agent. J. Photochem. Photobiol. B 2018, 189, 66–73. [Google Scholar] [CrossRef] [PubMed]
  143. Bocarando-Chacón, J.; Vargas-Vazquez, D.; Martinez-Suarez, F.; Flores-Juárez, C.; Cortez-Valadez, M. Surface-Enhanced Raman Scattering and Antibacterial Properties from Copper Nanoparticles Obtained by Green Chemistry. Appl. Phys. A Mater. Sci. Process. 2020, 126, 530. [Google Scholar] [CrossRef]
  144. Vasantharaj, S.; Shivakumar, P.; Sathiyavimal, S.; Senthilkumar, P.; Vijayaram, S.; Shanmugavel, M.; Pugazhendhi, A. Antibacterial Activity and Photocatalytic Dye Degradation of Copper Oxide Nanoparticles (CuONPs) Using Justicia Gendarussa. Appl. Nanosci. 2021, 13, 2295–2302. [Google Scholar] [CrossRef]
  145. Vasantharaj, S.; Sathiyavimal, S.; Saravanan, M.; Senthilkumar, P.; Gnanasekaran, K.; Shanmugavel, M.; Manikandan, E.; Pugazhendhi, A. Synthesis of Ecofriendly Copper Oxide Nanoparticles for Fabrication over Textile Fabrics: Characterization of Antibacterial Activity and Dye Degradation Potential. J. Photochem. Photobiol. B 2019, 191, 143–149. [Google Scholar] [CrossRef]
  146. Vaidehi, D.; Bhuvaneshwari, V.; Bharathi, D.; Sheetal, B.P. Antibacterial and Photocatalytic Activity of Copper Oxide Nanoparticles Synthesized Using Solanum Lycopersicum Leaf Extract. Mater. Res. Express 2018, 5, 85403. [Google Scholar] [CrossRef]
  147. Tamil Elakkiya, V.; Meenakshi, R.V.; Senthil Kumar, P.; Karthik, V.; Ravi Shankar, K.; Sureshkumar, P.; Hanan, A. Green Synthesis of Copper Nanoparticles Using Sesbania Aculeata to Enhance the Plant Growth and Antimicrobial Activities. Int. J. Environ. Sci. Technol. 2022, 19, 1313–1322. [Google Scholar] [CrossRef]
  148. Tahvilian, R.; Zangeneh, M.M.; Falahi, H.; Sadrjavadi, K.; Jalalvand, A.R.; Zangeneh, A. Green Synthesis and Chemical Characterization of Copper Nanoparticles Using Allium Saralicum Leaves and Assessment of Their Cytotoxicity, Antioxidant, Antimicrobial, and Cutaneous Wound Healing Properties. Appl. Organomet. Chem. 2019, 33, 268–2605. [Google Scholar] [CrossRef]
  149. Taherzadeh Soureshjani, P.; Shadi, A.; Mohammadsaleh, F. Algae-Mediated Route to Biogenic Cuprous Oxide Nanoparticles and Spindle-like CaCO3: A Comparative Study, Facile Synthesis, and Biological Properties. RSC Adv. 2021, 11, 10599–10609. [Google Scholar] [CrossRef]
  150. Sivaraj, R.; Rahman, P.K.S.M.; Rajiv, P.; Salam, H.A.; Venckatesh, R. Biogenic Copper Oxide Nanoparticles Synthesis Using Tabernaemontana Divaricate Leaf Extract and Its Antibacterial Activity against Urinary Tract Pathogen. Spectrochim. Acta A Mol. Biomol. Spectrosc. 2014, 133, 178–181. [Google Scholar] [CrossRef]
  151. Shende, S.; Ingle, A.P.; Gade, A.; Rai, M. Green Synthesis of Copper Nanoparticles by Citrus Medica Linn. (Idilimbu) Juice and Its Antimicrobial Activity. World J. Microbiol. Biotechnol. 2015, 31, 865–873. [Google Scholar] [CrossRef]
  152. Seetha, J.; Mallavarapu, U.; Mesa, A. In Situ Green Synthesis of Antibacterial Copper Nanocomposite Cotton Fabrics Using Achyranthes Aspera Leaf Extract. J. Appl. Pharm. Sci. 2020, 10, 104–109. [Google Scholar] [CrossRef]
  153. Roy, K.; Sarkar, C.K.; Ghosh, C.K. Antibacterial Mechanism of Biogenic Copper Nanoparticles Synthesized Using Heliconia Psittacorum Leaf Extract. Nanotechnol. Rev. 2016, 5, 529–536. [Google Scholar] [CrossRef]
  154. Ramzan, M.; Obodo, R.M.; Mukhtar, S.; Ilyas, S.Z.; Aziz, F.; Thovhogi, N. Green Synthesis of Copper Oxide Nanoparticles Using Cedrus Deodara Aqueous Extract for Antibacterial Activity. Mater. Today Proc. 2019, 36, 576–581. [Google Scholar] [CrossRef]
  155. Rafique, M.; Tahir, M.B.; Irshad, M.; Nabi, G.; Gillani, S.S.A.; Iqbal, T.; Mubeen, M. Novel Citrus Aurantifolia Leaves Based Biosynthesis of Copper Oxide Nanoparticles for Environmental and Wastewater Purification as an Efficient Photocatalyst and Antibacterial Agent. Optik 2020, 219, 165138. [Google Scholar] [CrossRef]
  156. Nagaraj, E.; Karuppannan, K.; Shanmugam, P.; Venugopal, S. Exploration of Bio-Synthesized Copper Oxide Nanoparticles Using Pterolobium Hexapetalum Leaf Extract by Photocatalytic Activity and Biological Evaluations. J. Clust. Sci. 2019, 30, 1157–1168. [Google Scholar] [CrossRef]
  157. Hosseinzadeh, R.; Mohadjerani, M.; Mesgar, S. Green Synthesis of Copper Oxide Nanoparticles Using Aqueous Extract of Convolvulus percicus L. as Reusable Catalysts in Cross-Coupling Reactions and Their Antibacterial Activity. IET Nanobiotechnol. 2017, 11, 725–730. [Google Scholar] [CrossRef]
  158. Gopalakrishnan, V.; Muniraj, S. Neem Flower Extract Assisted Green Synthesis of Copper Nanoparticles—Optimisation, Characterisation and Anti-Bacterial Study. Mater. Today Proc. 2019, 36, 832–836. [Google Scholar] [CrossRef]
  159. Gholami, M.; Azarbani, F.; Hadi, F.; Murthy, H.C.A. Eco-Friendly Synthesis of Copper Nanoparticles Using Mentha Pulegium Leaf Extract: Characterisation, Antibacterial and Cytotoxic Activities. Mater. Technol. 2022, 37, 1523–1531. [Google Scholar] [CrossRef]
  160. Fernandez, A.C.; KM, A.; Rajagopal, R. Green Synthesis, Characterization, Catalytic and Antibacterial Studies of Copper Iodide Nanoparticles Synthesized Using Brassica Oleracea Var. Capitata f. Rubra Extract. Chem. Data Collect. 2020, 29, 100538. [Google Scholar] [CrossRef]
  161. Fatma, S.; Kalainila, P.; Fatma, S.; Renganathan, S. Green Synthesis of Copper Nanoparticle from Passiflora Foetida Leaf Extract and Its Antibacterial Activity. Asian J. Pharm. Clin. Res. 2017, 10, 79–83. [Google Scholar] [CrossRef] [Green Version]
  162. Zaman, M.B.; Poolla, R.; Singh, P.; Gudipati, T. Biogenic Synthesis of CuO Nanoparticles Using Tamarindus indica L. and a Study of Their Photocatalytic and Antibacterial Activity. Environ. Nanotechnol. Monit. Manag. 2020, 14, 100346. [Google Scholar] [CrossRef]
  163. Velmurugan, P.; Jang, S.H.; Hong, S.C.; Yi, P.I.; Jung, E.S.; Park, J.S.; Sivakumar, S. Green Crystallization and Characterization of Copper Oxide (CuO) Nanoparticles Using Anacardium Occidentale Shell Liquid and Their Biomedical Applications. J. Nano Res. 2016, 40, 167–173. [Google Scholar] [CrossRef]
  164. Taha, J.H.; Abbas, N.K.; Al-Attraqchi, A.A.F. Green Synthesis and Evaluation of Copper Oxide Nanoparticles Using Fig Leaves and Their Antifungal and Antibacterial Activities. Int. J. Drug. Deliv. Technol. 2020, 10, 378–382. [Google Scholar] [CrossRef]
  165. Sharma, P.; Pant, S.; Dave, V.; Tak, K.; Sadhu, V.; Reddy, K.R. Green Synthesis and Characterization of Copper Nanoparticles by Tinospora Cardifolia to Produce Nature-Friendly Copper Nano-Coated Fabric and Their Antimicrobial Evaluation. J. Microbiol. Methods 2019, 160, 107–116. [Google Scholar] [CrossRef]
  166. Reddeppa, M.; Reddy, R.C.K.; Raj, Y.P.; Rani, T.S. Green Synthesis of Copper Nanoparticles: Evaluation of Catalytic and Antibacterial Activity. Asian J. Chem. 2019, 31, 622–626. [Google Scholar] [CrossRef]
  167. Jain, V.; Khusnud, A.; Tiwari, J.; Mishra, M.; Mishra, P.K. Biogenic Proceedings and Characterization of Copper-Gold Nanoalloy: Evaluation of Their Innate Antimicrobial and Catalytic Activities. Inorg. Nano-Met. Chem. 2021, 51, 230–238. [Google Scholar] [CrossRef]
  168. Fazal, A.; Ara, S.; Ishaq, M.T.; Sughra, K. Green Fabrication of Copper Oxide Nanoparticles: A Comparative Antibacterial Study Against Gram-Positive and Gram-Negative Bacteria. Arab. J. Sci. Eng. 2022, 47, 523–533. [Google Scholar] [CrossRef]
  169. Benassai, E.; Del Bubba, M.; Ancillotti, C.; Colzi, I.; Gonnelli, C.; Calisi, N.; Salvatici, M.C.; Casalone, E.; Ristori, S. Green and Cost-Effective Synthesis of Copper Nanoparticles by Extracts of Non-Edible and Waste Plant Materials from Vaccinium Species: Characterization and Antimicrobial Activity. Mater. Sci. Eng. C. 2021, 119, 111453. [Google Scholar] [CrossRef]
  170. Suramwar, N.V.; Thakare, S.R.; Karade, N.N.; Khaty, N.T. Green Synthesis of Predominant (1 1 1) Facet CuO Nanoparticles: Heterogeneous and Recyclable Catalyst for N-Arylation of Indoles. J. Mol. Catal. A Chem. 2012, 359, 28–34. [Google Scholar] [CrossRef]
  171. Manjari, G.; Saran, S.; Arun, T.; Vijaya Bhaskara Rao, A.; Devipriya, S.P. Catalytic and Recyclability Properties of Phytogenic Copper Oxide Nanoparticles Derived from Aglaia Elaeagnoidea Flower Extract. J. Saudi Chem. Soc. 2017, 21, 610–618. [Google Scholar] [CrossRef]
  172. Kiriyanthan, R.M.; Sharmili, S.A.; Balaji, R.; Jayashree, S.; Mahboob, S.; Al-Ghanim, K.A.; Al-Misned, F.; Ahmed, Z.; Govindarajan, M.; Vaseeharan, B. Photocatalytic, Antiproliferative and Antimicrobial Properties of Copper Nanoparticles Synthesized Using Manilkara Zapota Leaf Extract: A Photodynamic Approach. Photodiagnosis Photodyn. Ther. 2020, 32, 102058. [Google Scholar] [CrossRef]
  173. Wang, Z.L. Characterizing the Structure and Properties of Individual Wire-Like Nanoentities. Adv. Mater. 2000, 12, 1295–1298. [Google Scholar] [CrossRef]
  174. Nasrollahzadeh, M.; Sajadi, S.M.; Maham, M. Tamarix Gallica Leaf Extract Mediated Novel Route for Green Synthesis of CuO Nanoparticles and Their Application for N-Arylation of Nitrogen-Containing Heterocycles under Ligand-Free Conditions. RSC Adv. 2015, 5, 40628–40635. [Google Scholar] [CrossRef]
  175. Selvam, K.; Sudhakar, C.; Selvankumar, T.; Senthilkumar, B.; Selva Kumar, R.; Kannan, N. Biomimetic Synthesis of Copper Nanoparticles Using Rhizome Extract of Corallocarbus Epigaeus and Their Bactericidal with Photocatalytic Activity. SN Appl. Sci. 2020, 2, 1–7. [Google Scholar] [CrossRef]
  176. Dulta, K.; Koşarsoy Ağçeli, G.; Chauhan, P.; Jasrotia, R.; Chauhan, P.K.; Ighalo, J.O. Multifunctional CuO Nanoparticles with Enhanced Photocatalytic Dye Degradation and Antibacterial Activity. Sustain. Environ. Res. 2022, 32, 1–15. [Google Scholar] [CrossRef]
  177. Sorbiun, M.; Shayegan Mehr, E.; Ramazani, A.; Taghavi Fardood, S. Green Synthesis of Zinc Oxide and Copper Oxide Nanoparticles Using Aqueous Extract of Oak Fruit Hull (Jaft) and Comparing Their Photocatalytic Degradation of Basic Violet 3. Int. J. Environ. Res. 2018, 12, 29–37. [Google Scholar] [CrossRef]
  178. Pakzad, K.; Alinezhad, H.; Nasrollahzadeh, M. Green Synthesis of Ni/Fe3O4 and CuO Nanoparticles Using Euphorbia Maculata Extract as Photocatalysts for the Degradation of Organic Pollutants under UV-Irradiation. Ceram. Int. 2019, 45, 17173–17182. [Google Scholar] [CrossRef]
  179. Wang, G.; Zhao, K.; Gao, C.; Wang, J.; Mei, Y.; Zheng, X.; Zhu, P. Green Synthesis of Copper Nanoparticles Using Green Coffee Bean and Their Applications for Efficient Reduction of Organic Dyes. J. Environ. Chem. Eng. 2021, 9, 105331. [Google Scholar] [CrossRef]
  180. Kushwah, M.; Yadav, R.; Gaur, M.S.; Berlina, A.N. Copper Nanoparticles-Catalysed Reduction of Methylene Blue and High-Sensitive Chemiluminescence Detection of Mercury. Int. J. Environ. Anal. Chem. 2021, 103, 2385–2401. [Google Scholar] [CrossRef]
  181. Subha, V.; Thulasimuthu, E.; Ilangovan, R. Bactericidal Action of Copper Nanoparticles Synthesized from Methanolic Root Extract of Asparagus Racemosus. Mater. Today Proc. 2022, 64, 1761–1767. [Google Scholar] [CrossRef]
  182. Noman, M.; Shahid, M.; Ahmed, T.; Niazi, M.B.K.; Hussain, S.; Song, F.; Manzoor, I. Use of Biogenic Copper Nanoparticles Synthesized from a Native Escherichia Sp. as Photocatalysts for Azo Dye Degradation and Treatment of Textile Effluents. Environ. Pollut. 2020, 257, 113514. [Google Scholar] [CrossRef] [PubMed]
  183. Kumar, B.; Smita, K.; Debut, A.; Cumbal, L. Andean Sacha Inchi (Plukenetiavolubilis L.) Leaf-Mediated Synthesis of Cu2O Nanoparticles: A Low-Cost Approach. Bioengineering 2020, 7, 54. [Google Scholar] [CrossRef]
  184. Surendhiran, S.; Gowthambabu, V.; Balamurugan, A.; Sudha, M.; Senthil Kumar, V.B.; Suresh, K.C. Rapid Green Synthesis of CuO Nanoparticles and Evaluation of Its Photocatalytic and Electrochemical Corrosion Inhibition Performance. Mater. Today Proc. 2021, 47, 1011–1016. [Google Scholar] [CrossRef]
  185. Subha, V.; Kirubanandan, S.; Arulmozhi, M.; Renganathan, S. Green Synthesis of Copper Nanoparticles Using o Dina Woider Gum Extract and Their Effect on Photocatalytic Dye Degradation. Chemist 2018, 91, 9–19. [Google Scholar]
  186. Kerour, A.; Boudjadar, S.; Bourzami, R.; Allouche, B. Eco-Friendly Synthesis of Cuprous Oxide (Cu2O) Nanoparticles and Improvement of Their Solar Photocatalytic Activities. J. Solid. State Chem. 2018, 263, 79–83. [Google Scholar] [CrossRef]
  187. Sinha, T.; Ahmaruzzaman, M. Biogenic Synthesis of Cu Nanoparticles and Its Degradation Behavior for Methyl Red. Mater. Lett. 2015, 159, 168–171. [Google Scholar] [CrossRef]
  188. Al-Jubouri, A.K.; Al-Saadi, N.H.; Kadhim, M.A. Green Synthesis of Copper Nanoparticles from Myrtus Communis Leaves Extract: Characterization, Antioxidant and Catalytic Activity. Iraqi J. Agric. Sci. 2022, 53, 471–486. [Google Scholar] [CrossRef]
  189. Chowdhury, R.; Khan, A.; Rashid, M.H. Green Synthesis of CuO Nanoparticles Using: Lantana Camara Flower Extract and Their Potential Catalytic Activity towards the Aza-Michael Reaction. RSC Adv. 2020, 10, 14374–14385. [Google Scholar] [CrossRef] [Green Version]
  190. Phang, Y.-K.; Aminuzzaman, M.; Akhtaruzzaman, M.; Muhammad, G.; Ogawa, S.; Watanabe, A.; Tey, L.-H. Green Synthesis and Characterization of CuO Nanoparticles Derived from Papaya Peel Extract for the Photocatalytic Degradation of Palm Oil Mill Effluent (POME). Sustainability 2021, 13, 1–15. [Google Scholar] [CrossRef]
  191. Ssekatawa, K.; Byarugaba, D.K.; Angwe, M.K.; Wampande, E.M.; Ejobi, F.; Nxumalo, E.; Maaza, M.; Sackey, J.; Kirabira, J.B. Phyto-Mediated Copper Oxide Nanoparticles for Antibacterial, Antioxidant and Photocatalytic Performances. Front. Bioeng. Biotechnol. 2022, 10, 820218. [Google Scholar] [CrossRef] [PubMed]
  192. Haseena, S.; Shanavas, S.; Duraimurugan, J.; Ahamad, T.; Alshehri, S.M.; Acevedo, R.; Jayamani, N. Investigation on Photocatalytic and Antibacterial Ability of Green Treated Copper Oxide Nanoparticles Using Artabotrys Hexapetalus and Bambusa Vulgaris Plant Extract. Mater. Res. Express 2019, 6. [Google Scholar] [CrossRef]
  193. Rafique, M.; Shafiq, F.; Ali Gillani, S.S.; Shakil, M.; Tahir, M.B.; Sadaf, I. Eco-Friendly Green and Biosynthesis of Copper Oxide Nanoparticles Using Citrofortunella Microcarpa Leaves Extract for Efficient Photocatalytic Degradation of Rhodamin B Dye Form Textile Wastewater. Optik 2020, 208, 164053. [Google Scholar] [CrossRef]
  194. Saif, S.; Adil, S.F.; Khan, M.; Hatshan, M.R.; Khan, M.; Bashir, F. Adsorption Studies of Arsenic(V) by Cuo Nanoparticles Synthesized by Phyllanthus Emblica Leaf-Extract-Fueled Solution Combustion Synthesis. Sustainability 2021, 13, 2017. [Google Scholar] [CrossRef]
  195. Borah, R.; Saikia, E.; Bora, S.J.; Chetia, B. On-Water Synthesis of Phenols Using Biogenic Cu2O Nanoparticles without Using H2O2. RSC Adv. 2016, 6, 100443–100447. [Google Scholar] [CrossRef]
  196. Priya, D.D.; Roopan, S.M.; Singh, S.; Bansal, J.; Shanavas, S.; Khan, M.R.; Al-Dhabi, N.A.; Arasu, M.V.; Duraipandiyan, V. Phyto-Synthesis of CuO Nano-Particles and Its Catalytic Application in C-S Bond Formation. Mater. Lett. 2020, 266, 127486. [Google Scholar] [CrossRef]
  197. Veisi, H.; Hemmati, S.; Javaheri, H. N-Arylation of Indole and Aniline by a Green Synthesized CuO Nanoparticles Mediated by Thymbra Spicata Leaves Extract as a Recyclable and Heterogeneous Nanocatalyst. Tetrahedron Lett. 2017, 58, 3155–3159. [Google Scholar] [CrossRef]
  198. Veisi, H.; Karmakar, B.; Tamoradi, T.; Hemmati, S.; Hekmati, M.; Hamelian, M. Biosynthesis of CuO Nanoparticles Using Aqueous Extract of Herbal Tea (Stachys Lavandulifolia) Flowers and Evaluation of Its Catalytic Activity. Sci. Rep. 2021, 11, 1–13. [Google Scholar] [CrossRef]
  199. Singh, H.P.; Sharma, S.; Sharma, S.K.; Sharma, R.K. Biogenic Synthesis of Metal Nanocatalysts Using Mimosa Pudica Leaves for Efficient Reduction of Aromatic Nitrocompounds. RSC Adv. 2014, 4, 37816–37825. [Google Scholar] [CrossRef]
  200. Mohan, S.; Singh, Y.; Verma, D.K.; Hasan, S.H. Synthesis of CuO Nanoparticles through Green Route Using Citrus Limon Juice and Its Application as Nanosorbent for Cr(VI) Remediation: Process Optimization with RSM and ANN-GA Based Model. Process. Saf. Environ. Prot. 2015, 96, 156–166. [Google Scholar] [CrossRef]
  201. Singh, J.; Kumar, V.; Kim, K.-H.; Rawat, M. Biogenic Synthesis of Copper Oxide Nanoparticles Using Plant Extract and Its Prodigious Potential for Photocatalytic Degradation of Dyes. Environ. Res. 2019, 177, 108569. [Google Scholar] [CrossRef] [PubMed]
  202. Nasrollahzadeh, M.; Maham, M.; Mohammad Sajadi, S. Green Synthesis of CuO Nanoparticles by Aqueous Extract of Gundelia Tournefortii and Evaluation of Their Catalytic Activity for the Synthesis of N-Monosubstituted Ureas and Reduction of 4-Nitrophenol. J. Colloid. Interface Sci. 2015, 455, 245–253. [Google Scholar] [CrossRef] [PubMed]
  203. Nasrollahzadeh, M.; Mohammad Sajadi, S.; Rostami-Vartooni, A. Green Synthesis of CuO Nanoparticles by Aqueous Extract of Anthemis Nobilis Flowers and Their Catalytic Activity for the A3 Coupling Reaction. J. Colloid. Interface Sci. 2015, 459, 183–188. [Google Scholar] [CrossRef]
  204. Nasrollahzadeh, M.; Sajadi, S.M.; Rostami-Vartooni, A.; Hussin, S.M. Green Synthesis of CuO Nanoparticles Using Aqueous Extract of Thymus vulgaris L. Leaves and Their Catalytic Performance for N-Arylation of Indoles and Amines. J. Colloid. Interface Sci. 2016, 466, 113–119. [Google Scholar] [CrossRef]
  205. Ghosh, M.K.; Sahu, S.; Gupta, I.; Ghorai, T.K. Green Synthesis of Copper Nanoparticles from an Extract OfJatropha Curcasleaves: Characterization, Optical Properties, CT-DNA Binding and Photocatalytic Activity. RSC Adv. 2020, 10, 22027–22035. [Google Scholar] [CrossRef]
  206. Chandraker, S.K.; Lal, M.; Ghosh, M.K.; Tiwari, V.; Ghorai, T.K.; Shukla, R. Green Synthesis of Copper Nanoparticles Using Leaf Extract of Ageratum Houstonianum Mill. and Study of Their Photocatalytic and Antibacterial Activities. Nano Express 2020, 1, 010033. [Google Scholar] [CrossRef]
  207. Nasrollahzadeh, M.; Sajadi, S.M.; Khalaj, M. Green Synthesis of Copper Nanoparticles Using Aqueous Extract of the Leaves of Euphorbia Esula L and Their Catalytic Activity for Ligand-Free Ullmann-Coupling Reaction and Reduction of 4-Nitrophenol. RSC Adv. 2014, 4, 47313–47318. [Google Scholar] [CrossRef]
  208. Nasrollahzadeh, M.; Mohammad Sajadi, S. Green Synthesis of Copper Nanoparticles Using Ginkgo biloba L. Leaf Extract and Their Catalytic Activity for the Huisgen [3+2] Cycloaddition of Azides and Alkynes at Room Temperature. J. Colloid. Interface Sci. 2015, 457, 141–147. [Google Scholar] [CrossRef]
  209. Nasrollahzadeh, M.; Sajadi, S.M.; Mirzaei, Y. An Efficient One-Pot Synthesis of 1,4-Disubstituted 1,2,3-Triazoles at Room Temperature by Green Synthesized Cu NPs Using Otostegia Persica Leaf Extract. J. Colloid. Interface Sci. 2016, 468, 156–162. [Google Scholar] [CrossRef] [PubMed]
  210. Nasrollahzadeh, M.; Momeni, S.S.; Sajadi, S.M. Green Synthesis of Copper Nanoparticles Using Plantago Asiatica Leaf Extract and Their Application for the Cyanation of Aldehydes Using K 4 Fe(CN) 6. J. Colloid. Interface Sci. 2017, 506, 471–477. [Google Scholar] [CrossRef]
  211. Ullah, H.; Ullah, Z.; Fazal, A.; Irfan, M. Use of Vegetable Waste Extracts for Controlling Microstructure of CuO Nanoparticles: Green Synthesis, Characterization, and Photocatalytic Applications. J. Chem. 2017, 2017, 2721798. [Google Scholar] [CrossRef] [Green Version]
  212. Nazar, N.; Bibi, I.; Kamal, S.; Iqbal, M.; Nouren, S.; Jilani, K.; Umair, M.; Ata, S. Cu Nanoparticles Synthesis Using Biological Molecule of P. Granatum Seeds Extract as Reducing and Capping Agent: Growth Mechanism and Photo-Catalytic Activity. Int. J. Biol. Macromol. 2018, 106, 1203–1210. [Google Scholar] [CrossRef]
  213. Sharma, P.; Pant, S.; Poonia, P.; Kumari, S.; Dave, V.; Sharma, S. Green Synthesis of Colloidal Copper Nanoparticles Capped with Tinospora Cordifolia and Its Application in Catalytic Degradation in Textile Dye: An Ecologically Sound Approach. J. Inorg. Organomet. Polym. Mater. 2018, 28, 2463–2472. [Google Scholar] [CrossRef]
  214. Olajire, A.A.; Ifediora, N.F.; Bello, M.D.; Benson, N.U. Green Synthesis of Copper Nanoparticles Using Alchornea Laxiflora Leaf Extract and Their Catalytic Application for Oxidative Desulphurization of Model Oil. Iran. J. Sci. Technol. Trans. A Sci. 2018, 42, 1935–1946. [Google Scholar] [CrossRef]
  215. Bagherzadeh, M.; Safarkhani, M.; Ghadiri, A.M.; Kiani, M.; Fatahi, Y.; Taghavimandi, F.; Daneshgar, H.; Abbariki, N.; Makvandi, P.; Varma, R.S.; et al. Bioengineering of CuO Porous (Nano)Particles: Role of Surface Amination in Biological, Antibacterial, and Photocatalytic Activity. Sci. Rep. 2022, 12, 15351. [Google Scholar] [CrossRef]
  216. Ismail, M.; Gul, S.; Khan, M.I.; Khan, M.A.; Asiri, A.M.; Khan, S.B. Green Synthesis of Zerovalent Copper Nanoparticles for Efficient Reduction of Toxic Azo Dyes Congo Red and Methyl Orange. Green. Process. Synth. 2019, 8, 135–143. [Google Scholar] [CrossRef]
  217. Kayalvizhi, S.; Sengottaiyan, A.; Selvankumar, T.; Senthilkumar, B.; Sudhakar, C.; Selvam, K. Eco-Friendly Cost-Effective Approach for Synthesis of Copper Oxide Nanoparticles for Enhanced Photocatalytic Performance. Optik 2020, 202, 163507. [Google Scholar] [CrossRef]
  218. Devi, T.B.; Ahmaruzzaman, M. Facile Preparation of Copper Nanaoparticles Using Coccinia Grandis Fruit Extract and Its Application towards the Reduction of Toxic Nitro Compound. Mater. Today Proc. 2018, 5, 2098–2104. [Google Scholar] [CrossRef]
  219. Devi, T.B.; Ahmaruzzaman, M. Removal of Perilous Nitrocompound from Aqueous Phase Using Biogenic Copper Nanoparticles as a Catalyst. Indian. J. Chem. Technol. 2018, 25, 561–564. [Google Scholar]
  220. Buazar, F.; Sweidi, S.; Badri, M.; Kroushawi, F. Biofabrication of Highly Pure Copper Oxide Nanoparticles Using Wheat Seed Extract and Their Catalytic Activity: A Mechanistic Approach. Green. Process. Synth. 2019, 8, 691–702. [Google Scholar] [CrossRef]
  221. Siddiqi, K.S.; Rashid, M.; Rahman, A.; Tajuddin; Husen, A.; Rehman, S. Green Synthesis, Characterization, Antibacterial and Photocatalytic Activity of Black Cupric Oxide Nanoparticles. Agric. Food Secur. 2020, 9, 17. [Google Scholar] [CrossRef]
  222. Karuppannan, S.K.; Ramalingam, R.; Mohamed Khalith, S.B.; Dowlath, M.J.H.; Darul Raiyaan, G.I.; Arunachalam, K.D. Characterization, Antibacterial and Photocatalytic Evaluation of Green Synthesized Copper Oxide Nanoparticles. Biocatal. Agric. Biotechnol. 2021, 31, 101904. [Google Scholar] [CrossRef]
  223. Mahmoud, A.E.D.; Al-Qahtani, K.M.; Alflaij, S.O.; Al-Qahtani, S.F.; Alsamhan, F.A. Green Copper Oxide Nanoparticles for Lead, Nickel, and Cadmium Removal from Contaminated Water. Sci. Rep. 2021, 11, 12547. [Google Scholar] [CrossRef]
  224. Roy, K.; Ghosh, C.K.; Sarkar, C.K. Degradation of Toxic Textile Dyes and Detection of Hazardous Hg2+ by Low-Cost Bioengineered Copper Nanoparticles Synthesized Using Impatiens Balsamina Leaf Extract. Mater. Res. Bull. 2017, 94, 257–262. [Google Scholar] [CrossRef]
  225. Khani, R.; Roostaei, B.; Bagherzade, G.; Moudi, M. Green Synthesis of Copper Nanoparticles by Fruit Extract of Ziziphus Spina-Christi (L.) Willd.: Application for Adsorption of Triphenylmethane Dye and Antibacterial Assay. J. Mol. Liq. 2018, 255, 541–549. [Google Scholar] [CrossRef]
  226. Roselyn Maheo, A.; Scholastica Mary Vithiya, B.; Augustine Arul Prasad, T.; Tamizhdurai, P.; Mangesh, V.L. Biosynthesis, Characterization, Biological and Photo Catalytic Investigations of Elsholtzia Blanda and Chitosan Mediated Copper Oxide Nanoparticles: Biosynthesis, Characterization, Biological and Photo Catalytic Investigations. Arab. J. Chem. 2022, 15, 103661. [Google Scholar] [CrossRef]
  227. Raina, S.; Roy, A.; Bharadvaja, N. Degradation of Dyes Using Biologically Synthesized Silver and Copper Nanoparticles. Environ. Nanotechnol. Monit. Manag. 2020, 13, 100278. [Google Scholar] [CrossRef]
  228. Prabhu, S.; Thangadurai, T.D.; Bharathy, P.V.; Kalugasalam, P. Investigation on the Photocatalytic and Antibacterial Activities of Green Synthesized Cupric Oxide Nanoparticles Using Clitoria Ternatea. Iran. J. Catal. 2022, 12, 1–11. [Google Scholar] [CrossRef]
  229. Tamuly, C.; Saikia, I.; Hazarika, M.; Das, M.R. Reduction of Aromatic Nitro Compounds Catalyzed by Biogenic CuO Nanoparticles. RSC Adv. 2014, 4, 53229–53236. [Google Scholar] [CrossRef]
  230. Alinezhad, H.; Pakzad, K. C-S Cross-Coupling Reaction Using Novel and Green Synthesized CuO Nanoparticles Assisted by Euphorbia Maculata Extract. Appl. Organomet. Chem. 2019, 33, e5144. [Google Scholar] [CrossRef]
  231. Batool, M.; Qureshi, M.Z.; Hashmi, F.; Mehboob, N.; Daoush, W.M. Adsorption of Congo Red (Acid Red 28) Azodye on Biosynthesized Copper Oxide Nanoparticles. Asian J. Chem. 2019, 31, 707–713. [Google Scholar] [CrossRef]
  232. Reddeppa, M.; Srujana, P.; Chenna Krishna Reddy, R.; Reddy, T.V.; Rani, T.S. Phyto-Mediated Green Synthesis of Copper Nanoparticles: Study of Catalytic Performance and Antibacterial Activity. Rasayan J. Chem. 2020, 13, 1885–1893. [Google Scholar] [CrossRef]
  233. Nasrollahzadeh, M.; Sajadi, S.M.; Maham, M.; Dasmeh, H.R. In Situ Green Synthesis of Cu Nanoparticles Supported on Natural Natrolite Zeolite for the Reduction of 4-Nitrophenol, Congo Red and Methylene Blue. IET Nanobiotechnol. 2017, 11, 538–545. [Google Scholar] [CrossRef]
  234. Hussain, N.; Gogoi, P.; Azhaganand, V.K.; Shelke, M.V.; Das, M.R. Green Synthesis of Stable Cu(0) Nanoparticles onto Reduced Graphene Oxide Nanosheets: A Reusable Catalyst for the Synthesis of Symmetrical Biaryls from Arylboronic Acids under Base-Free Conditions. Catal. Sci. Technol. 2015, 5, 1251–1260. [Google Scholar] [CrossRef]
  235. Pary, F.F.; Addanki Tirumala, R.T.; Andiappan, M.; Nelson, T.L. Copper(i) Oxide Nanoparticle-Mediated C-C Couplings for Synthesis of Polyphenylenediethynylenes: Evidence for a Homogeneous Catalytic Pathway. Catal. Sci. Technol. 2021, 11, 2414–2421. [Google Scholar] [CrossRef]
  236. Ley, S.V.; Thomas, A.W. Modern Synthetic Methods for Copper-Mediated C(Aryl)-O, C(Aryl)-N, and C(Aryl)-S Bond Formation. Angew. Chem. Int. Ed. 2003, 42, 5400–5449. [Google Scholar] [CrossRef]
  237. Barman, K.; Dutta, P.; Chowdhury, D.; Baruah, P.K. Green Biosynthesis of Copper Oxide Nanoparticles Using Waste Colocasia Esculenta Leaves Extract and Their Application as Recyclable Catalyst Towards the Synthesis of 1,2,3-Triazoles. Bionanoscience 2021, 11, 189–199. [Google Scholar] [CrossRef]
  238. Borah, R.; Saikia, E.; Bora, S.J.; Chetia, B. Banana Pulp Extract Mediated Synthesis of Cu2O Nanoparticles: An Efficient Heterogeneous Catalyst for the Ipso-Hydroxylation of Arylboronic Acids. Tetrahedron Lett. 2017, 58, 1211–1215. [Google Scholar] [CrossRef]
  239. Wang, B.; Xu, M.; Chi, C.; Wang, C.; Mneg, D. Degradation of Methyl Orange Using Dielectric Barrier Discharge Water Falling Film Reactor. J. Adv. Oxid. Technol. 2017, 20, 20170021. [Google Scholar] [CrossRef]
  240. Yadav, S.; Chauhan, M.; Mathur, D.; Jain, A.; Malhotra, P. Sugarcane Bagasse-Facilitated Benign Synthesis of Cu2O Nanoparticles and Its Role in Photocatalytic Degradation of Toxic Dyes: A Trash to Treasure Approach. Environ. Dev. Sustain. 2021, 23, 2071–2091. [Google Scholar] [CrossRef]
  241. Yasin, A.; Fatima, U.; Shahid, S.; Mansoor, S.; Inam, H.; Javed, M.; Iqbal, S.; Alrbyawi, H.; Somaily, H.H.; Pashameah, R.A.; et al. Fabrication of Copper Oxide Nanoparticles Using Passiflora edulis Extract for the Estimation of Antioxidant Potential and Photocatalytic Methylene Blue Dye Degradation. Agronomy 2022, 12, 2315. [Google Scholar] [CrossRef]
  242. Zhuang, X.; Kang, Y.; Zhao, L.; Guo, S. Design and Synthesis of Copper Nanoparticles for the Treatment of Human Esophageal Cancer: Introducing a Novel Chemotherapeutic Supplement. J Exp. Nanosci. 2022, 17, 274–284. [Google Scholar] [CrossRef]
Figure 1. VOSviewer visualization of a term co-occurrence network based on the title and abstract fields, narrowing 30 keywords and 29 co-occurrences.
Figure 1. VOSviewer visualization of a term co-occurrence network based on the title and abstract fields, narrowing 30 keywords and 29 co-occurrences.
Molecules 28 04838 g001
Figure 2. Antioxidant effect of CuNPs synthesized with biological extracts; (A) antioxidant activity (%), (B) IC50 (µg/mL), and (C) frequency of salt and concentrations of articles reviewed.
Figure 2. Antioxidant effect of CuNPs synthesized with biological extracts; (A) antioxidant activity (%), (B) IC50 (µg/mL), and (C) frequency of salt and concentrations of articles reviewed.
Molecules 28 04838 g002
Figure 3. Antitumoral effect (IC50) of CuNPs synthesized with biological extracts in (A) breast, (B) cervical, and (C) lung cancer cell lines in vitro.
Figure 3. Antitumoral effect (IC50) of CuNPs synthesized with biological extracts in (A) breast, (B) cervical, and (C) lung cancer cell lines in vitro.
Molecules 28 04838 g003
Figure 4. (A) Frequency (%) of papers reviewed of reaction time (min) on synthesis/catalysis/degradation/reduction approach; (B) Frequency (%) of reaction time (min) on dye removal approach; (C) Salt type frequency on nanoparticles synthesis (%), and (D) Salt concentration (M) frequency on nanoparticles synthesis.
Figure 4. (A) Frequency (%) of papers reviewed of reaction time (min) on synthesis/catalysis/degradation/reduction approach; (B) Frequency (%) of reaction time (min) on dye removal approach; (C) Salt type frequency on nanoparticles synthesis (%), and (D) Salt concentration (M) frequency on nanoparticles synthesis.
Molecules 28 04838 g004
Figure 5. Photocatalytic activity of MB, RhB, and CR using CuNPs under similar conditions. (A) Step 1: degradation process in dark conditions, (B) step 2: degradation process under UV Light, (C) spectrophotometer UV-Vis, and (D) synthesis of MB. Step 3: spectrophotometer UV-Vis.
Figure 5. Photocatalytic activity of MB, RhB, and CR using CuNPs under similar conditions. (A) Step 1: degradation process in dark conditions, (B) step 2: degradation process under UV Light, (C) spectrophotometer UV-Vis, and (D) synthesis of MB. Step 3: spectrophotometer UV-Vis.
Molecules 28 04838 g005
Figure 6. PRISMA (preferred reporting items for systematic reviews and meta-analysis).
Figure 6. PRISMA (preferred reporting items for systematic reviews and meta-analysis).
Molecules 28 04838 g006
Table 1. The antioxidant effect of biogenically synthetized CuNPs, salt, concentration, and biological source.
Table 1. The antioxidant effect of biogenically synthetized CuNPs, salt, concentration, and biological source.
PlantSalt Concentration
(mM)
Part of the PlantResultsReference
Antioxidant Activity (%) aIC50 (µg/mL)
Copper acetate
Azadirachta indica200Leaves--[19]
Eclipta prostrata3Leaves53-[20]
Monotheca buxifolia50Leaves53.40-[21]
Borreria hispida10Plant-0.6[22]
Cissus vitiginea-Leaves-45.2[23]
Copper nitrate
Eryngium caucasicum10Leaves40.02-[24]
Juglans regia-Leaves78.80–90-[25]
Solanum nigrum100Leaves90-[26]
Galeopsidis herba-Extract-4.12[27]
Allium noeanum0.1Leaves-160[28]
Abutilon indicum-Leaves-84[29]
Allium sativum100Extract-40[30]
Plectranthus amboinicus100Leaves-40.10[31]
Solanum nigrum10Leaves-60[26]
Tinospora cordifolia-Leaves-566[32]
Copper sulfate
Ocimum basilicum100Extract--[33]
Sargassum longifolium10Seaweed--[34]
Thymbra spicata1Leaves--[35]
Berberis thunbergii200Leaves--[17]
Achillea nobilis100Branch44–48-[36]
Prunus mahaleb50Leaves15.90-[37]
Cissus vitiginea10Leaves21-[38]
Persea americana-Seed23-[16]
Mangifera indica3Leaves36.9-[39]
Cissus arnotiana10Leaves21-[14]
Protoparmeliopsis muralis100Lichen31.02-[18]
Falcaria vulgaris40Leaves-190[40]
Malus domestica1000Leaves-45.90[41]
Cystoseira trinodis (algae)---543[42]
Salvia hispanica100Leaves--[43]
Zingiber officinale280 Rhizome81-[44]
Moringa oleifera Leaves60–70-[45]
Artemisia haussknechtii100Leaves74.45-[46]
Actin- omycetes (bacteria)20---[47]
Actin- omycetes (bacteria)20---[47]
Laurus nobilis1Leaves--[48]
Triticum aestivum40Herb--[49]
Pleurotus ostreatus2Seed--[50]
Urtica dioica10Leaves--[51]
Pleurotus ostreatus2Biomass--[50]
Copper chloride
Koelreuteria paniculata10Seeds14.54-[52]
Camellia sinensis250Leaves-14[53]
a Percentage compared to standards antioxidants.
Table 3. The antibacterial effect of biogenically synthetized CuNPs, salt, concentration, and biological source.
Table 3. The antibacterial effect of biogenically synthetized CuNPs, salt, concentration, and biological source.
Natural Extract SourcePart of the PlantSize (nm)ShapeNP
Concentration
Micro-OrganismsReference
Copper acetate
Sargassum swartziiWhole32Spherical25 μg/mLV. parahaemolyticus[120]
Averrhoa carambolaFruit15Spherical, square, and hexagonal20 μg/mLS. aureus, Bacillius spp., Pseudomona spp.[118]
Cylindrospermum stagnaleBiomass100Spherical8 mMC. albicans, K. pneumonia, E. cloacae, P. aeruginosa, E. coli[76]
Penicillium chrysogenumFiltrate10–190Crystalline50 μg/mLK. oxytoca, E. coli, S. aureus, B. cereus[121]
Ocimum tenuiflorumLeaves12–44Spherical3125–12,500 μg/mLB. subtilis, S. aureus, E.coli[87]
Aloe-veraLeaves45–95Elliptical1562 μg/mLB. subtilis, S. aureus, E.coli[122]
Camellia SinensisLeaves25–32Crystalline1000 µg/mLS. aureus, B. subtilis, E. coli, K. pneumonia[123]
Mimosa hamataFlower40-0.1 g/mLE. coli and B. cereus[124]
Polyalthia longifoliaLeaves50–60Quasi-spherical100–1000 µg/mLS. pyogenes, S. aureus, E. coli, and P. aeruginosa, E. floccosum, C. albicans, A. niger, A. clavatus[125]
Botryococcus brauniiBiomass10–70Spherical and cubical250 µg/mLP. aeruginosa, E. coli, K. pneumoniae, S. aureus, F. oxysporum[126]
Clematis orientalisLeaves13–53Crystalline0.25 MS. aureus, B. subtilis, E. coli, P. aeruginosa, K. pneumoniae[127]
Copper nitrate
Allium sativumRoot20–40Circular150 μg/mLE. coli, S. aureus, B. subtilis, S. pyogenes, P. aeruginosa, K. pneumoniae[30]
Zingiber officinale and Allium sativumRoot20–45Spherical, agglomerated1000 μg/mLMDR S. aureus[111]
Tinospora cordifoliaLeaves10Spherical1001 μg/mLK. aerogenes, E. coli, P. desmolyticum, S. aureus[32]
Withania somniferaRoot5–9Spherical100 μg/mLE. coli and S. aureus[128]
Morus alba L. Fruit50–200Spherical and non-regular2500 μg/mLE.coli and L. monocytogenes[129]
Alpinia galangalRhizome20–60Irregular spherical10 mg/mLS. mutans, B. cereus, P. vulgaris, S. marcences[130]
Cordia sebestenaFlower20–35Spherical75 µg/mLB. subtilis, S. aureus and E. coli, K. pneumoniae[131]
Cassia fistula and Melia azedarachLeaves43.8Spherical0.5 mg/mLK. pneumonia and H. pylori[132]
Gloriosa superba L. Leaves5–10Spherical1000 µg/mLK. aerogenes, P. desmolyticum, E. coli, S. aureus[133]
Solanum nigrumLeaves25Spherical100 μg/mLB. subtilis, S. saprohyticus, E. coli, P. aeruginosa[26]
Rauvolfia serpentinaLeaves10–20Crystalline1000 μg/mLE. coli, P. desmolyticum, S. aureus[134]
Aloe veraLeaves20Crystalline100 μg/mLA. hydrophila, P. fluorescens, F. branchiophilum[135]
Hibiscus cannabinusLeaves10–40Crystalline5–31 mg/mLB. cereus, S. aureus, E. coli, K. pneumoniae[136]
Capsicum frutescensLeaves20–40Spherical and rectangular150 μg/mLK. pneumoniae, B. anthracis, L. monocytogenes[137]
Eryngium caucasicumLeaves40Spherical30–100 μg/mLE. coli, S. typhimurium, B. cereus and S. aureus[24]
MentheBiomass22–25Cubic, crystalline250 µg/mLE. coli, B. subtilis[138]
Saccharum officinarumStem5–140Spherical100 µg/mLE. coli, P. aeruginosa, S. aureus, B. subtilis[139]
Catha edulisLeaves28Crystalline40 mg/mLS. aureus, S. pyogenes, E. coli, K. pneumonia[140]
Solanum tuberosumTuber54Spherical200–1000 µg/mLB. cereus, S sonnei, S. aureus, S. epidermidis, Enterococcus spp., P. aeruginosa, E. coli[141]
Madhuca longifoliaFlower/seeds30–100Spherical10 mg/mLE.coli, B. subtilis, S. aureus[142]
Opuntia ficus-indicaLeaves3–10Spherical100 µg/mLE. coli[143]
Copper sulfate
Falcaria vulgarisLeaf20–25Spherical2 mg/mLS. pneumonia, B. subtilis, C. guilliermondii, C. krusei[40]
4 mg/mLP. aeruginosa, S. aureus, C. albicans, C. glabrata
8 mg/mLS. typhimurium, E. coli
4 mg/mLS. pneumonia, B. subtilis, C. guilliermondii, C. kruse
8 mg/mLP. aeruginosa, S. aureus, C. albicans, C. glabrata
16 mg/mLS. typhimurium, E. coli
Syzygium alternifoliumBarks50–100Spherical20 μg/mLB. subtilis, S. aureus, E.coli, K. pneumonia, P. vulgaris, P. aeruginosa, S. typhimurium[81]
A. solani, A. flavus, A. niger, P. chrysogenum, T. harzianum
80 μg/mLBacteria and fungi tested
Cissus vitigineaLeaves5–20Spherical75 μg/mLE. coli, Enterococcus sp., Proteus sp., Klebsiella sp.[38]
Citrus AurantifoliaLeaves10Spherical-crystalline50 μg/mLK. pneumoniae, S. aureus[102]
Illicium verumFruit150–220Undefined100 μg/mLS. aureus[102]
Myristica fragransFruit210–270Undefined100 μg/mLS. aureus[102]
Justicia gendarussaLeaves50–100Flower-shaped75 μg/mLE. coli, S. aureus[144]
Ruellia tuberosaLeaves82Spherical, cylindrical, and cubical75 μg/mLS. aureus, K. pneumoniae, E. coli[145]
Solanum lycopersicumLeaves20–40Spherical300 μg/mLB. subtilis, S. aureus, E.coli[146]
Sesbania aculeataLeaves--40 μg/mLP. destructive and C. lunata[147]
Allium saralicumLeaves45–50Spherical8 mg/mLC. albicans, C. glabrata, C. krusei, C. guilliermondii, P. aeruginosa, E.coli, B. subtilis, S. aureus, S. typhimurium, S. pneumoniae[148]
Cystoseira myrica, Sargassum latifolium and Padina australisLeaves40Spherical250 μg/mLE. coli and S. aureus[149]
Tabernaemontana divaricateLeaves46Spherical25 μg/mLE. coli[150]
Acalypha indicaLeaves26Spherical25 μg/mlA. indica and C. albicans[86]
Citrus medicaFruits10–60-100 mME. coli, P. acne, K. pneumoniae, S. typhi, and P. aeruginosa, F. oxysporum, F. graminearum and F. culmorum[151]
Achillea NobilisFlower15–25Hexagonal50 μg/mLE. coli and S. aureus[36]
Achyranthes asperaLeaves95Spherical250 mMS. aureus and Gram-negative P. aeruginosa[152]
Heliconia psittacorumFlower12Spherical50 μg/mLS. aureus, P. putida, E. coli[153]
Cedrus deodaraLeaves-Spherical150 μg/mLE. Coli, S. Aureus, S. Enterica, L. Monocytogenes[154]
Sargassum longifoliumBiomass40–60Spherical100 μg/mLA. hydrophilla, V. harveyi, V. parahaemolyticus, S. marcescens[34]
Cissus arnotianaLeaves60–90Spherical50 μg/mLE. coli, Streptococcus sp., Rhizobium sp., Klebsiella sp.[14]
Citrus aurantifoliaLeaves22Crystalline150 μg/mLS. aureus and E. coli[155]
Pterolobium hexapetalumLeaves10–50Crystalline50 µg/mLS. aureus, B. subtilis, E. coli[156]
Convolvulus percicus L. Leaves15–30Crystalline6.25 μg/mLS. aureus, E. coli[157]
Cystoseira trinodisBiomass6–7.8Crystalline10 μg/mLE.coli, E. faecalis, S. typhimurium, S. aureus, B. subtilis, S. faecalis[42]
Azadirachta indicaFlower5Spherical40 μg/mLE. faecalis, P. mirabilis, K. pneumonia, S. aureus.[158]
Mentha pulegiumLeaves21–48Spherical1000 μg/mLS. aureus, B. cereus, E.coli, K. pneumoniae[159]
Brassica oleracea var. capitata f. rubraLeaves77.5Spherical50 μg/mLE. coli, S. aureus[160]
Passiflora foetida 24.5Crystalline1000 µg/mLE.coli, S. typhimurium, A. aceti[161]
Thymbra spicataLeaves21–26.8Spherical100 µg/mLB. cereus, S. aureus, E.coli, S. typhimurium[35]
Prunus mahaleb L. Whole10–50Spherical0.5–2 mg/mLS. aureus, K. pneumonia, P. aeruginosa, E. coli, P. aeruginosa[37]
Artemisia haussknechtiiLeaves35Spherical0.1 ME. coli, S. aureus[46]
Copper chloride
Tamarindus indica L. Fruit and leaves50–100Spherical60 μL NPL. acidophilus[162]
60 μL NPE. coli
40 μL NPS. typhi
Anacardium occidentaleShell100 40 μg/mLB. linens, P. acnes, B. cereus, S. epidermidis[163]
Ficus caricaLeaves41.5Spherical200 μg/mLCandida spp., Aspergillus spp., S aureus, A. baumanii[164]
Tinospora cardifoliaLeaves63–143Spherical175 μg/mLS. aureus and E. coli[165]
Vitex negundoRoot40–60Spherical, cubic and hexagonal500 μg/mLB. subtilis, S. aureus, E. coli, P. aeruginosa[166]
Cardiospermum halicacabumLeaves40Hexagonal50 µg/mLP. aeruginosa, E. coli, S. aureus[98]
Aspergillus nigerBiomass23–199Crystalline2.5 mg/mLP. aeruginosa, E. faecalis, E. coli, K. pneumonia, P. vulgari, S. aureus, C. albicans, A. niger[167]
Brassica oleracea, Solanum tuberosum, Pisum sativumPeels32.5, 40.75, 47.2Spherical and cubical45 µg/mLP. aeruginosa, E. coli, B. subtilus, S. aureus[168]
Vaccinium myrtillus L. Fruit2–10Spherical3.125 mME. coli, C. albicans, S. cerevisiae[169]
Table 5. AMSTAR criteria table.
Table 5. AMSTAR criteria table.
Item
1Was an ‘‘a priori’’ design provided?
2Was there duplicate study selection and data extraction?
3Was a comprehensive literature search performed?
4Was the status of publication (i.e., grey literature) used as an inclusion criterion?
5Was a list of studies (included and excluded) provided?
6Were the characteristics of the included studies provided?
7Was the scientific quality of the included studies assessed and documented?
8Was the scientific quality of the included studies used appropriately in formulating the conclusions?
9Were the methods used to combine the findings of studies appropriate?
10Was the likelihood of publication bias assessed?
11Were potential conflicts of interest included?
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Luque-Jacobo, C.M.; Cespedes-Loayza, A.L.; Echegaray-Ugarte, T.S.; Cruz-Loayza, J.L.; Cruz, I.; de Carvalho, J.C.; Goyzueta-Mamani, L.D. Biogenic Synthesis of Copper Nanoparticles: A Systematic Review of Their Features and Main Applications. Molecules 2023, 28, 4838. https://doi.org/10.3390/molecules28124838

AMA Style

Luque-Jacobo CM, Cespedes-Loayza AL, Echegaray-Ugarte TS, Cruz-Loayza JL, Cruz I, de Carvalho JC, Goyzueta-Mamani LD. Biogenic Synthesis of Copper Nanoparticles: A Systematic Review of Their Features and Main Applications. Molecules. 2023; 28(12):4838. https://doi.org/10.3390/molecules28124838

Chicago/Turabian Style

Luque-Jacobo, Cristina M., Andrea L. Cespedes-Loayza, Talia S. Echegaray-Ugarte, Jacqueline L. Cruz-Loayza, Isemar Cruz, Júlio Cesar de Carvalho, and Luis Daniel Goyzueta-Mamani. 2023. "Biogenic Synthesis of Copper Nanoparticles: A Systematic Review of Their Features and Main Applications" Molecules 28, no. 12: 4838. https://doi.org/10.3390/molecules28124838

Article Metrics

Back to TopTop