Next Article in Journal
A Comprehensive Review of the Current Knowledge of Chlorfenapyr: Synthesis, Mode of Action, Resistance, and Environmental Toxicology
Previous Article in Journal
Phenopyrrolizins A and B, Two Novel Pyrrolizine Alkaloids from Marine-Derived Actinomycetes Micromonospora sp. HU138
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Recent Updates on Multifunctional Nanomaterials as Antipathogens in Humans and Livestock: Classification, Application, Mode of Action, and Challenges

by
Samreen Sadiq
1,
Iltaf Khan
2,
Zhenyu Shen
1,
Mengdong Wang
1,
Tao Xu
1,
Sohail Khan
3,
Xuemin Zhou
1,
Ali Bahadur
4,
Madiha Rafiq
5,
Sumreen Sohail
6 and
Ping Wu
1,*
1
School of Biotechnology, Jiangsu University of Science and Technology, Zhenjiang 212100, China
2
School of Environmental & Chemical Engineering, Jiangsu University of Science and Technology, Zhenjiang 212100, China
3
Department of Pharmacy, University of Swabi, Khyber Pakhtunkhwa 94640, Pakistan
4
College of Science, Mathematics, and Technology, Wenzhou-Kean University, Wenzhou 325060, China
5
Key Laboratory for Preparation and Application of Ordered Structural Materials of Guangdong Province, Department of Chemistry, Shantou University, Shantou 515063, China
6
Department of Information Technology, Careerera, Beltsville, MD 20705, USA
*
Author to whom correspondence should be addressed.
Molecules 2023, 28(22), 7674; https://doi.org/10.3390/molecules28227674
Submission received: 10 October 2023 / Revised: 11 November 2023 / Accepted: 14 November 2023 / Published: 20 November 2023

Abstract

:
Pathogens cause infections and millions of deaths globally, while antipathogens are drugs or treatments designed to combat them. To date, multifunctional nanomaterials (NMs), such as organic, inorganic, and nanocomposites, have attracted significant attention by transforming antipathogen livelihoods. They are very small in size so can quickly pass through the walls of bacterial, fungal, or parasitic cells and viral particles to perform their antipathogenic activity. They are more reactive and have a high band gap, making them more effective than traditional medications. Moreover, due to some pathogen’s resistance to currently available medications, the antipathogen performance of NMs is becoming crucial. Additionally, due to their prospective properties and administration methods, NMs are eventually chosen for cutting-edge applications and therapies, including drug administration and diagnostic tools for antipathogens. Herein, NMs have significant characteristics that can facilitate identifying and eliminating pathogens in real-time. This mini-review analyzes multifunctional NMs as antimicrobial tools and investigates their mode of action. We also discussed the challenges that need to be solved for the utilization of NMs as antipathogens.

Graphical Abstract

1. Introduction

Pathogens are microorganisms, including bacteria, viruses, fungi, and parasites, that may induce acute and chronic infections in their hosts after entering the body through ingestion, inhalation, or direct contact. These pathogens are directly responsible for millions of annual cases of infections and fatalities worldwide [1]. Nanomaterials (NMs) with dimensions under 100 nm offer unique characteristics that are suitable for various applications [2]. Currently, NMs are attracting significant attention, transforming antipathogen livelihoods. NMs can quickly enter bacterial, fungal, and protozoal cell walls and viral particles to perform their antipathogenic activity, owing to their ultra-small size, improved surface-to-volume ratio, greater reactivity, and high band gap [3,4]. Thus, researchers are designing nanostructures with several antipathogenic advantages. Ultimately, silver NPs (Ag-NPs) were among the first NPs with remarkable antipathogenic effects [5]. Transition metals (Ag, Cu, Zn), metal oxides (Fe2O3, TiO2, ZnO2), and carbon-based NMs also have intrinsic antipathogenic effects [6]. Furthermore, because of some pathogen’s resistance to currently available medications, the antibacterial, antiviral, antifungal, and antiparasitic performances of NMs are becoming important [7,8].
Hence, NMs offer the potential for developing new therapeutic strategies, such as drug administration and diagnostic tools for antipathogens [9]. Herein, NMs have significant characteristics that facilitate the identification of pathogens and their elimination in real time. For example, the utilization of NMs as drug carriers is an efficient way to fight against several pathogens [10,11]. NMs as a novel drug delivery method allow drugs to cross cell membranes and enter the cytoplasm, effectively killing intracellular infections and improving drug effectiveness against pathogens while minimizing negative consequences in humans and animals [12,13]. In brief, NMs, with their antimicrobial properties, are promising for emerging applications, but their function, structural characteristics, and therapeutic effectiveness remain unexplored, necessitating additional research for optimal execution [14].
This mini review aims to provide a detailed analysis of the classification, production, characterization, mode of action, and outcomes of NMs as antimicrobial agents. The main objective of this study is to explore the potential of NMs in overcoming drug resistance and enhancing the effectiveness of therapeutic interventions. In a more comprehensive approach, our study delved into the considerations pertaining to the long-term sustainability and adaptability of NMs across various environments. This review also provides insights and perspectives that will contribute to the understanding of the pathogens proliferation mechanism as well as the potential for expanding nanotechnology-based techniques for combating pathogens.

2. Overview of Nanomaterials

2.1. Classification of Nanomaterials

The classification of NMs is diverse. Based on their spatial features on the nanometer scale, NMs can be divided into zero-dimensional NMs, one-dimensional NMs, two-dimensional NMs, and nanostructured materials. Based on their morphology, NMs can be classified into nano powder materials, nano-bulk materials, nanofiber materials, nanofilm materials, and nano-liquid materials [15]. Based on their function, NMs can be categorized into nano-magnetic materials, nano-biomaterials, nano-pharmaceutical materials, nano-catalytic materials, nano-wave-absorbing materials, and so on. Based on their chemical composition, NMs can be divided into organic NMs, inorganic NMs, and nanocomposites [16,17].

2.1.1. Organic Nanomaterials

Organic NMs are carbon-based compounds with covalent bonds, which provide mechanical strength, electrical conductivity, and thermal stability. They are suitable for various applications, such as portable electronics and medication delivery systems. Some examples are carbon nanotubes (CNTs), graphene, polymer NPs, and lipid-based NPs for gene therapy [18,19,20].

2.1.2. Inorganic Nanomaterials

Inorganic NMs are non-carbon-based elements with unique physiochemical characteristics that are used in various domains. Metals like Au, Ag, iron (Fe), and platinum (Pt) are used in medication delivery, imaging, and biosensors. Metal oxide NMs (TiO2, Fe3O4, and ZnO) are used in photocatalysis, while QDs and semiconductor NMs have size-dependent optical features for next-generation technologies and photovoltaic panels [19,21,22].

2.1.3. Nanocomposites

Nanocomposite, also known as hybrid NMs that refers to the combination of two or more distinct components such as polymer, metallic, or ceramic nanomaterials. These nanocomposites are used in numerous fields, such as automobiles, aerospace, and architecture. For example, Mg/CNT offers a higher tensile strength, fatigue resistance, and wear resistance, while thermoplastic/thermoset nanocomposites offer easy recycling, longevity, and chemical resistance. The Al2O3/SiO2 nanocomposite is a popular electrical insulator in electronics, aviation, and cars, superior to conventional polymer composites in the electric and healthcare sectors [23,24,25].

2.2. Synthesis and Characterization of Nanomaterials

NMs are generally manufactured through two different approaches: top-down and bottom-up methods. Top-down methods involve applying physical, chemical, or mechanical treatments to reduce large-scale materials to the nanoscale level [26,27]. Examples of top-down methods are ball milling, plasma arc synthesis, and lithography techniques. On the other hand, in the bottom-up process, NMs are synthesized from smaller building blocks like atoms, molecules, or NPs; examples of bottom-up methods are chemical synthesis, sol–gel synthesis, and vapor phase deposition approaches [28,29]. Material characterization helps in the design of new materials by understanding their chemical, mechanical, physical, and microstructural properties. Tools for characterization, include structural characterization, surface characterization, elemental analysis, surface charge characterization, crystallinity, pore structure characterization, and thermal stability. Moreover, Fourier transform infrared spectroscopy (FT-IR), nuclear magnetic resonance (NMR), UV-Vis spectroscopy, transmission electron microscopy (TEM), scanning electron microscopy (SEM), X-ray photoelectron spectroscopy (XPS), and X-ray crystallography (XRD) are the most popular techniques to improve advanced, cutting-edge materials [20,30,31].

2.3. Multiple Functions of Nanomaterials

Significantly, NMs possess remarkable characteristics including their surface area, porosity, pore volume, biocompatibility, non-toxicity, electromagnetic properties, and biodegradability. These properties enable innovation in various domains, including biomedical, agriculture, and industrial applications [32]. For example, NM-based diagnostic systems offer high specificity, low detection limits, and portability, making them valuable tools for pathogen detection and surveillance [33]. They revolutionize medicine by enabling targeted drug delivery, regenerative medicine, and biosensors [34]. NMs are known to improve livestock efficiency, carcass characteristics, intestinal microbiota, and prevent oxidative harm due to their growth-promoting, immune-stimulating, and antimicrobial properties when added to animal feed [35]. In addition, NMs are essential for energy technologies, environmental remediation, electronics, and solar system manufacturing, enabling miniaturization, higher computing power, and enhanced device performance [36,37,38]. In short, NMs are expected to play a significant role in fostering future advancements as shown in Figure 1.

2.4. Role of Nanomaterials as Delivery Systems That Enhance the Antimicrobial Activity of Potential Agents

Nanomaterials hold great promise in drug delivery systems, biomedicine, and environmental protection due to their unique properties such as high surface area, large pore volume, high porosity, and quantum effects. By establishing the optimal storage and delivery conditions, they can significantly enhance the efficacy of antibacterial medicines. Nanomaterials can encapsulate antimicrobial agents, enabling them to bypass cell walls, enter cells, or attach directly to microbes. Their magnetic responsiveness and photothermal properties regulate medication release, increase drug stability, and enhance efficacy. However, nanoparticles may have harmful effects on cells and microbes, and their interaction may alter their pharmacokinetic features [39]

3. Research Advances of Nanomaterials as Antipathogens

Recently, NMs have gained significant attention for their antimicrobial properties, which have been shown to be effective in combating various pathogen-related diseases, including those caused by bacteria, viruses, fungi, and parasites. These NMS provide enhanced antimicrobial activity, reduced antibiotic resistance, controlled drug delivery, surface disinfection, and rapid diagnostics [40]. Summary of nanomaterials for antimicrobial activities is given below in Table 1.

3.1. Nanomaterials for Antibacterial Applications

3.1.1. Organic NMs

NMs are a promising method for combating bacteria and resistant microbes, offering antibacterial properties in organic, inorganic, and hybrid forms. NMs diverse chemical structures, particularly nanosized ones, are designed to combat highly antibiotic-resistant biofilms. Moreover, NMs have improved antimicrobial medicine’s effectiveness by dissolving and dispersing biofilms, which pose a significant barrier in clinical settings. For example, Liu and colleagues [41] demonstrated polymeric micelles as nanocarriers for hydrophobic antimicrobials such as Triclosan. Triclosan is a pH-responsive shell that targets S. aureus biofilms, allowing bacterial lipases to break down biofilms and release encapsulated medicines. Antibiotic-resistant bacteria demand innovative antibacterial medications, with organic NMs gaining attention for their tailored size and large surface-to-volume ratio [101]. In contrast, Costa et al. [102] explored rifampicin-containing poly-lactic acid (PLA) NPs functionalized with poly-L-lysine (PLL), which reversed negative charges to improve antibiotic delivery in S. aureus biofilms, enhancing carrier retention capacity and treatment efficacy. Moreover, Hoque and colleagues [42] found that N-(2-hydroxypropyl)-3-trimethylammonium chitosan chlorides effectively kill MDR bacteria by disrupting the bacterial membrane and exhibiting minimal resistance. In vitro, results confirmed their non-toxic behavior, low skin tissue inflammation, and reduced methicillin-resistant S. aureus (MRSA) burden in superficial skin infections without adverse effects.
In addition, nanotechnology and NMs have significantly impacted the field of livestock antibacterial medicine [43]. Multiple studies have shown that ampicillin-loaded chitosan NPs can suppress E. coli growth, prolong ampicillin release, and improve animal welfare. Liposomes, an amphiphilic delivery system, enhance meat preservation by encapsulating active compounds, extending shelf life, and promoting antibacterial and antioxidant effects [103]. For instance, Singh and coworkers [44] demonstrated that liposomes and solid-lipid NPs can improve meat ingredient surface quality, effectiveness, stability, sensory quality, and bioavailability, while essential oil nano emulsions offer antibacterial and antioxidant properties. Recently, Li and colleagues [45] discovered that administering liposome-associated fimbriae antigens to chickens at 8 and 10 weeks of age increased their IgA and IgG responses and reduced Salmonella enterica excretion. Additionally, researchers have developed NPs for enhanced feed detection in poultry. They have developed a nanomaterial-based technology for detecting nanoparticles in poultry feed. This technology can improve the sensitivity and accuracy of detection, resulting in better control of feed quality and safety. NPs with a polystyrene base, PEG linker, and mannose-attracting biomolecule could potentially replace antibiotics and reduce antibiotic-resistant bacteria [104].

3.1.2. Inorganic NMs

Inorganic NMs exhibit enhanced antibacterial activity, biosensing, broad spectrum, and drug delivery capabilities against both Gram-positive and Gram-negative bacteria [105]. Biosensors use enzymes or antibodies to assess the effectiveness of sterile agents, enabling real-time detection of drug-resistant strains for effective therapies and infection control. Accordingly, Fouda and his research team [46] discovered selenium NPs (Se-NPs) as versatile therapeutic agents, biocides, antioxidants, catalysts, and photoreactive substances. They offer broad-spectrum defense against bacteria, cancer, fungi, and pathogens, exhibit photocatalytic performance, and can be recycled five times. Additionally, inorganic NMs are photothermal candidates that convert light energy into heat to target bacteria, killing germs without destroying healthy tissues [106]. Chen and colleagues [47] found that Y-4-produced palladium NPs have broad NIR absorption, making microorganism eradication easy, economical, and sustainable against Bacillus megaterium.These NPs improve dispersity, light utilization stability, biocompatibility, and photothermal efficacy against S. aureus and E. coli.
In addition, Adegbeye and coworkers [107] demonstrated that inorganic NPs like Ag and Cu can improve feed efficiency, prevent periodontal disease in horses, and address issues like environmental pollution, antibiotic resistance, digestive disorders, and gut health management. Additionally, ZnO-NPs have the potential to be used as antibiotic and anticoccidial replacements due to their bioavailability, characteristics, and impact on veterinary biological systems [108]. In this regard, Yusof and his team [48] highlighted the effectiveness of ZnO-NPs as an alternative antibiotic against multidrug-resistant bacteria in the poultry industry, inhibiting the growth of Salmonella spp., E. coli, and Staphylococcus aureus. Moreover, Hasssan et al. [49] reported that NMs, like ZnO, can improve animal health by promoting development and reducing diarrhea in piglets and dairy cows with recessive mastitis. Furthermore, Tsakmakidis et al. [50] study on FeO and Ag-NPs on ram sperm found that Ag-NPs demonstrated superior antibacterial activity and cytotoxicity, indicating potential for sperm therapy. Accordingly, Kot et al. [51] explored the effectiveness of metal NPs (Ag-NPs, Cu-NPs, Au-NPs, Pt-NPs, and Fe-NPs) in combating digital dermatitis in cows caused by Treponema bacteria, with Ag-NPs and Cu-NPs showing the most biocidal effect. Interestingly, Au-NP-based diagnostics in veterinary science have significantly improved the detection of pathogens and toxins in poultry and cattle, including bacterial infections like anthrax and brucellosis, thereby enhancing the quality of veterinary care [52]. Moreover, quantum dots (QDs) are being used to study livestock gamete biology and reproductive challenges. These biocompatible, photo-stable NPs can provide either targeted or non-targeted imaging with higher signal intensity than organic fluorescent molecules [53]. Based on this, Chatterjee and coworkers [54] proposed QDs-NPs activated by light to create superoxides as a treatment for drug-resistant bacterial infections, reducing viability by seven times. Additionally, researchers have developed QDFM immunochromatography for biological and chemical detection, offering a fast, efficient, specific, high-sensitivity, and simple operation, making it a potential immunolabeling technology [109].

3.1.3. Hybrid NMs

Researchers have developed antibacterial drugs using NMs like graphene and polymers as matrix materials for metal NMs (Au, Ag, ZnO, Cu, and TiO2), enhancing biosensing and cell death [110,111]. Accordingly, Kaushal et al. [55] developed PEG@GO-decorated hybrid antibody biosensors for fast, specific, and higher sensitivity detection of foodborne bacteria like E. coli and Salmonella typhimurium, enabling faster NIR illumination and visual detection. Ahghari and coworkers [56] study on the sustainable synthesis of silver iodate NPs and chitosan (chitosan-AgIO3) showed high bacterial eradication rates against E. coli, Klebsiella pneumoniae, Staphylococcus saprophyticus, Pseudomonas aeruginosa, and S. aureus, indicating the potential of this green, inexpensive, and effective antibacterial agent for biomedical and therapeutic applications. In addition, polymeric NPs offer advantages over lipid-based NPs, including structural integrity, stability, and controlled release capabilities for drug delivery [112]. A potential antibacterial agent for preventing biofilms and intracellular bacterial growth and membrane formation was recently produced by Qiu and colleagues [113] using phosphatidylcholine-chitosan hybrid NMs doped with gentamicin antibiotics. Cui and colleagues [57] found that tea tree oil and liposome-loaded chitosan electro-spun nanofilms effectively inhibited Salmonella in chicken meat, while also preserving the sensory properties of the chicken meat, demonstrating their antibacterial potential in livestock food production. In another instance, Cui et al. [114] developed a chitosan edible film with liposome-encapsulated phage, enhancing phage stability and exhibiting high antibacterial activity against E. coli O157:H7, making it a promising antibacterial packaging for beef preservation. Pabast and colleagues [115] developed a biodegradable coating of chitosan with nano-encapsulated Satureja khuzestanica essential oils (SKEO) to improve food quality and extend shelf-life. The coatings effectively retarded microbial growth, delayed SKEO release, and enhanced sensory attributes, making them a promising candidate for lamb meat shelf-life extension. Furthermore, Amjadi and his team [58] developed betanin nanoliposomes (G/CH NF/ZnO NPs/B NLPs) using gelatin, chitosan nanofiber, and ZnO-NPs in a bio-nanocomposite film for meat preservation. The film effectively inhibited bacterial growth, lipid oxidation, pH changes, and color changes in beef samples, demonstrating its potential for meat preservation. Additionally, Huang and colleagues [116] manufactured chrysanthemum essential oil encapsulated with chitosan and pectin, which reduced oil release and demonstrated sustained antibacterial activity against Campylobacter jejuni in broilers through liposomal delivery.

3.2. Nanomaterials for Antiviral Applications

3.2.1. Organic NMs

Biomedicine is advancing with organic NPs like liposomes, dendrimers, polymer micelles, and carbon-based NPs, promising antiviral candidates due to their viricidal activity, drug carrier properties, selective administration, and regulated release [117]. Accordingly, Bhattacharya and colleagues [59] used membrane-derived vesicles from human corneal epithelial cells, Vero, and CHO cells to combat HSV-1. These liposomes have receptors and neutralizing particles, but limitations in their drug carrier delivery require further research for optimal evaluation and production. Moreover, polyamidoamine dendrimers, naturally antiviral, prevent virus proteins from spreading, invading, and growing. They have the potential to combat diseases like H1N1, HIV, SARS, and Ebola [118,119]. Kandeel et al. [60] studied cationic and anionic dendrimers against MERS-CoV in vivo. The study found that anionic dendrimers reduced MERS-CoV by 40%, while cationic dendrimers assassinated Vero cells. Polyanionic dendrimers can improve targeted antiviral drug delivery.
According to recent studies, alternative treatments like early innate responses and Toll-like receptor ligands have being explored to prevent viral diseases in poultry animals [120]. On this basis, Bavananthasivam et al. [61] found that encapsulating TLR ligands in PLGA-NPs enhances IFN-γ and IL-1β expression, promoting prolonged innate responses and systemic immune responses against Marek’s disease virus (MDV) in chickens. Additionally, Singh and coworkers [121] found that PLGA-NPs effectively combat H9N2 virus in chickens, with nonencapsulated formulations generating higher antibody and mucosal responses. Moreover, Dhakal and colleagues [122] documented a new drug delivery platform using mucoadhesive chitosan NPs. The inactivated swine influenza A virus (SwIAV) vaccine, encapsulated in chitosan NPs, elicited strong immune responses in pigs, reducing viral shedding and lung virus titers, suggesting it as an ideal pig vaccine. Accordingly, Renukaradhya and coworkers [62] demonstrated polyanhydride-NPs, encapsulated in killed SwIAV, being effective as a vaccine in pigs, promoting virus-specific lymphocyte proliferation, fever protection, and reduced viral antigens for pigs. Furthermore, Huang and his research group [123] illustrated that mannosylated gelatin NPs (MnG-NPs) with inactivated Porcine Reproductive and Respiratory Syndrome (PRRSV) in vitro induce T cell-mediated immunity, enhancing monocyte dendritic cell uptake, cytokine expression, and cell activation, making it a significant PRSV vaccine for piglets.

3.2.2. Inorganic NMs

Remarkably, Innocenzi et al. [63] identified that graphene, fullerenes, and carbon dots are promising antiviral agents due to their unique physicochemical characteristics. Graphene oxide has a large surface area and excellent sorption properties, while carbon dots are suitable for viral therapies like HSV-1, HIV, and RSV due to their high aspect ratio and superior mechanical properties. Ag-NPs have been studied for their antiviral effects on several viral infections, including respiratory syncytial virus (RSV), dengue virus (DENV), influenza, hepatitis (HSV-1), poliovirus (PV), and coronaviruses (CoV) [64]. Fruitfully, Yoo and coworkers [124] fabricated a heating filter membrane (HFM) decorated with plasmonic Au-NPs to eliminate H1N1pdm09 virus infectivity. The HFM reduced virus titers by over 99.9% in 10 min, and SARS-CoV-2 virus infectivity by 99% using the photothermal method. This meth by utilizing localized surface plasmon resonance, effectively inactivated the virus, making it suitable for air quality control, viral particle capture, and qRT-PCR genetic information extraction.
Additionally, inorganic NPs are widely used as antiviral agents in domestic animals [65]. Dung and his colleagues [66] found that Ag-NPs can effectively combat African swine fever virus (ASFV) in piglets, thereby reducing viral contamination in pig houses, indicating their potential as a disinfectant. Recently, Zeedan and colleagues [67] reported the biosynthesis of ZnO-NPs and Ag-NPs as antiviral agents against bovine herpesvirus-1 (BoHV-1) in cattl, demonstrating safety in Madin-Darby canine kidney cell culture and experimental animals with minimal cytotoxicity levels. Interestingly, Bai et al. [68] manufactured hollow mesoporous silica-NPs to induce persistent humoral immunity against foot and mouth disease virus-like particles in guinea pigs, enhancing T-lymphocyte proliferation and IFN-γ production, making them a promising nano-adjuvant for vaccines. On the other hand, Fawzy and colleagues [69] found that Au-NPs conjugated with foot and mouth disease virus (FMDV) capsid protein VP1 increased antibody production, IFN-γ production, and macrophage activity in guinea pigs.

3.2.3. Hybrid NMs

Currently, hybrid NMs are progressively used in antiviral approaches due to their integration of antiviral substances, physical barriers, and photothermal or photocatalytic activity [125]. Recently, Ghaffari and his research team [126] investigated that ZnO-NPs with PEGylated coatings effectively inhibited H1N1 by decreasing MDCK-SIAT cell toxicity and improving antiviral activity. PEGylated ZnO-NPs showed 94.6% viral inhibition rates and decreased fluorescence emission intensity. In an additional study, Hodek et al. [127] fabricated a hybrid surface protection of Ag, Cu, and Zn on transparent glass or polymethylmethacrylate (PMMA) plates to combat viral transmission. The coating reduced HIV-1 titers by 99.5–100% after 20 min, while PMMA plates showed 75–100% and 98–100% inactivation after 120 min. The coating targets enveloped viruses, including SARS-CoV-2, and is sterile, safe for Vero and HeLa cells, and minimally cytotoxic. Interestingly, NMs provide targeted antiviral drug delivery with enhanced stability, controlled release, multifunctionality, and biological barriers, enhancing treatment outcomes. Recently, Smith and colleagues [128] developed hybrid poloxamer–lipid NPs to improve antiretroviral lamivudine delivery against HIV-1. M23TC, a phosphoramidite pronucleotide, improved the intracellular delivery and antiretroviral and pharmacokinetic profiles in MDM and CD4+ cells. Likewise, hybrid NMs improve antiviral delivery, clearance, and treatment strategies. In this regard, Abdel-Bar and his research team [129] employed lipid polymer hybrid NPs (LPH-NPs) in combination with piroxicam to administer azithromycin or niclosamide to counter the Corona virus. This system showed entrapment efficiencies, a dose-dependent cellular uptake, and enhanced antiviral efficacy.
Additionally, Zhou et al. [70] reported that GSH-ZnS NPs modified with zinc sulfide demonstrated significant antiviral activity against PRRSV in pigs, indicating potential for antiviral NM development and host restriction factor investigation. Interestingly, Zhou et al. [71] documented that MES-coated tellurium NPs (Te/BSA NPs) inhibited internalization, suppressing virus infection in PRRSV models and demonstrating higher antiviral activity against cattle and pigs. Recently, Du and colleagues [72] developed a method for fabricating virus-like particles using calcium phosphate-biomineralized core immunogen shell NPs, which were used to produce FMDV VLPs, suggesting it as an effective vaccine production method for cattle, sheep, and pigs. Likewise, Chen et al. [130] investigated the antiviral properties of graphene oxide sheets and GO sheets with Ag-NPs against feline coronavirus and infectious bursal disease virus in chickens.

3.3. Nanomaterials for Antifungal Applications

3.3.1. Organic NMs

Organic NMs like micelles, dendrimers, liposomes, graphene, fullerene CNTs, and chitosan offer potential for antifungal therapy due to their large surface area, biocompatibility, targeted delivery, and biodegradability [131]. Leal and coworkers [73] confirmed that itraconazole encapsulated with liposomes had a synergistic effect against Aspergillus in vivo experiments. Adult female Wistar rats were exposed to A. flavus, and itraconazole encapsulated with liposomes showed higher antifungal activity. This drug could be used in clinical settings due to its cost-effectiveness and low cytotoxicity. For another illustration, Helal and his colleagues [132] highlighted the use of organic NMs against fungus-resistant strains and loaded antifungal drugs like nystatin and fluconazole. They found that the biological conjugation and encapsulation of NMs with drugs reduces the toxicity risk and offers promising antifungal therapy.
In addition, organic NPs such as polymeric NPs are being explored as potential antimicrobial drug delivery agents due to their efficient dissolving, entrapment, biocompatibility, low toxicity, and synergistic therapy capabilities in livestock [133]. On this basis, Maldonado et al. [74] proposed synthetic polymeric NPs and rapamycin, which can induce immune tolerance against Streptomyces hygroscopicus, potentially treating allergies, autoimmune diseases, and preventing antidrug antibodies in animal husbandry. Recently, the therapeutic potential of liposomal amphotericin B against A. fumigatus-induced pulmonary mycotic infections in livestock was reported by Siopi and coworkers [75]. Yet, Ahmed et al. [134] found that chitosan NPs effectively inhibited the growth of fungal-like oomycetes Aphanomyces invadans and Saprolegnia parasitica in fish, with the strongest concentration inhibiting 90 % of visible mycelial growth.

3.3.2. Inorganic NMs

Inorganic NMs with a green synthesis approach have antibacterial, antifungal, and antioxidant properties. Metal and metal oxide NMs exhibit potential antifungal activity against Candida, Aspergillus, and dermatophytes. Amin et al. [135] prepared copper oxide (CuO) using Aerva javanica leaf extract to combat fungal infections. In vitro studies of CuO-NPs coupled with amphotericin B showed a higher MIC concentration (160 μg/mL), broad-spectrum activity, minimal toxicity, and a cost-effective approach. Inorganic NMs are also useful for antifungal therapy due to their increased solubility, stability, regulated release, and targeted administration to the infection site [136]. Significantly, these NMs have the potential to overcome the drawbacks of traditional antifungal medicines, including inadequate absorption and resistance to drugs. Recently, Gignone and his research team [76] incorporated clotrimazole into mesoporous silica using theoretical and analytic strategies, evaluating drug behavior through drug adsorption simulation and identifying high-loading-capacity configurations.
On the other hand, Hassan et al. [137] demonstrated that metal NPs like Fe, Zn, Ag, and Se have antimicrobial and antifungal properties, inhibiting mold growth and preventing mycotoxin production, and protecting against aflatoxins and mycotoxins in animals. Recently, Tawab and colleagues [77] demonstrated the antifungal effect of Fe2O3 and Fe3O4 NPs on Aspergillus flavus, isolated from broiler feed. These NPs were synthesized using the co-precipitate method, having a potent antifungal effect. In another study, Nabawy and his team [78] reported that higher concentrations of ZnO and Fe2O3 NPs inhibited A. flavus strains and decreased aflatoxin B1 production in cattle diseases, compared to commercial feed additives. Additionally, Alagawany and colleagues [138] reported that giving Japanese quail Se-NPs improved their growth, blood-related factors, corpse features, state of antioxidant immunity, and gastrointestinal flora, reducing their consumption of feed, as well as having antifungal activities.

3.3.3. Hybrid NMs

Metal and metal oxide NMs such as Ag, Au, or ZnO can be combined with organic molecules, polymers, or carbon-based materials to create hybrid NMs for antifungal activities [139]. For instance, Reda et al. [79] used a sol–gel technique to create calcium-doped zinc oxide ceramic NPs (ZnO-CaO) for combating Candida auris. The ceramics showed better bioactivity and effectiveness in combating multidrug-resistant C. auris, as they release Zn2+, causing oxidative stress and DNA replication and ultimately killing the target microbe. Similarly, Hamad et al. [80] developed thiolated PEGylated cholesterol and PEG-SH nanocomplexes with Au nanorods in a poloxamer 407 hydrogel with fluconazole. These nanocomplexes reduced fungal proliferation (C. albicans) and improved cargo delivery by 14-fold, with minimal cytotoxicity towards human dermal fibroblasts. Also, Hernandez and his coworkers [81] developed titanium-doped copper dioxide/copper iodide (TiO2-Cu2+/Cul) composite NMs using the sol–gel and co-precipitation methods, with minimal inhibitory and fungicidal concentrations for Candida parapsilosis and Aspergillus niger making them cost efficient, and facile for the environment for biomedicine and environmental remediation. More specifically, hybrid NMs effectively disintegrate fungal cell membranes due to their high surface area, reducing proliferation and improving antifungal activity [140]. In this regard, Mohaptara and colleagues [141] prepared a green Ag-ZnO nanocomposite against Schizosaccharomyces pombe, reducing cell proliferation with minimal cytotoxicity, indicating potential antifungal activity in biomedicine and healthcare settings.
In addition, Masry and colleagues [142] found that nanobiotechnological applications in mycotoxicology are promising due to their size-dependent properties. They demonstrated that metal nanocomposites (Fe3O4/CuO/ZnO) can counteract ochratoxin residues in broilers by decreasing body weight, immunological responses, and oxidative stress, while enhancing kidney function. In another instance, Arias et al. [82] developed a miconazole nanocarrier using iron oxide NPs and chitosan, which demonstrated superior antifungal activity against C. albicans and Candida glabrata biofilms in veterinary applications, reducing CFU and metabolism and preventing external magnetic field effects. Interestingly, Atef et al. [143] found that ZnO-NPs and cinnamon oils effectively inhibited fungal growth in cattle mastitis, demonstrating a synergistic effect on the significant inhibition of fungal growth. Kalinska et al. [83] found that Ag-NPs, when combined with Cu-NPs, demonstrated strong antifungal activity against Candida albicans in dairy cows and goats; notably, Ag-NPs showed stronger activity than the Ag-Cu complex.

3.4. Nanomaterials for Antiparasitic Applications

3.4.1. Organic NMs

Recently, standard treatments for parasite infections are facing resistance and poor functionality, prompting the development of organic nanomedicines as potential antiparasitic therapies. These materials reduce drug dosage and cytotoxicity and improve pharmacological potency [144]. In this regard, Moles and colleagues [84] developed an immunoliposome with antibodies targeting RBC surface protein glycophorin A, targeting naive and Plasmodium-infected RBCs. The liposomes loaded with chloroquinoline effectively transferred the drugs, inhibiting parasite growth. Furthermore, combination therapy encapsulates antiparasitic medications with immune modulators, improving treatment outcomes and enhancing bioavailability and therapeutic value [145]. Accordingly, Moles and his research group [85] confirmed immuno-PEG-liposomes for targeted drug delivery in a murine malaria model, efficiently encapsulating amphiphilic drugs like chloroquine and primaquine using a pH gradient. This method effectively inhibited parasite growth and improved drug activity after 15 min of exposure. Sawicka et al. [86] documented that liposome-based vaccines have strong immune responses against parasitic pathogens like Toxoplasma gondii. This reported that the intramuscular injection of MIC3 plasmids induced a significant and effective immune response against T. gondii, increasing serum levels of IgG2 and IgG1. Additionally, Zhang and his research group [87] reported a new approach to anti-coccidiosis drug formulation that involves using 3-carboxyphenylboronic acid-modified chitosan conjugates and diclazuril for site-specific drug release in chicken intestinal tracts. In another instance, a self-nanoemulsifying system (SNEDDS) has been developed to improve the solubilization capacity of buparvaquone (BPQ), a veterinary drug, for treating visceral leishmaniasis. The system, adsorbable on chitosan polymers, has shown enhanced oral bioavailability and potent in vitro efficacy in inhibiting parasite replication in the spleen and liver [88].

3.4.2. Inorganic NMs

Moreover, artemisinin-based combination therapy (ACT) effectively treats mild malaria by targeting molecular markers and studying resistance genetics for improved results [146]. On this basis, Foko et al. [89] optimized and characterized the green synthesis of Ag-NPs using A. cordifolia leaves for potential medical uses. These polycrystalline, stable spheres showed strong antiplasmodial action against P. falciparum strains, making them safe for blood use. Green nanotechnology offers alternative malaria drug/insecticide development. Likewise, inorganic NPs aid in detecting and diagnosing parasitic diseases by interacting with receptors or biomarkers [147]. Additionally, inorganic NMs including mesoporous silica, metals (Cu, Ag, and Au), and metal oxides (TiO2 and ZnO) are gaining attention for their improved therapeutic efficacy against parasites like malaria, leishmaniasis, and toxoplasmosis [90,91,92]. In this regard, Adeyemi et al. [93] found that Au, Ag, and platinum (Pt) NPs have promising anti-Toxoplasma gondii therapeutic activity. Au-NPs and Ag-NPs showed a 13-fold increase in parasite killing compared to host cells, while Pt-NPs showed a 75% reduction in parasite growth. Tsamesidis et al. [148] studied silica-based NPs (Si-NPs) for improved drug delivery against malaria and leishmania parasites. They found that Si-NPs reduced leishmania activity but increased resistance to certain antileishmanial drugs. Furthermore, Jahani and colleagues [94] manufactured Au-NPs with labeled antigen B that can detect antibodies against the hydatid cyst disease of domestic animals, which is caused by Echinococcus granulosus, making it a simple, cost-effective, and selective early detection method. Additionally, cattle and buffalo are susceptible to Toxocariasis due to the gastrointestinal worm Toxocara vitulorum. Mohamed and colleagues [149] investigated the anthelmintic effects of Ag-NPs on both male and female worms as a result of drug resistance. Changes in body structure and the possible intake of drugs were identified. Recently, Ag-NPs synthesized from Azadirachta indica showed potent anthelmintic properties against Haemonchus contortus, a common parasite of domestic animals [95]. Another study by Aydin and colleagues [96] demonstrated that ZnO and FeO-NPs have anthelmintic effects on Toxocara vitulorum in cattle; these NPs caused oxidative/nitrosative stress, leading to the increased mortality of protozoans in the host.

3.4.3. Hybrid NMs

Significantly, hybrid NMs combine organic and inorganic components for enhanced antiparasitic activity, drug delivery, stability, and bioavailability, improving treatment outcomes and vaccine formulation [150]. Very recently, investigations have aimed to identify immunogenic sites and reduce autoimmune and allergic reactions for effective parasitic vaccines. Oxidoreductase is a promising target in the SDR family for Toxoplasma gondii prevention [151]. In this regard, Yu et al. [97] developed TgSDRO-pVAX1, a DNA vaccine combining SDR family oxidoreductase, chitosan NPs, and PLGA. The vaccine demonstrated Th1/Th2 immunity, a transformed antibody production, dendritic cell development, and CD4+ and CD8+ T cell development in immunized mice, and that photodynamic therapy offers an alternative for treating localized lesions. Additionally, Sepúlveda et al. [98] synthesized TiO2 doped with Zn using solution combustion and hypericin (HY) for enhanced photodynamic activity against cutaneous leishmaniasis. The nanocomposite showed the highest fluorescence intensity and in vivo effects on the parasite load.
Elfeky et al. [99] developed cellulose nanocrystal (CNC) and ZnO/CuO nanostructures using the sol–chemical and hydrolysis approaches. The CNC/ZnO/CuO nanostructures showed better larvicidal efficacy towards Anopheles stephensi linked to CNC and ZnO/CuO nanostructures. Additionally, Shehu and coworkers [100] documented the biosynthesis of ZnO-CuO nanoporous composites using gum arabic; this composite has been efficaciously employed to control Culex quinquefasciatus, a vector of filariasis. Furthermore, Yang and colleagues [152] developed a magnetic field controllable and disposable electrochemical immunosensor for the detection of clenbuterol in pork samples. These sensors use graphene sheets, Nafion film, and Fe3O4@Au-NPs coated with bovine serum albumin–CLB conjugates, is sensitive, rapid, low-sample-consumable, and disposable.

4. The Mechanism of Nanomaterials for Antipathogens

Pathogens like bacteria, viruses, fungi, and parasites pose a significant threat to living organisms by causing infectious diseases and malignancies. While drugs are used to combat these resistant pathogens, nanotechnology and NMs offer potential solutions due to their antimicrobial properties [153]. Nonetheless, NM’s antimicrobial mechanisms remain unclear; current theories suggest that those mechanisms may involve direct contact, intracellular localization, and oxidative stress. NMs properties are influenced by physical, chemical, and morphological characteristics, leading to distinct modes of action [154]. Several key antipathogenic mechanisms of NMs are briefly discussed below.

4.1. Mode of Action of Nanomaterials for Antibacterial Activity

4.1.1. Disruption to the Cell Membranes

The cell wall and membrane play a crucial role in maintaining the stability of the substances in the bacteria and protect the bacteria from harm [155,156]. NPs with antibacterial properties can attach to the negatively charged cell membrane due to their positive charge when they come into contact with bacteria [157]. The integrity of the cell membrane is damaged, weakening the interaction between lipopolysaccharide layers on the outer membrane. As a result, most lipopolysaccharides and proteins are released from bacteria, enhancing cell permeability and affecting material exchange inside and outside the cell [158]. In addition, NPs can penetrate the bacterial outer wall and accumulate in their inner membrane, causing instability, damage, increased membrane permeability, cell contents leakage, and death. For example, Au-NPs can continuously release ions that adhere to cell walls and membranes, altering the membrane permeability and causing the destruction of the bacterial envelope. Au-NPs can also cause damage through electrostatic attraction with bacterial cell walls, resulting in cell wall rupture and bacterial death [159,160,161].

4.1.2. Production of Reactive Oxygen Species (ROS)

ROS are partially reduced oxygen derivatives with a strong oxidation capacity, including superoxide anions (O2), hydrogen peroxide (H2O2), hydroxyl radicals (•OH), and singlet oxygen (1O2) [162,163]. Maintaining ROS at an appropriate level positively affects cells [164]. However, excessive ROS can have a negative effect and cause serious damage to bacteria [165]. The overproduction of ROS causes oxidative stress, which affects the structure and function of most biomolecules. For example, lipid peroxidation and protein oxidation are significantly increased [166,167], causing plasma membrane damage and cell apoptosis [168]. ROS mainly inactivate bacteria through two mechanisms: (i) the bacterial cell wall is destroyed, resulting in the leakage of cell contents or damage to the normal membrane transport system function, and the normal structure of related proteases is damaged to inactivate them [169]; (ii) ROS damage the sugar components and bases in genetic DNA, causing the double helix structure to be destroyed, causing normal bacterial proliferation and metabolism [170]. For example, Karunakaran and colleagues [171] found that positively charged 2H-MoS2 NPs can effectively attach to the surface of bacteria and stimulate more ROS production within bacterial cells. Qing et al. [172] demonstrated that Au-NPs can induce intracellular ROS production, potentially leading to bacterial death through protein aggregation and DNA destruction.

4.1.3. Interaction with Cell Contents and Damage to DNA

Due to the small particle size of NPs, the antibacterial metal ions released in the solution, such as Ag+, Zn2+, etc., can enter and penetrate the cell, interacting with the cell contents such as proteins, enzymes, and genetic material to inactivate the cell [173]. Studies suggest that Ag+ reacts with protein sulfhydryl groups, inactivating proteins and inhibiting the activity of bacterial respiratory chain dehydrogenase [174,175,176]. NPs can inhibit bacterial replication and induce cell death by attaching to and binding to bacterial DNA, blocking DNA unwinding during transcription and preventing pathogen proliferation [177]. Lee and colleagues [178] confirmed that Au-NPs can induce DNA fragmentation and apoptotic-like cell death, independent of intracellular ROS. The key methods for various NM functions are summarized in Figure 2.

4.2. Mode of Action of Nanomaterials for Antiviral Activity

4.2.1. NMs Directly Interact with the Viruses to Prevent Their Entry

The virus invades into the cell in three different stages: (i) the virus makes contact with the cell membrane, then enters the intracellular space, and subsequently releases the viral genome into the cell; (ii) the proliferation of the viral genome and its expression; (iii) the assembly of new viruses and their release into the extracellular space. NMs can directly influence virus replication, and they can also affect virus replication through immune responses [179]. Ag-NPs effectively combat viruses like HIV-1 and monkeypox [180,181] by binding to gp120 through electrostatic interactions. Additionally, two disulfide links in the carboxyl half of the HIV-1 gp120 glycoprotein are linked by Ag-NPs involved with sulfhydryl groups simultaneously, triggering protein denaturation by reducing disulfide bonds in the CD4 binding zone and inhibiting virus attachment to the host cell membrane [181]. Park and colleagues [182] synthesized a magnetic hybrid colloid loaded with Ag-NPs of varying sizes and found that Ag ions can bind to the sulfhydryl protein on the surface of the virus, thereby damaging the viral envelope and inhibiting the virus. Iron oxide NPs (Fe2O3 and Fe3O4) with glycine have been shown to reduce biotoxicity and inhibit the H1N1 influenza virus. The NMs with smaller diameters and higher surface areas demonstrated specific spatial resistance, effectively preventing virus attachment to host receptors [183]. The research conducted by Abo-Zeid [184] revealed that IO-NPs (Fe2O3 and Fe3O4) successfully interact with the SARS-CoV-2 spike protein receptor binding domain and HCV glycoproteins. Notably, Fe3O4 forms a stable complex that disrupts the adsorption of the virus with host receptors.

4.2.2. NMs Inhibit Viral Genome Replication

NMs can enter host cells, obstruct viral replication, and attach to viral genomes. Capping agents like polymers and surfactants enhance NM’s effectiveness, with capped Ag-NPs being highly efficient [185,186]. Ye and his research group [187] reported that graphene oxide (GO) inhibited virus replication against pseudorabies and porcine epidemic diarrhea in a cell culture. Negatively charged GO caused DNA damage and viral growth inhibition, while, when conjugated with nonionic PVP, it blocked viral infection. Additionally, Ghaffari and coworkers [126] demonstrated that surface-modified ZnO-NPs and PEGylated NPs effectively suppressed the HSV-1 and H1N1 influenza virus replication at maximum non-toxic concentrations. The release of Zn2+ ions from an aqueous dissolution leads to cell apoptosis and potential oxidative stress and DNA damage in viruses.

4.2.3. NMs Prevent Viruses Assembly and Release

Research has shown that metal ions can form chemical bonds with viral nucleic acids or proteins. This disrupts their structure or causes irreversible conformational changes in viral proteins, thereby achieving the goal of inhibiting viral replication. There are two plausible mechanisms that can account for the toxicity of Cu-NPs on viruses. The first mechanism is dissolution-independent, involving Cu2+capture. The second mechanism involves NPs instability, leading to the generation of large levels of Cu2+ [188]. Cu2+ ions can cause capsid disintegration, protein inactivation, and damage to the viral genome, effectively counteracting various viruses by impeding their entry into cells. Additionally, Cu-NPs can deactivate viral proteins in HSV-1 through oxidation and genome destruction, releasing them into the extracellular space [189]. The virus attaches to the host cell, transcribes its genome, initiates replication, synthesizes mRNA and proteins, and aids in the reassembly of progeny virions [190].

4.2.4. Activation of Immune System by Drugs That Can Hinder the Spread of Viruses

Upon entering a cellular environment, viruses undergo essential processes like unpacking, replication, and translation, leading to the production of RNA/DNA molecules and proteins [191]. The host immune system can be stimulated by two mechanisms with respect to viral entrance into the host cell: (i) directly by NPs or (ii) by coating NPs to the drugs. Azharuddin et al. [192] documented that Au-NPs can trigger immunological responses, including humoral and cell-mediated responses, and produce M2e-specific IgG serum antibodies to prevent the spread of influenza virus by regulating cytokine generation and stimulating immune cells. On the other hand, NPs can stimulate the immune system when combined with drugs, inhibiting viral replication and their spread [193]. Figure 3 provides a summary of the key methods for the various NM antiviral function mechanisms.
With this regard, Dungdung et al. [194] utilized the ZnS quantum point as a drug carrier and loaded it with mycophenolic acid (MPA), an immunosuppressant against dengue virus. The study revealed a higher neutralization rate, enhancing the inhibitory effect and increasing the selective index by two-fold. Antiviral drugs can reduce virus infection rates, but the blood–brain barrier limits the drug scope. In this regard, Nair and coworkers [195] demonstrated the release of azidothymidine 5′-triphosphate, an anti-human immunodeficiency virus drug decorated with CoFe2O4@BaTiO3. This triggered release process is intrinsic, dissipation-free, and energy-efficient, achieving release at the intrinsic level without intermediate materials that help to prevent the viral spread into the host cell.

4.3. Mode of Action of NMs for Antifungal Activity

Metal-based NMs exhibit antifungal activity through three mechanisms: membrane rupture, interference with functions, and surface-dependent interactions with fungal cells, making them promising agents [196,197]. On this basis, Salah et al. [198] reported that Co-NPs can inhibit antifungal activity by preventing copper ion invasion and membrane degradation. However, they can also interfere with essential cellular functions, leading to cell death and affecting cell division and protein synthesis. In a different study, Matras and colleagues [199] found Ag-NPs potent antifungal properties in in vitro experiments on F. avenaceum and F. equiseti by disrupting the cell membrane structure, hindering budding activity, and preventing cytotoxicity. Munir and his research group [200] found that titanium ions permeate cell membranes and bind with DNA, while Cu-, Cr-, and Ni-doped TiO2 binds to fungal cells, enhancing its antifungal activity. Accordingly, Morsy and colleagues [201] discovered that CuO-NPs significantly impact broiler chickens growth, immune status, DNA status, and histological structures, with dose-dependent increases in malondialdehyde levels, Cu contents, and the DNA fragmentation percent.

4.4. Mode of Action of NMs for Antiparasitic Activity

NMs are currently being investigated for their potential antiparasitic effectiveness against various parasites by breaking down cell membranes, producing reactive oxygen species, transporting medication, stopping responses, regulating neurotransmission and enzyme activity, and activating the immune system [202]. In this regard, Villiers and his colleagues [203] found that chloroquine deposits in parasites digestive vacuoles prevented the detoxification of heme, leading to toxic hemozoin accumulation. Antiparasitic drugs can impact vital parasitic functions, hindering enzymes, blocking metabolic routes, depleting vital ions, and inhibiting immune defenses, ultimately causing parasite death. Khadragy and colleagues [204] found that biosynthesized Ag-NPs effectively combat Leishmania major infection, reducing cutaneous lesions and enhancing antioxidant enzyme activities in animals. In another instance, Torres et al. [205] documented metronidazole and riluzole to treat Entamoeba histolytica, causing DNA degradation, neurotransmitter interference, protozoan disintegration, nitric oxide generation, and parasitic cell death.

5. Issues and Challenges Need to Be Solved for the Utilizations of NMs as Antipathogens

NMs are used to improve human and animal health through disease identification, prognosis, prevention, and treatment. Still, Researchers are exploring NM-based antipathogenic activity, which is crucial in medicine and agriculture. However, their use in biomedical applications faces challenges due to their adverse effects on living organisms, as shown in Figure 4.

5.1. Biocompatibility and Toxicity

Biocompatibility and toxicity are important factors to consider while using NMs in biomedical applications. Cytotoxicity is influenced by factors such as physicochemical properties, concentration, and exposure duration [206]. Genotoxicity, on the other hand, is affected by factors such as size, shape, surface charge, and composition, which can impact host cell and tissue interactions [207]. Ag-NPs could potentially cause DNA damage and chromosomal aberrations. Most specifically, NMs face challenges in blood contact, including protein adsorption, interference, and nanotoxicity, which can compromise their antimicrobial activity and cause adverse effects [208]. In vitro studies, using cell viability assays can identify NMs potential toxicity and safe concentrations, while in vivo studies evaluate their toxicological effects in complex biological systems using techniques like histopathology, immunohistochemistry, and biochemical analysis [209].

5.2. Appropriate Selection of Nanomaterials

Pathogen identification is a complex process that is influenced by various factors such as behaviors, strains, and genetic mutations. Understanding the target characteristics is crucial for developing effective NPs, avoiding harmful microorganisms [210]. Remarkably, an inappropriate selection of NPs can disrupt organelle distribution, affecting cellular processes like metabolism, protein synthesis, and waste disposal. Small NMs like Au-NPs and QDs have high stability and slow clearance rates, which increase the risk of long-term toxicity or bioaccumulation [211,212].

5.3. Surface Functionalization

The surface functionalization of NMs improves their stability, solubility, and antimicrobial properties, but precise control is challenging due to reaction conditions, surface impurities, and aggregation [213]. Functionalized NMs can disrupt living organisms, causing cytotoxicity, inflammation, and disrupting processes. For instance, Ag, Cu, TiO2, and ZnO-NPs possess antimicrobial properties but can cause toxic effects on living organisms, depending on characteristics including functional groups, dosage, and exposure duration [214,215].

5.4. Storage

Proper storage conditions are essential for NM’s stability and durability, as they are sensitive to environmental factors like temperature, moisture, and light. Improper storage can result in cytotoxicity, oxidative stress, inflammation, DNA damage, and potential health effects, including cancer development. For instance, metal and metal oxide-based NMs are prone to oxidation, while others are sensitive to air or humidity. Hence, understanding material-specific factors and storage conditions is crucial for long-term stability [216,217].

5.5. Dose Optimization

The dose is a crucial factor in antipathogenic applications, particularly, for antimicrobial treatments. It determines the efficacy and safety of therapeutic interventions. The optimal dosage is essential to balance antimicrobial activity and minimize negative effects, including toxicity, immune response, and bioaccumulation [218]. For instance, Ag-NPs, ZnO-NPs, TiO2-NPs, and CNTs have potential antimicrobial properties, but can cause toxicity, inflammation, organ dysfunction, and lung toxicity [219,220,221,222].

5.6. Stability and Aggregation

During antipathogenic activity, hostile hosts encounter stability challenges with nanoparticles, such as aggregation, precipitation, and dissolution. Aggregation reduces NPs effectiveness, while precipitation disrupts suspension stability due to factors like pH, temperature, or ionic strength, resulting in less effective particles, while the dissolution of NPs can compromise structural integrity, release toxic ions, and disrupt cellular processes.

5.7. Drug resistance Development

Drug resistance in pathogenic microorganisms poses significant challenges to traditional antimicrobial therapies. Novel antipathogenic agents face various obstacles, including evolution, biofilm resistance, cross-resistance, safety concerns, and environmental impact, so adaptive resistance is crucial in combating these challenges [223,224]. Multiple studies have revealed that bacteria can resist Ag-NPs through extracellular precipitation, destruction, or modification, similar to drug modification [225], while Cu-NPs induce antibiotic resistance by upregulating efflux pumps and membrane permeability [226]. Nevertheless, a lack of understanding in designing NMs hinders the design of rational strategies for drug resistance and antipathogenic activities, raises environmental concerns, and requires stricter regulations for regulatory approval and large-scale production [227,228].

5.8. Recyclability

NMs are effective in combating microbial infections, but they also face potential toxicity due to interactions with microorganisms, biological interactions, and aggregation. These interactions can lead to increased toxicity, potentially due to factors like size and shape. These interactions can result in the formation of a complex web of molecules that can interact with and bind to microorganisms, posing significant risks to the effectiveness of nanoparticles in treating microbial infections. Recyclability is crucial for utilizing NMs for antipathogenic activities and sustainable development and environmental protection. However, it can be challenging due to their small particle sizes, molecular penetration, aggregation tendencies, and time-consuming recycling processes [229,230,231].

6. Prospects

Antipathogenic activity based on NMs has immense potential for improving disease prevention, diagnosis, and therapy. Researchers are exploring novel NMs that can transport antimicrobial drugs directly to infected regions, improving localized infection treatment. These NMs can encapsulate drugs, protect them from degradation, and enhance their stability. Customization with ligands or antibodies can reduce dosage and improve drug delivery specificity and therapeutic benefits by boosting solubility, cellular absorption, or generating synergistic effects when coupled with drugs. In addition, nanofabrication techniques like 3D printing and bottom-up self-assembly offer promising methods for producing antipathogenic materials. These methods allow for precise control over their structure, composition, and characteristics, reducing resistance development, aiding in scalability, and enabling larger NM manufacturing for broader applications. In short, advancements in NMs and antipathogenic strategies have the potential to improve human and animal health and contribute to environmental remediation.

7. Conclusions

Nanotechnology and NMs offer customized tools for preventive and therapeutic purposes, addressing challenges in traditional antipathogen pathophysiology. These multifunctional NMs can overcome medicinal solubility, toxic exposures, uncontrolled pharmacokinetic issues, and biostability. Drug resistance occurs when high doses are insufficient to rapidly kill microbes, leading to the widespread distribution of untargeted drugs. However, nanocarriers offer molecular-level precision in targeting infected cells, allowing them to deliver multiple antigens to immune cells, which further allows for the development of better vaccines. In this respect, metal and metal oxide-based NMs and liposome-based NMs have been proven effective in preventing pathogen activity. However, production costs remain a significant concern, and clinicians must collaborate with the medical sector to adapt the technology for effective therapy.

Funding

The authors are grateful for financial support from the National Natural Science Foundation of China (No. 32372948), the Jiangsu Provincial Key Research and Development Program (No. BE2022355), and the China Agriculture Research System of MARA.

Conflicts of Interest

Author Sumreen Sohail was employed by the company Careerera. The remaining authors declare that the research was conducted in the absence of any commercial or financial relationships that could be construed as a potential conflict of interest.

References

  1. Chen, Z.K.; Lin, S.; Wu, Y.X.; Zhao, Z.M.; Zhou, X.M.; Sadiq, S.; Zhang, Z.D.; Guo, X.J.; Wu, P. Hsp90 could promote BmNPV proliferation by interacting with Actin-4 and enhance its expression. Dev. Comp. Immunol. 2023, 142, 104667. [Google Scholar] [CrossRef] [PubMed]
  2. Khan, I.; Kang, K.; Khan, A.; Jiyuan, G.; Khan, S.; Khan, S.; Basir, A.; Sadiq, S. Efficient CO2 conversion and organic pollutants degradation over Sm3+ doped and rutile TiO2 nanorods decorated-GdFeO3 nanorods. Int. J. Hydrogen Energy 2023, 48, 32756–32770. [Google Scholar] [CrossRef]
  3. Khan, I.; Yuan, A.; Khan, S.; Khan, A.; Khan, S.; Shah, S.; Luo, M.S.; Yaseen, W.; Shen, X.; Yaseen, M. Graphitic Carbon Nitride Composites with Gold and ZIF-67 Nanoparticles as Visible-Light-Promoted Catalysts for CO2 Conversion and Bisphenol A Degradation. ACS Appl. Nano Mater. 2022, 5, 13404–13416. [Google Scholar] [CrossRef]
  4. Zaman, S.; Khan, I.; Zhang, F.-M.; Khan, S.; Khan, A.; Khan, S.; Sadiq, S.; Rafiq, M.; Saghir, S.; Sun, X.-J. Synthesis of mediator free hollow BiFeO3 spheres/porous g-C3N4 Z-scheme photocatalysts for CO2 conversion and Alizarin Red S degradation. Mater. Sci. Semicond. Process. 2023, 162, 107534. [Google Scholar] [CrossRef]
  5. Ullah, S.; Khalid, R.; Rehman, M.F.; Irfan, M.I.; Abbas, A.; Alhoshani, A.; Anwar, F.; Amin, H.M.A. Biosynthesis of phyto-functionalized silver nanoparticles using olive fruit extract and evaluation of their antibacterial and antioxidant properties. Front. Chem. 2023, 11, 1202252. [Google Scholar] [CrossRef] [PubMed]
  6. Kubo, A.L.; Rausalu, K.; Savest, N.; Žusinaite, E.; Vasiliev, G.; Viirsalu, M.; Plamus, T.; Krumme, A.; Merits, A.; Bondarenko, O. Antibacterial and Antiviral Effects of Ag, Cu and Zn Metals, Respective Nanoparticles and Filter Materials Thereof against Coronavirus SARS-CoV-2 and Influenza A Virus. Pharmaceutics 2022, 14, 2549. [Google Scholar] [CrossRef] [PubMed]
  7. Roy, A.; Bulut, O.; Some, S.; Mandal, A.K.; Yilmaz, M.D. Green synthesis of silver nanoparticles: Biomolecule-nanoparticle organizations targeting antimicrobial activity. RSC Adv. 2019, 9, 2673–2702. [Google Scholar] [CrossRef]
  8. Khan, S.; Qi, K.; Khan, I.; Wang, A.; Liu, J.; Humayun, M.; Khan, A.; Bahadur, A.; Alanazi, A.F.; Bououdina, M. Eco-friendly graphitic carbon nitride nanomaterials for the development of innovative biomaterials: Preparation, properties, opportunities, current trends, and future outlook. J. Saudi Chem. Soc. 2023, 27, 101753. [Google Scholar] [CrossRef]
  9. Khan, S.S.; Ullah, I.; Ullah, S.; An, R.; Xu, H.; Nie, K.; Liu, C.; Liu, L. Recent Advances in the Surface Functionalization of Nanomaterials for Antimicrobial Applications. Materials 2021, 14, 6932. [Google Scholar] [CrossRef]
  10. Liu, X.; Zhong, X.; Li, C. Challenges in cell membrane-camouflaged drug delivery systems: Development strategies and future prospects. Chin. Chem. Lett. 2021, 32, 2347–2358. [Google Scholar] [CrossRef]
  11. Anbu, P.; Gopinath, S.C.; Jayanthi, S. Synthesis of gold nanoparticles using Platycodon grandiflorum extract and its antipathogenic activity under optimal conditions. Nanomater. Nanotechnol. 2020, 10, 1847980420961697. [Google Scholar] [CrossRef]
  12. Kirtane, A.R.; Verma, M.; Karandikar, P.; Furin, J.; Langer, R.; Traverso, G. Nanotechnology approaches for global infectious diseases. Nat. Nanotechnol. 2021, 16, 369–384. [Google Scholar] [CrossRef] [PubMed]
  13. Khan, M.; Khan, S.; Salar, U.; Khan, K.; Rehman, G.; Gul, N.; Khan, I. Biology-Oriented Drug Synthesis (BIODS), Structural Characterization and Bioactivities of Novel Albendazole Derivatives. Lett. Drug Des. Discov. 2019, 16, 1329–1338. [Google Scholar] [CrossRef]
  14. Jabbar, A.; Abbas, A.; Assad, N.; Naeem-ul-Hassan, M.; Alhazmi, H.A.; Najmi, A.; Zoghebi, K.; Al Bratty, M.; Hanbashi, A.; Amin, H.M.A. A highly selective Hg2+ colorimetric sensor and antimicrobial agent based on green synthesized silver nanoparticles using Equisetum diffusum extract. RSC Adv. 2023, 13, 28666–28675. [Google Scholar] [CrossRef] [PubMed]
  15. Asghar, H.; Khan, I.; Saeed, M.; Wu, P.; Khan, A. Synthesis of g-C3N4/SmFeO3 nanosheets Z-scheme based nanocomposites as efficient visible light photocatalysts for CO2 reduction and Congo red degradation. J. Mater. Res. 2023, 38, 2986–2997. [Google Scholar] [CrossRef]
  16. Khan, I.; Luo, M.s.; Khan, S.; Asghar, H.; Saeed, M.; Khan, S.; Khan, A.; Humayun, M.; Guo, L.; Shi, B. Green synthesis of SrO bridged LaFeO3/g-C3N4 nanocomposites for CO2 conversion and bisphenol A degradation with new insights into mechanism. Environ. Res. 2022, 207, 112650. [Google Scholar] [CrossRef] [PubMed]
  17. Saeed, M.; Adeel, M.; Khan, I.; Akram, N.; Muneer, M. Synthesis of p-n CoO-ZnO Heterojunction for Enhanced Visible-Light Assisted Photodegradation of Methylene Blue. Environ. Sci. Pollut. Res. 2021. [Google Scholar] [CrossRef]
  18. Fang, F.; Li, M.; Zhang, J.; Lee, C.-S. Different Strategies for Organic Nanoparticle Preparation in Biomedicine. ACS Mater. Lett. 2020, 2, 531–549. [Google Scholar] [CrossRef]
  19. Khan, I.; Wang, C.; Khan, S.; Chen, J.; Khan, A.; Shah, S.; Yuan, A.; Khan, S.; Butt, M.; Asghar, H. Bio-capped and green synthesis of ZnO/g-C3N4 nanocomposites and its improved antibiotic and photocatalytic activities: An exceptional approach towards environmental remediation. Chin. J. Chem. Eng. 2022, 56, 215–224. [Google Scholar] [CrossRef]
  20. Shah, S.; Khan, I.; Yuan, A. MoS2 asv a Co-Catalyst for Photocatalytic Hydrogen Production: A Mini Review. Molecules 2022, 27, 3289. [Google Scholar] [CrossRef]
  21. Ghosh, S. Chapter 12—Promising inorganic nanomaterials for future generation. In Applications of Multifunctional Nanomaterials; Thomas, S., Kalarikkal, N., Abraham, A.R., Eds.; Elsevier: Amsterdam, The Netherlands, 2023; pp. 247–263. [Google Scholar] [CrossRef]
  22. Song, J.; Zhao, K.; Yin, X.; Liu, Y.; Khan, I.; Liu, S.-Y. Photocatalytic degradation of tetracycline hydrochloride with g-C3N4/Ag/AgBr composites. Front. Chem. 2022, 10, 1069816. [Google Scholar] [CrossRef] [PubMed]
  23. Namkoong, M.; Guo, H.; Rahman, M.S.; Wang, D.; Pfeil, C.J.; Hager, S.; Tian, L. Moldable and transferrable conductive nanocomposites for epidermal electronics. npj Flex. Electron. 2022, 6, 41. [Google Scholar] [CrossRef] [PubMed]
  24. Camargo, P.; Satyanarayana, K.G.; Wypych, F. Nanocomposites: Synthesis, Structure, Properties and New Application Opportunities. Mater. Res. 2009, 12, 1–39. [Google Scholar] [CrossRef]
  25. Khan, I.; Luo, M.S.; Lin, G.; Khan, S.; Shah, S.; Khan, I.; Khan, A.; Chunjuan, W.; Ai, B.; Zaman, S. Synthesis of phosphate-bridged g-C3N4/LaFeO3 nanosheets Z-scheme nanocomposites as efficient visible photocatalysts for CO2 reduction and malachite green degradation. Appl. Catal. A Gen. 2021, 629, 118418. [Google Scholar] [CrossRef]
  26. Qi, K.; Lv, W.; Khan, I.; Liu, S.-Y. Photocatalytic H2 generation via CoP quantum-dot-modified g-C3N4 synthesized by electroless plating. Chin. J. Catal. 2020, 41, 114–121. [Google Scholar] [CrossRef]
  27. Zhao, K.; Khan, I.; Qi, K.; Liu, Y.; Khataee, A. Ionic liquid assisted preparation of phosphorus-doped g-C3N4 photocatalyst for decomposition of emerging water pollutants. Mater. Chem. Phys. 2020, 253, 123322. [Google Scholar] [CrossRef]
  28. Khan, S.; Wan, C.; Chen, J.; Khan, I.; Luo, M.S.; Wang, C. Eriobotrya japonica assisted green synthesis of g-C3N4 nanocomposites and its exceptional photoactivities for doxycycline and rhodamine B degradation with mechanism insight. J. Chin. Chem. Soc. 2021, 68, 2093–2102. [Google Scholar] [CrossRef]
  29. Yaseen, W.; Meng, S.; Li, W.; Xie, M.; Rafiq, M.; Yusuf, B.; Shah, S.; Khan, I.; Xie, J.; Xu, Y. Facile synthesis of CoMoO4/CoMoB/boron-doped carbon nanocomposite as a highly durable bifunctional electrocatalyst for overall water splitting. Int. J. Hydrogen Energy, 2023, in press. [CrossRef]
  30. Khan, I.; Sun, N.; Wang, Y.; Zhijun, L.; Qu, Y.; Jing, L. Synthesis of SnO2/yolk-shell LaFeO3 nanocomposites as efficient visible-light photocatalysts for 2,4-dichlorophenol degradation. Mater. Res. Bull. 2020, 127, 110857. [Google Scholar] [CrossRef]
  31. Saeed, M.; Albalawi, K.; Khan, I.; Akram, N.; Abd El-Rahim, I.; Al-hag, S. Synthesis of p-n NiO-ZnO heterojunction for photodegradation of crystal violet dye. Alex. Eng. J. 2022, 65, 561–574. [Google Scholar] [CrossRef]
  32. Hayat, A.; Rahman, M.; Khan, I.; Khan, J.; Sohail, M.; Yasmeen, H.; Liu, S.; Qi, K.; Lv, W. Conjugated electron donor-acceptor hybrid polymeric carbon nitride as a photocatalyst for CO2 reduction. Molecules 2019, 24, 1779. [Google Scholar] [CrossRef] [PubMed]
  33. Adeel, M.; Saeed, M.; Khan, I.; Muneer, M.; Akram, N. Synthesis and Characterization of Co–ZnO and Evaluation of Its Photocatalytic Activity for Photodegradation of Methyl Orange. ACS Omega 2021, 6, 1426–1435. [Google Scholar] [CrossRef] [PubMed]
  34. Wu, J.; Cui, Y.; Li, X.; Khan, I.; Liu, X.; Xu, Y.; Song, Y.; Xie, H. Single-atom Pt anchored thiophene ring doped carbon nitride nanosheets for enhanced visible-light photocatalytic H2 evolution and ciprofloxacin degradation. Int. J. Hydrogen Energy, 2023, in press. [CrossRef]
  35. Geetha, K.; Chellapandian, M.; Arulnathan, N.; Ramanathan, A. Nano zinc oxide—An alternate zinc supplement for livestock. Vet. World 2020, 13, 121–126. [Google Scholar] [CrossRef] [PubMed]
  36. Ai, B.; Luo, M.s.; Khan, I. Effective Malachite Green Degradation over the Noble Metal-Doped and MOF-Coupled CsSnBr3 Nanocomposite Catalyst. Processes 2023, 11, 1398. [Google Scholar] [CrossRef]
  37. Taha, T.B.; Barzinjy, A.A.; Hussain, F.H.S.; Nurtayeva, T. Nanotechnology and Computer Science: Trends and advances. Mem.—Mater. Devices Circuits Syst. 2022, 2, 100011. [Google Scholar] [CrossRef]
  38. Kang, K.; Hu, Y.; Khan, I.; He, S.; Fetahi, P. Recent advances in the synthesis and application of magnetic biochar for wastewater treatment. Bioresour. Technol. 2023, 390, 129786. [Google Scholar] [CrossRef]
  39. Sharmin, S.; Rahaman, M.M.; Sarkar, C.; Atolani, O.; Islam, M.T.; Adeyemi, O.S. Nanoparticles as antimicrobial and antiviral agents: A literature-based perspective study. Heliyon 2021, 7, e06456. [Google Scholar] [CrossRef]
  40. Karnwal, A.; Kumar, G.; Pant, G.; Hossain, K.; Ahmad, A.; Alshammari, M.B. Perspectives on Usage of Functional Nanomaterials in Antimicrobial Therapy for Antibiotic-Resistant Bacterial Infections. ACS Omega 2023, 8, 13492–13508. [Google Scholar] [CrossRef]
  41. Liu, Y.; Busscher, H.J.; Zhao, B.; Li, Y.; Zhang, Z.; van der Mei, H.C.; Ren, Y.; Shi, L. Surface-Adaptive, Antimicrobially Loaded, Micellar Nanocarriers with Enhanced Penetration and Killing Efficiency in Staphylococcal Biofilms. ACS Nano 2016, 10, 4779–4789. [Google Scholar] [CrossRef]
  42. Hoque, J.; Adhikary, U.; Yadav, V.; Samaddar, S.; Konai, M.M.; Prakash, R.G.; Paramanandham, K.; Shome, B.R.; Sanyal, K.; Haldar, J. Chitosan Derivatives Active against Multidrug-Resistant Bacteria and Pathogenic Fungi: In Vivo Evaluation as Topical Antimicrobials. Mol. Pharm. 2016, 13, 3578–3589. [Google Scholar] [CrossRef]
  43. Ramachandraiah, K.; Hong, G.-P. Polymer Based Nanomaterials for Strategic Applications in Animal Food Value Chains. Food Rev. Int. 2022, 38, 1577–1606. [Google Scholar] [CrossRef]
  44. Singh, H. Nanotechnology Applications in Functional Foods; Opportunities and Challenges. Prev. Nutr. Food Sci. 2016, 21, 1–8. [Google Scholar] [CrossRef] [PubMed]
  45. Li, W.; Watarai, S.; Iwasaki, T.; Kodama, H. Suppression of Salmonella enterica serovar Enteritidis excretion by intraocular vaccination with fimbriae proteins incorporated in liposomes. Dev. Comp. Immunol. 2004, 28, 29–38. [Google Scholar] [CrossRef] [PubMed]
  46. Fouda, A.; Hassan, S.E.-D.; Eid, A.M.; Abdel-Rahman, M.A.; Hamza, M.F. Light enhanced the antimicrobial, anticancer, and catalytic activities of selenium nanoparticles fabricated by endophytic fungal strain, Penicillium crustosum EP-1. Sci. Rep. 2022, 12, 11834. [Google Scholar] [CrossRef] [PubMed]
  47. Chen, Y.; Chen, Z.; Yang, D.; Zhu, L.; Liang, Z.; Pang, Y.; Zhou, L. Novel Microbial Palladium Nanoparticles with a High Photothermal Effect for Antibacterial Applications. ACS Omega 2023, 8, 1534–1541. [Google Scholar] [CrossRef]
  48. Mohd Yusof, H.; Abdul Rahman, N.A.; Mohamad, R.; Hasanah Zaidan, U.; Samsudin, A.A. Antibacterial Potential of Biosynthesized Zinc Oxide Nanoparticles against Poultry-Associated Foodborne Pathogens: An In Vitro Study. Animals 2021, 11, 2093. [Google Scholar] [CrossRef] [PubMed]
  49. Hassan, A.A.; Sayed El-Ahl, R.M.H.; Oraby, N.H.; El-Hamaky, A.M.A.; Mansour, M.K. Chapter 25—Zinc nanomaterials: Toxicological effects and veterinary applications. In Zinc-Based Nanostructures for Environmental and Agricultural Applications; Abd-Elsalam, K.A., Ed.; Elsevier: Amsterdam, The Netherlands, 2021; pp. 509–541. [Google Scholar] [CrossRef]
  50. Tsakmakidis, I.; Samaras, T.; Anastasiadou, S.; Basioura, A.; Ntemka, A.; Michos, I.A.; Simeonidis, K.; Karagiannis, I.; Tsousis, G.; Angelakeris, M.; et al. Toxic and Microbiological Effects of Iron Oxide and Silver Nanoparticles as Additives on Extended Ram Semen. Animals 2021, 11, 1011. [Google Scholar] [CrossRef]
  51. Kot, M.; Kalińska, A.; Jaworski, S.; Wierzbicki, M.; Smulski, S.; Gołębiewski, M. In Vitro Studies of Nanoparticles as a Potentially New Antimicrobial Agent for the Prevention and Treatment of Lameness and Digital Dermatitis in Cattle. Int. J. Mol. Sci. 2023, 24, 6146. [Google Scholar] [CrossRef]
  52. Manhas, P.K.; Quintela, I.A.; Wu, V.C.H. Enhanced Detection of Major Pathogens and Toxins in Poultry and Livestock With Zoonotic Risks Using Nanomaterials-Based Diagnostics. Front. Vet. Sci. 2021, 8, 673718. [Google Scholar] [CrossRef]
  53. Hill, E.K.; Li, J. Current and future prospects for nanotechnology in animal production. J. Anim. Sci. Biotechnol. 2017, 8, 26. [Google Scholar] [CrossRef] [PubMed]
  54. McCollum, C.R.; Levy, M.; Bertram, J.R.; Nagpal, P.; Chatterjee, A. Photoexcited Quantum Dots as Efficacious and Nontoxic Antibiotics in an Animal Model. ACS Biomater. Sci. Eng. 2021, 7, 1863–1875. [Google Scholar] [CrossRef] [PubMed]
  55. Kaushal, S.; Pinnaka, A.K.; Soni, S.; Singhal, N.K. Antibody assisted graphene oxide coated gold nanoparticles for rapid bacterial detection and near infrared light enhanced antibacterial activity. Sens. Actuators B Chem. 2021, 329, 129141. [Google Scholar] [CrossRef]
  56. Ahghari, M.A.; Ahghari, M.R.; Kamalzare, M.; Maleki, A. Design, synthesis, and characterization of novel eco-friendly chitosan-AgIO3 bionanocomposite and study its antibacterial activity. Sci. Rep. 2022, 12, 10491. [Google Scholar] [CrossRef]
  57. Cui, H.; Bai, M.; Li, C.; Liu, R.; Lin, L. Fabrication of chitosan nanofibers containing tea tree oil liposomes against Salmonella spp. in chicken. LWT 2018, 96, 671–678. [Google Scholar] [CrossRef]
  58. Amjadi, S.; Nazari, M.; Alizadeh, S.A.; Hamishehkar, H. Multifunctional betanin nanoliposomes-incorporated gelatin/chitosan nanofiber/ZnO nanoparticles nanocomposite film for fresh beef preservation. Meat Sci. 2020, 167, 108161. [Google Scholar] [CrossRef] [PubMed]
  59. Bhattacharya, I.; Yadavalli, T.; Wu, D.; Shukla, D. Plasma Membrane-Derived Liposomes Exhibit Robust Antiviral Activity against HSV-1. Viruses 2022, 14, 799. [Google Scholar] [CrossRef]
  60. Kandeel, M.; Al-Taher, A.; Park, B.K.; Kwon, H.J.; Al-Nazawi, M. A pilot study of the antiviral activity of anionic and cationic polyamidoamine dendrimers against the Middle East respiratory syndrome coronavirus. J. Med. Virol. 2020, 92, 1665–1670. [Google Scholar] [CrossRef]
  61. Bavananthasivam, J.; Alkie, T.N.; Astill, J.; Abdul-Careem, M.F.; Wootton, S.K.; Behboudi, S.; Yitbarek, A.; Sharif, S. In ovo administration of Toll-like receptor ligands encapsulated in PLGA nanoparticles impede tumor development in chickens infected with Marek’s disease virus. Vaccine 2018, 36, 4070–4076. [Google Scholar] [CrossRef]
  62. Dhakal, S.; Goodman, J.; Bondra, K.; Lakshmanappa, Y.S.; Hiremath, J.; Shyu, D.L.; Ouyang, K.; Kang, K.I.; Krakowka, S.; Wannemuehler, M.J.; et al. Polyanhydride nanovaccine against swine influenza virus in pigs. Vaccine 2017, 35, 1124–1131. [Google Scholar] [CrossRef]
  63. Innocenzi, P.; Stagi, L. Carbon-based antiviral nanomaterials: Graphene, C-dots, and fullerenes. A perspective. Chem. Sci. 2020, 11, 6606–6622. [Google Scholar] [CrossRef]
  64. Huy, T.Q.; Hien Thanh, N.T.; Thuy, N.T.; Chung, P.V.; Hung, P.N.; Le, A.-T.; Hong Hanh, N.T. Cytotoxicity and antiviral activity of electrochemical—Synthesized silver nanoparticles against poliovirus. J. Virol. Methods 2017, 241, 52–57. [Google Scholar] [CrossRef]
  65. Talebian, S.; Wallace, G.G.; Schroeder, A.; Stellacci, F.; Conde, J. Nanotechnology-based disinfectants and sensors for SARS-CoV-2. Nat. Nanotechnol. 2020, 15, 618–621. [Google Scholar] [CrossRef] [PubMed]
  66. Thi Ngoc Dung, T.; Nang Nam, V.; Thi Nhan, T.; Ngoc, T.T.B.; Minh, L.Q.; Nga, B.T.T.; Phan Le, V.; Viet Quang, D. Silver nanoparticles as potential antiviral agents against African swine fever virus. Mater. Res. Express 2019, 6, 1250g9. [Google Scholar] [CrossRef]
  67. Zeedan, G.; El-Razik, K.; Allam, A.; Abdalhamed, A.; Abou-Zeina, H. Evaluations of Potential Antiviral Effects of Green Zinc Oxide and Silver Nanoparticles against Bovine Herpesvirus-1. Adv. Anim. Vet. Sci. 2020, 8, 433–443. [Google Scholar] [CrossRef]
  68. Bai, M.; Dong, H.; Su, X.; Jin, Y.; Sun, S.; Zhang, Y.; Yang, Y.; Guo, H. Hollow mesoporous silica nanoparticles as delivery vehicle of foot-and-mouth disease virus-like particles induce persistent immune responses in guinea pigs. J. Med. Virol. 2019, 91, 941–948. [Google Scholar] [CrossRef] [PubMed]
  69. Fawzy, M.; Khairy, G.M.; Hesham, A.; Rabaan, A.A.; El-Shamy, A.G.; Nagy, A. Nanoparticles as a novel and promising antiviral platform in veterinary medicine. Arch. Virol. 2021, 166, 2673–2682. [Google Scholar] [CrossRef] [PubMed]
  70. Zhou, Y.; Tong, T.; Jiang, X.; Fang, L.; Wu, Y.; Liang, J.; Xiao, S. GSH-ZnS Nanoparticles Exhibit High-Efficiency and Broad-Spectrum Antiviral Activities via Multistep Inhibition Mechanisms. ACS Appl. Bio Mater. 2020, 3, 4809–4819. [Google Scholar] [CrossRef]
  71. Zhou, Y.; Jiang, X.; Tong, T.; Fang, L.; Wu, Y.; Liang, J.; Xiao, S. High antiviral activity of mercaptoethane sulfonate functionalized Te/BSA nanostars against arterivirus and coronavirus. RSC Adv. 2020, 10, 14161–14169. [Google Scholar] [CrossRef]
  72. Du, P.; Liu, R.; Sun, S.; Dong, H.; Zhao, R.; Tang, R.; Dai, J.; Yin, H.; Luo, J.; Liu, Z.; et al. Biomineralization improves the thermostability of foot-and-mouth disease virus-like particles and the protective immune response induced. Nanoscale 2019, 11, 22748–22761. [Google Scholar] [CrossRef]
  73. Leal, A.F.; Leite, M.C.; Medeiros, C.S.; Cavalcanti, I.M.; Wanderley, A.G.; Magalhães, N.S.; Neves, R.P. Antifungal activity of a liposomal itraconazole formulation in experimental Aspergillus flavus keratitis with endophthalmitis. Mycopathologia 2015, 179, 225–229. [Google Scholar] [CrossRef] [PubMed]
  74. Maldonado, R.A.; LaMothe, R.A.; Ferrari, J.D.; Zhang, A.H.; Rossi, R.J.; Kolte, P.N.; Griset, A.P.; O’Neil, C.; Altreuter, D.H.; Browning, E.; et al. Polymeric synthetic nanoparticles for the induction of antigen-specific immunological tolerance. Proc. Natl. Acad. Sci. USA 2015, 112, E156–E165. [Google Scholar] [CrossRef] [PubMed]
  75. Siopi, M.; Mouton, J.W.; Pournaras, S.; Meletiadis, J. In Vitro and In Vivo Exposure-Effect Relationship of Liposomal Amphotericin B against Aspergillus fumigatus. Antimicrob. Agents Chemother. 2019, 63, e02673-18. [Google Scholar] [CrossRef]
  76. Gignone, A.; Delle Piane, M.; Corno, M.; Ugliengo, P.; Onida, B. Simulation and Experiment Reveal a Complex Scenario for the Adsorption of an Antifungal Drug in Ordered Mesoporous Silica. J. Phys. Chem. C 2015, 119, 13068–13079. [Google Scholar] [CrossRef]
  77. El-Tawab, A.A.A.; El-Hofy, F.I.; Metwally, A. Comparative study on antifungal activity of Fe2O3, and Fe3O4 nanoparticles. Int. J. Adv. Res. 2018, 6, 189–194. [Google Scholar] [CrossRef]
  78. Nabawy, G.A.; Hassan, A.A.; El-Ahl, R.H.S.; Refai, M.K. Effect of metal nanoparticles in comparison with commercial antifungal feed additives on the growth of aspergillus flavus and aflatoxin b1 production. J. Glob. Biosci. 2014, 3, 954–971. [Google Scholar]
  79. Reda, A.E.; Fayed, B. The synthesis of calcium doped zinc oxide ceramic nanoparticles via sol–gel effective against the emerging multidrug-resistant Candida auris. J. Aust. Ceram. Soc. 2023. [Google Scholar] [CrossRef]
  80. Hamad, K.M.; Mahmoud, N.N.; Al-Dabash, S.; Al-Samad, L.A.; Abdallah, M.; Al-Bakri, A.G. Fluconazole conjugated-gold nanorods as an antifungal nanomedicine with low cytotoxicity against human dermal fibroblasts. RSC Adv. 2020, 10, 25889–25897. [Google Scholar] [CrossRef]
  81. Hernandez, R.; Jiménez Chávez, J.; De Vizcaya-Ruiz, A.; Lozano-Alvarez, J.; Escalante, K.; Medina Ramirez, I. Synthesis of TiO2-Cu2+/CuI Nanocomposites and Evaluation of Antifungal and Cytotoxic Activity. Nanomaterials 2023, 13, 1900. [Google Scholar] [CrossRef]
  82. Arias, L.S.; Pessan, J.P.; de Souza Neto, F.N.; Lima, B.H.R.; de Camargo, E.R.; Ramage, G.; Delbem, A.C.B.; Monteiro, D.R. Novel nanocarrier of miconazole based on chitosan-coated iron oxide nanoparticles as a nanotherapy to fight Candida biofilms. Colloids Surf. B Biointerfaces 2020, 192, 111080. [Google Scholar] [CrossRef]
  83. Kalińska, A.; Jaworski, S.; Wierzbicki, M.; Gołębiewski, M. Silver and Copper Nanoparticles—An Alternative in Future Mastitis Treatment and Prevention? Int. J. Mol. Sci. 2019, 20, 1672. [Google Scholar] [CrossRef] [PubMed]
  84. Moles, E.; Urbán, P.; Jiménez-Díaz, M.B.; Viera-Morilla, S.; Angulo-Barturen, I.; Busquets, M.A.; Fernàndez-Busquets, X. Immunoliposome-mediated drug delivery to Plasmodium-infected and non-infected red blood cells as a dual therapeutic/prophylactic antimalarial strategy. J. Control. Release 2015, 210, 217–229. [Google Scholar] [CrossRef] [PubMed]
  85. Moles, E.; Galiano, S.; Gomes, A.; Quiliano, M.; Teixeira, C.; Aldana, I.; Gomes, P.; Fernàndez-Busquets, X. ImmunoPEGliposomes for the targeted delivery of novel lipophilic drugs to red blood cells in a falciparum malaria murine model. Biomaterials 2017, 145, 178–191. [Google Scholar] [CrossRef]
  86. Hiszczyńska-Sawicka, E.; Li, H.; Xu, J.; Akhtar, M.; Holec-Gąsior, L.; Kur, J.; Bickerstaffe, R.; Stankiewicz, M. Induction of immune responses in sheep by vaccination with liposome-entrapped DNA complexes encoding Toxoplasma gondii MIC3 gene. Pol. J. Vet. Sci. 2012, 15, 3–9. [Google Scholar] [CrossRef] [PubMed]
  87. Zhang, X.; Xu, G.; Gadora, K.; Cheng, H.; Peng, J.; Ma, Y.; Guo, Y.; Chi, C.; Zhou, J.; Ding, Y. Dual-sensitive chitosan derivative micelles for site-specific drug release in the treatment of chicken coccidiosis. RSC Adv. 2018, 8, 14515–14526. [Google Scholar] [CrossRef]
  88. Smith, L.; Serrano, D.R.; Mauger, M.; Bolás-Fernández, F.; Dea-Ayuela, M.A.; Lalatsa, A. Orally Bioavailable and Effective Buparvaquone Lipid-Based Nanomedicines for Visceral Leishmaniasis. Mol. Pharm. 2018, 15, 2570–2583. [Google Scholar] [CrossRef]
  89. Kojom Foko, L.P.; Hawadak, J.; Verma, V.; Belle Ebanda Kedi, P.; Eboumbou Moukoko, C.E.; Kamaraju, R.; Pande, V.; Singh, V. Phytofabrication and characterization of Alchornea cordifolia silver nanoparticles and evaluation of antiplasmodial, hemocompatibility and larvicidal potential. Front. Bioeng. Biotechnol. 2023, 11, 1109841. [Google Scholar] [CrossRef]
  90. Prasanna, P.; Kumar, P.; Kumar, S.; Rajana, V.K.; Kant, V.; Prasad, S.R.; Mohan, U.; Ravichandiran, V.; Mandal, D. Current status of nanoscale drug delivery and the future of nano-vaccine development for leishmaniasis—A review. Biomed. Pharmacother. 2021, 141, 111920. [Google Scholar] [CrossRef]
  91. Neves Borgheti-Cardoso, L.; San Anselmo, M.; Lantero, E.; Lancelot, A.; Serrano, J.L.; Hernández-Ainsa, S.; Fernàndez-Busquets, X.; Sierra, T. Promising nanomaterials in the fight against malaria. J. Mater. Chem. B 2020, 8, 9428–9448. [Google Scholar] [CrossRef]
  92. Adeyemi, O.S.; Ishii, K.; Kato, K. L-tryptophan-titanium oxide nanoparticles showed selective anti-Toxoplasma gondii activity and improved host biocompatibility. Biomed. Pharmacother. 2023, 162, 114597. [Google Scholar] [CrossRef]
  93. Adeyemi, O.S.; Murata, Y.; Sugi, T.; Kato, K. Inorganic nanoparticles kill Toxoplasma gondii via changes in redox status and mitochondrial membrane potential. Int. J. Nanomed. 2017, 12, 1647–1661. [Google Scholar] [CrossRef]
  94. Jahani, Z.; Meshgi, B.; Rajabi-Bzl, M.; Jalousian, F.; Hasheminasab, S. Improved serodiagnosis of hydatid cyst disease using gold nanoparticle labeled antigen B in naturally infected sheep. Iran. J. Parasitol. 2014, 9, 218–225. [Google Scholar]
  95. Tomar, R.S.; Preet, S. Evaluation of anthelmintic activity of biologically synthesized silver nanoparticles against the gastrointestinal nematode, Haemonchus contortus. J. Helminthol. 2017, 91, 454–461. [Google Scholar] [CrossRef] [PubMed]
  96. Aydin, A.; Göz, Y.; Yüksek, N.; Ayaz, E. Prevalence of Toxocara vitulorum in Hakkari eastern region of Turkey. Bull. Vet. Inst. Pulawy 2006, 50, 51–54. [Google Scholar]
  97. Yu, Z.; Cao, W.; Gao, X.; Aleem, M.T.; Liu, J.; Luo, J.; Yan, R.; Xu, L.; Song, X.; Li, X. With Chitosan and PLGA as the Delivery Vehicle, Toxoplasma gondii Oxidoreductase-Based DNA Vaccines Decrease Parasite Burdens in Mice. Front. Immunol. 2021, 12, 726615. [Google Scholar] [CrossRef] [PubMed]
  98. Sepúlveda, A.A.L.; Arenas Velásquez, A.M.; Patiño Linares, I.A.; de Almeida, L.; Fontana, C.R.; Garcia, C.; Graminha, M.A.S. Efficacy of photodynamic therapy using TiO2 nanoparticles doped with Zn and hypericin in the treatment of cutaneous Leishmaniasis caused by Leishmania amazonensis. Photodiagnosis Photodyn. Ther. 2020, 30, 101676. [Google Scholar] [CrossRef]
  99. Elfeky, A.S.; Salem, S.S.; Elzaref, A.S.; Owda, M.E.; Eladawy, H.A.; Saeed, A.M.; Awad, M.A.; Abou-Zeid, R.E.; Fouda, A. Multifunctional cellulose nanocrystal /metal oxide hybrid, photo-degradation, antibacterial and larvicidal activities. Carbohydr. Polym. 2020, 230, 115711. [Google Scholar] [CrossRef]
  100. Shehu, Z.; Abba, E.; Danbature, W.; Yoriyo, K.; Abubakar, Z.; Kenneth, Z.; Usiju, Z.; Abubakar, A. Bio-Fabrication of ZnO-CuO Nanoporous Composite and Its Application as Nanolarvicidal Agent for Malaria Vectors. J. Pharm. Res. Int. 2020, 32, 31–39. [Google Scholar] [CrossRef]
  101. Jiang, Y.; Zheng, W.; Tran, K.; Kamilar, E.; Bariwal, J.; Ma, H.; Liang, H. Hydrophilic nanoparticles that kill bacteria while sparing mammalian cells reveal the antibiotic role of nanostructures. Nat. Commun. 2022, 13, 197. [Google Scholar] [CrossRef]
  102. Da Costa, D.; Exbrayat-Héritier, C.; Rambaud, B.; Megy, S.; Terreux, R.; Verrier, B.; Primard, C. Surface charge modulation of rifampicin-loaded PLA nanoparticles to improve antibiotic delivery in Staphylococcus aureus biofilms. J. Nanobiotechnol. 2021, 19, 12. [Google Scholar] [CrossRef]
  103. Wu, J.; Guan, R.; Cao, G.; Liu, Z.; Wang, Z.; Shen, H.; Xia, Q. Antioxidant and Antimicrobial Effects of Catechin Liposomes on Chinese Dried Pork. J. Food Prot. 2018, 81, 827–834. [Google Scholar] [CrossRef] [PubMed]
  104. Ghaffar, S.; Abbas, A.; Naeem-ul-Hassan, M.; Assad, N.; Sher, M.; Ullah, S.; Alhazmi, H.A.; Najmi, A.; Zoghebi, K.; Al Bratty, M.; et al. Improved Photocatalytic and Antioxidant Activity of Olive Fruit Extract-Mediated ZnO Nanoparticles. Antioxidants 2023, 12, 1201. [Google Scholar] [CrossRef] [PubMed]
  105. Groß, S.; Panter, F.; Pogorevc, D.; Seyfert, C.E.; Deckarm, S.; Bader, C.D.; Herrmann, J.; Müller, R. Improved broad-spectrum antibiotics against Gram-negative pathogens via darobactin biosynthetic pathway engineering. Chem. Sci. 2021, 12, 11882–11893. [Google Scholar] [CrossRef] [PubMed]
  106. Bai, X.; Yang, Y.; Zheng, W.; Huang, Y.; Xu, F.; Bao, Z. Synergistic photothermal antibacterial therapy enabled by multifunctional nanomaterials: Progress and perspectives. Mater. Chem. Front. 2023, 7, 355–380. [Google Scholar] [CrossRef]
  107. Adegbeye, M.; Elghandour, M.; Barbabosa-Pliego, A.; Monroy, J.; Mellado, M.; Reddy, P.R.K.; Salem, A.Z. Nanoparticles in Equine Nutrition: Mechanism of Action and Application as Feed Additives. J. Equine Vet. Sci. 2019, 78, 29–37. [Google Scholar] [CrossRef]
  108. Michalak, I.; Dziergowska, K.; Alagawany, M.; Farag, M.R.; El-Shall, N.A.; Tuli, H.S.; Emran, T.B.; Dhama, K. The effect of metal-containing nanoparticles on the health, performance and production of livestock animals and poultry. Vet. Q. 2022, 42, 68–94. [Google Scholar] [CrossRef]
  109. Ding, R.; Chen, Y.; Wang, Q.; Wu, Z.; Zhang, X.; Li, B.; Lin, L. Recent advances in quantum dots-based biosensors for antibiotics detection. J. Pharm. Anal. 2022, 12, 355–364. [Google Scholar] [CrossRef]
  110. Mukherjee, A.; Bose, S.; Shaoo, A.; Das, S.K. Nanotechnology based therapeutic approaches: An advanced strategy to target the biofilm of ESKAPE pathogens. Mater. Adv. 2023, 4, 2544–2572. [Google Scholar] [CrossRef]
  111. Mobed, A.; Hasanzadeh, M.; Seidi, F. Anti-bacterial activity of gold nanocomposites as a new nanomaterial weapon to combat photogenic agents: Recent advances and challenges. RSC Adv. 2021, 11, 34688–34698. [Google Scholar] [CrossRef]
  112. Asem, H.; Zheng, W.; Nilsson, F.; Zhang, Y.; Hedenqvist, M.S.; Hassan, M.; Malmström, E. Functional Nanocarriers for Drug Delivery by Surface Engineering of Polymeric Nanoparticle Post-Polymerization-Induced Self-Assembly. ACS Appl. Bio Mater. 2021, 4, 1045–1056. [Google Scholar] [CrossRef]
  113. Qiu, Y.; Xu, D.; Sui, G.; Wang, D.; Wu, M.; Han, L.; Mu, H.; Duan, J. Gentamicin decorated phosphatidylcholine-chitosan nanoparticles against biofilms and intracellular bacteria. Int. J. Biol. Macromol. 2020, 156, 640–647. [Google Scholar] [CrossRef] [PubMed]
  114. Cui, H.; Yuan, L.; Lin, L. Novel chitosan film embedded with liposome-encapsulated phage for biocontrol of Escherichia coli O157:H7 in beef. Carbohydr. Polym. 2017, 177, 156–164. [Google Scholar] [CrossRef]
  115. Pabast, M.; Shariatifar, N.; Beikzadeh, S.; Jahed, G. Effects of chitosan coatings incorporating with free or nano-encapsulated Satureja plant essential oil on quality characteristics of lamb meat. Food Control 2018, 91, 185–192. [Google Scholar] [CrossRef]
  116. Huang, L.; Teng, W.; Cao, J.; Wang, J. Liposomes as Delivery System for Applications in Meat Products. Foods 2022, 11, 3017. [Google Scholar] [CrossRef] [PubMed]
  117. Patoo, T.S.; Khanday, F.; Qurashi, A. Prospectus of advanced nanomaterials for antiviral properties. Mater. Adv. 2022, 3, 2960–2970. [Google Scholar] [CrossRef]
  118. Paunovska, K.; Loughrey, D.; Dahlman, J.E. Drug delivery systems for RNA therapeutics. Nat. Rev. Genet. 2022, 23, 265–280. [Google Scholar] [CrossRef]
  119. Ullah, A.; Qazi, J.; Rahman, L.; Kanaras, A.G.; Khan, W.S.; Hussain, I.; Rehman, A. Nanoparticles-assisted delivery of antiviral-siRNA as inhalable treatment for human respiratory viruses: A candidate approach against SARS-CoV-2. Nano Sel. 2020, 1, 612–621. [Google Scholar] [CrossRef]
  120. Akbar, H.; Fasick, J.J.; Ponnuraj, N.; Jarosinski, K.W. Purinergic signaling during Marek’s disease in chickens. Sci. Rep. 2023, 13, 2044. [Google Scholar] [CrossRef]
  121. Singh, S.M.; Alkie, T.N.; Abdelaziz, K.T.; Hodgins, D.C.; Novy, A.; Nagy, É.; Sharif, S. Characterization of Immune Responses to an Inactivated Avian Influenza Virus Vaccine Adjuvanted with Nanoparticles Containing CpG ODN. Viral Immunol. 2016, 29, 269–275. [Google Scholar] [CrossRef]
  122. Dhakal, S.; Renu, S.; Ghimire, S.; Shaan Lakshmanappa, Y.; Hogshead, B.T.; Feliciano-Ruiz, N.; Lu, F.; HogenEsch, H.; Krakowka, S.; Lee, C.W.; et al. Mucosal Immunity and Protective Efficacy of Intranasal Inactivated Influenza Vaccine Is Improved by Chitosan Nanoparticle Delivery in Pigs. Front. Immunol. 2018, 9, 934. [Google Scholar] [CrossRef]
  123. Huang, J.; Liu, H.; Wang, M.; Bai, X.; Cao, J.; Zhang, Z.; Wang, Q. Mannosylated gelatin nanoparticles enhanced inactivated PRRSV targeting dendritic cells and increased T cell immunity. Vet. Immunol. Immunopathol. 2021, 235, 110237. [Google Scholar] [CrossRef] [PubMed]
  124. Yoo, S.; Yoon, S.W.; Jung, W.N.; Chung, M.H.; Kim, H.; Jeong, H.; Yoo, K.H. Photothermal inactivation of universal viral particles by localized surface plasmon resonance mediated heating filter membrane. Sci. Rep. 2022, 12, 1724. [Google Scholar] [CrossRef] [PubMed]
  125. Soares, D.C.F.; Poletto, F.; Eberhardt, M.J.; Domingues, S.C.; De Sousa, F.B.; Tebaldi, M.L. Polymer-hybrid nanosystems for antiviral applications: Current advances. Biomed. Pharmacother. 2022, 146, 112249. [Google Scholar] [CrossRef] [PubMed]
  126. Ghaffari, H.; Tavakoli, A.; Moradi, A.; Tabarraei, A.; Bokharaei-Salim, F.; Zahmatkeshan, M.; Farahmand, M.; Javanmard, D.; Kiani, S.J.; Esghaei, M.; et al. Inhibition of H1N1 influenza virus infection by zinc oxide nanoparticles: Another emerging application of nanomedicine. J. Biomed. Sci. 2019, 26, 70. [Google Scholar] [CrossRef] [PubMed]
  127. Hodek, J.; Zajícová, V.; Lovětinská-Šlamborová, I.; Stibor, I.; Müllerová, J.; Weber, J. Protective hybrid coating containing silver, copper and zinc cations effective against human immunodeficiency virus and other enveloped viruses. BMC Microbiol. 2016, 16 (Suppl. 1), 56. [Google Scholar] [CrossRef]
  128. Smith, N.; Bade, A.N.; Soni, D.; Gautam, N.; Alnouti, Y.; Herskovitz, J.; Ibrahim, I.M.; Wojtkiewicz, M.S.; Dyavar Shetty, B.L.; McMillan, J.; et al. A long acting nanoformulated lamivudine ProTide. Biomaterials 2019, 223, 119476. [Google Scholar] [CrossRef] [PubMed]
  129. Abdel-Bar, H.M.; Abdallah, I.A.; Fayed, M.A.A.; Moatasim, Y.; Mostafa, A.; El-Behairy, M.F.; Elimam, H.; Elshaier, Y.; Abouzid, K.A.M. Lipid polymer hybrid nanocarriers as a combinatory platform for different anti-SARS-CoV-2 drugs supported by computational studies. RSC Adv. 2021, 11, 28876–28891. [Google Scholar] [CrossRef]
  130. Chen, Y.N.; Hsueh, Y.H.; Hsieh, C.T.; Tzou, D.Y.; Chang, P.L. Antiviral Activity of Graphene-Silver Nanocomposites against Non-Enveloped and Enveloped Viruses. Int. J. Environ. Res. Public Health 2016, 13, 430. [Google Scholar] [CrossRef]
  131. Yazdani, S.; Mozaffarian, M.; Pazuki, G.; Hadidi, N.; Gallego, I.; Puras, G.; Pedraz, J.L. Design of double functionalized carbon nanotube for amphotericin B and genetic material delivery. Sci. Rep. 2022, 12, 21114. [Google Scholar] [CrossRef]
  132. Helal, S.H.; Abdel-Aziz, H.M.M.; El-Zayat, M.M.; Hasaneen, M.N.A. Preparation, characterization and properties of three different nanomaterials either alone or loaded with nystatin or fluconazole antifungals. Sci. Rep. 2022, 12, 22110. [Google Scholar] [CrossRef]
  133. Halwani, A.A. Development of Pharmaceutical Nanomedicines: From the Bench to the Market. Pharmaceutics 2022, 14, 106. [Google Scholar] [CrossRef]
  134. Ahmed, F.; Soliman, F.M.; Adly, M.A.; Soliman, H.A.M.; El-Matbouli, M.; Saleh, M. In vitro assessment of the antimicrobial efficacy of chitosan nanoparticles against major fish pathogens and their cytotoxicity to fish cell lines. J. Fish Dis. 2020, 43, 1049–1063. [Google Scholar] [CrossRef] [PubMed]
  135. 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]
  136. Chen, J.-N.; Wu, L.-T.; Song, K.; Zhu, Y.-S.; Ding, W. Nonphytotoxic copper oxide nanoparticles are powerful “nanoweapons” that trigger resistance in tobacco against the soil-borne fungal pathogen Phytophthora nicotianae. J. Integr. Agric. 2022, 21, 3245–3262. [Google Scholar] [CrossRef]
  137. Hassan, A.A.; Sayed-Elahl, R.M.; Oraby, N.H.; El-Hamaky, A.M.A. Chapter 11—Metal nanoparticles for management of mycotoxigenic fungi and mycotoxicosis diseases of animals and poultry. In Nanomycotoxicology; Rai, M., Abd-Elsalam, K.A., Eds.; Academic Press: Cambridge, MA, USA, 2020; pp. 251–269. [Google Scholar] [CrossRef]
  138. Alagawany, M.; Qattan, S.Y.A.; Attia, Y.A.; El-Saadony, M.T.; Elnesr, S.S.; Mahmoud, M.A.; Madkour, M.; Abd El-Hack, M.E.; Reda, F.M. Use of Chemical Nano-Selenium as an Antibacterial and Antifungal Agent in Quail Diets and Its Effect on Growth, Carcasses, Antioxidant, Immunity and Caecal Microbes. Animals 2021, 11, 3027. [Google Scholar] [CrossRef] [PubMed]
  139. Abbasi, R.; Shineh, G.; Mobaraki, M.; Doughty, S.; Tayebi, L. Structural parameters of nanoparticles affecting their toxicity for biomedical applications: A review. J. Nanoparticle Res. 2023, 25, 43. [Google Scholar] [CrossRef]
  140. Pathakumari, B.; Liang, G.; Liu, W. Immune defence to invasive fungal infections: A comprehensive review. Biomed. Pharmacother. 2020, 130, 110550. [Google Scholar] [CrossRef]
  141. Mohapatra, B.; Sharma, N. Molecular insights into the antifungal activity of biosynthesized Ag-ZnO hybrid nanostructures against Schizosaccharomyces pombe. Mater. Today Commun. 2023, 36, 106597. [Google Scholar] [CrossRef]
  142. Al Shap, N.F.; El-Sherbeny, E.M.E.; El Masry, D.M.A. The efficacy of metal nanocomposite (Fe3O4/CuO/ZnO) to ameliorate the toxic effects of ochratoxin in broilers. BMC Vet. Res. 2022, 18, 312. [Google Scholar] [CrossRef]
  143. Hassan, A.; Abo-Zaid, K.; Oraby, N. Molecular and Conventional Detection of Antimicrobial Activity of Zinc Oxide Nanoparticles and Cinnamon Oil against Escherichia coli and Aspergillus flavus. Adv. Anim. Vet. Sci. 2020, 8, 839–847. [Google Scholar] [CrossRef]
  144. Nafari, A.; Cheraghipour, K.; Sepahvand, M.; Shahrokhi, G.; Gabal, E.; Mahmoudvand, H. Nanoparticles: New agents toward treatment of leishmaniasis. Parasite Epidemiol. Control 2020, 10, e00156. [Google Scholar] [CrossRef] [PubMed]
  145. Sun, L.; Li, M.; Yang, J.; Li, J. Cell Membrane-Coated Nanoparticles for Management of Infectious Diseases: A Review. Ind. Eng. Chem. Res. 2022, 61, 12867–12883. [Google Scholar] [CrossRef]
  146. Uwimana, A.; Legrand, E.; Stokes, B.H.; Ndikumana, J.M.; Warsame, M.; Umulisa, N.; Ngamije, D.; Munyaneza, T.; Mazarati, J.B.; Munguti, K.; et al. Emergence and clonal expansion of in vitro artemisinin-resistant Plasmodium falciparum kelch13 R561H mutant parasites in Rwanda. Nat. Med. 2020, 26, 1602–1608. [Google Scholar] [CrossRef] [PubMed]
  147. Patra, S.; Singh, M.; Wasnik, K.; Pareek, D.; Gupta, P.S.; Mukherjee, S.; Paik, P. Polymeric Nanoparticle Based Diagnosis and Nanomedicine for Treatment and Development of Vaccines for Cerebral Malaria: A Review on Recent Advancement. ACS Appl. Bio Mater. 2021, 4, 7342–7365. [Google Scholar] [CrossRef]
  148. Tsamesidis, I.; Lymperaki, E.; Egwu, C.O.; Pouroutzidou, G.K.; Kazeli, K.; Reybier, K.; Bourgeade-Delmas, S.; Valentin, A.; Kontonasaki, E. Effect of Silica Based Nanoparticles against Plasmodium falciparum and Leishmania infantum parasites. J. Xenobiotics 2021, 11, 155–162. [Google Scholar] [CrossRef]
  149. Bahaaeldine, M.A.; El Garhy, M.; Fahmy, S.R.; Mohamed, A.S. In vitro anti-Toxocara vitulorum effect of silver nanoparticles. J. Parasit. Dis. 2022, 46, 409–420. [Google Scholar] [CrossRef]
  150. Asthana, S.; Jaiswal, A.K.; Gupta, P.K.; Dube, A.; Chourasia, M.K. Th-1 biased immunomodulation and synergistic antileishmanial activity of stable cationic lipid–polymer hybrid nanoparticle: Biodistribution and toxicity assessment of encapsulated amphotericin B. Eur. J. Pharm. Biopharm. 2015, 89, 62–73. [Google Scholar] [CrossRef]
  151. Zaheer, T.; Pal, K.; Zaheer, I. Topical review on nano-vaccinology: Biochemical promises and key challenges. Process Biochem. 2021, 100, 237–244. [Google Scholar] [CrossRef]
  152. Yang, X.; Wu, F.; Chen, D.-Z.; Lin, H.-W. An electrochemical immunosensor for rapid determination of clenbuterol by using magnetic nanocomposites to modify screen printed carbon electrode based on competitive immunoassay mode. Sens. Actuators B Chem. 2014, 192, 529–535. [Google Scholar] [CrossRef]
  153. Khezerlou, A.; Alizadeh-Sani, M.; Azizi-Lalabadi, M.; Ehsani, A. Nanoparticles and their antimicrobial properties against pathogens including bacteria, fungi, parasites and viruses. Microb. Pathog. 2018, 123, 505–526. [Google Scholar] [CrossRef]
  154. Babayevska, N.; Przysiecka, Ł.; Iatsunskyi, I.; Nowaczyk, G.; Jarek, M.; Janiszewska, E.; Jurga, S. ZnO size and shape effect on antibacterial activity and cytotoxicity profile. Sci. Rep. 2022, 12, 8148. [Google Scholar] [CrossRef] [PubMed]
  155. Strahl, H.; Errington, J. Bacterial Membranes: Structure, Domains, and Function. Annu. Rev. Microbiol. 2017, 71, 519–538. [Google Scholar] [CrossRef] [PubMed]
  156. Shan, L.; Qin, W.; Panunzio, M.; Biondi, S. Antibacterial Agents Targeting the Bacterial Cell Wall. Curr. Med. Chem. 2020, 27, 2902–2926. [Google Scholar] [CrossRef] [PubMed]
  157. Le Ouay, B.; Stellacci, F. Antibacterial activity of silver nanoparticles: A surface science insight. Nano Today 2015, 10, 339–354. [Google Scholar] [CrossRef]
  158. Amro, N.A.; Kotra, L.P.; Wadu-Mesthrige, K.; Bulychev, A.; Mobashery, S.; Liu, G.-Y. High-Resolution Atomic Force Microscopy Studies of the Escherichia coli Outer Membrane:  Structural Basis for Permeability. Langmuir 2000, 16, 2789–2796. [Google Scholar] [CrossRef]
  159. Zhang, Y.; Shareena Dasari, T.P.; Deng, H.; Yu, H. Antimicrobial Activity of Gold Nanoparticles and Ionic Gold. J. Environ. Sci. Health Part C Environ. Carcinog. Ecotoxicol. Rev. 2015, 33, 286–327. [Google Scholar] [CrossRef]
  160. Dasari, T.P.S.; Zhang, Y.; Yu, H. Antibacterial Activity and Cytotoxicity of Gold (I) and (III) Ions and Gold Nanoparticles. Biochem. Pharmacol. Open Access 2015, 4, 199. [Google Scholar] [CrossRef]
  161. Gualdi, S.; Agnoli, K.; Vitale, A.; Higgins, S.; Eberl, L. Identification of genes required for gold and silver tolerance in Burkholderia cenocepacia H111 by transposon sequencing. Environ. Microbiol. 2022, 24, 737–751. [Google Scholar] [CrossRef]
  162. Mourenza, Á.; Gil, J.A.; Mateos, L.M.; Letek, M. Oxidative Stress-Generating Antimicrobials, a Novel Strategy to Overcome Antibacterial Resistance. Antioxidants 2020, 9, 361. [Google Scholar] [CrossRef]
  163. Hegemann, D.; Hanselmann, B.; Zuber, F.; Pan, F.; Gaiser, S.; Rupper, P.; Maniura-Weber, K.; Ruffieux, K.; Ren, Q. Plasma-deposited AgOx-doped TiOx coatings enable rapid antibacterial activity based on ROS generation. Plasma Process. Polym. 2022, 19, 2100246. [Google Scholar] [CrossRef]
  164. Yoo, Y.; Park, J.-C.; Cho, M.-H.; Yang, J.; Kim, C.-Y.; Jung, K.-H.; Jeon, J.-S.; An, G.; Lee, S.-W. Lack of a Cytoplasmic RLK, Required for ROS Homeostasis, Induces Strong Resistance to Bacterial Leaf Blight in Rice. Front. Plant Sci. 2018, 9, 577. [Google Scholar] [CrossRef] [PubMed]
  165. Qayyum, S.; Oves, M.; Khan, A.U. Obliteration of bacterial growth and biofilm through ROS generation by facilely synthesized green silver nanoparticles. PLoS ONE 2017, 12, e0181363. [Google Scholar] [CrossRef] [PubMed]
  166. Seixas, A.F.; Quendera, A.P.; Sousa, J.P.; Silva, A.F.Q.; Arraiano, C.M.; Andrade, J.M. Bacterial Response to Oxidative Stress and RNA Oxidation. Front. Genet. 2022, 12, 821535. [Google Scholar] [CrossRef] [PubMed]
  167. Elwakil, B.H.; Toderas, M.; El-Khatib, M. Arc discharge rapid synthesis of engineered copper oxides nano shapes with potent antibacterial activity against multi-drug resistant bacteria. Sci. Rep. 2022, 12, 20209. [Google Scholar] [CrossRef] [PubMed]
  168. Zhang, Y.-M.; Rock, C.O. Membrane lipid homeostasis in bacteria. Nat. Rev. Microbiol. 2008, 6, 222–233. [Google Scholar] [CrossRef]
  169. Kashef, N.; Huang, Y.Y.; Hamblin, M.R. Advances in antimicrobial photodynamic inactivation at the nanoscale. Nanophotonics 2017, 6, 853–879. [Google Scholar] [CrossRef]
  170. Hamblin, M.R.; Hasan, T. Photodynamic therapy: A new antimicrobial approach to infectious disease? Photochem. Photobiol. Sci. 2004, 3, 436–450. [Google Scholar] [CrossRef]
  171. Karunakaran, S.; Pandit, S.; Basu, B.; De, M. Simultaneous Exfoliation and Functionalization of 2H-MoS2 by Thiolated Surfactants: Applications in Enhanced Antibacterial Activity. J. Am. Chem. Soc. 2018, 140, 12634–12644. [Google Scholar] [CrossRef]
  172. Qing, Y.; Cheng, L.; Li, R.; Liu, G.; Zhang, Y.; Tang, X.; Wang, J.; Liu, H.; Qin, Y. Potential antibacterial mechanism of silver nanoparticles and the optimization of orthopedic implants by advanced modification technologies. Int. J. Nanomed. 2018, 13, 3311–3327. [Google Scholar] [CrossRef]
  173. Kumar, R.; Umar, A.; Kumar, G.; Nalwa, H.S. Antimicrobial properties of ZnO nanomaterials: A review. Ceram. Int. 2017, 43, 3940–3961. [Google Scholar] [CrossRef]
  174. Eckhardt, S.; Brunetto, P.S.; Gagnon, J.; Priebe, M.; Giese, B.; Fromm, K.M. Nanobio Silver: Its Interactions with Peptides and Bacteria, and Its Uses in Medicine. Chem. Rev. 2013, 113, 4708–4754. [Google Scholar] [CrossRef] [PubMed]
  175. Zheng, K.; Setyawati, M.I.; Leong, D.T.; Xie, J. Antimicrobial silver nanomaterials. Coord. Chem. Rev. 2018, 357, 1–17. [Google Scholar] [CrossRef]
  176. Holt, K.B.; Bard, A.J. Interaction of Silver(I) Ions with the Respiratory Chain of Escherichia coli: An Electrochemical and Scanning Electrochemical Microscopy Study of the Antimicrobial Mechanism of Micromolar Ag+. Biochemistry 2005, 44, 13214–13223. [Google Scholar] [CrossRef] [PubMed]
  177. Arafa, M.G.; El-Kased, R.F.; Elmazar, M.M. Thermoresponsive gels containing gold nanoparticles as smart antibacterial and wound healing agents. Sci. Rep. 2018, 8, 13674. [Google Scholar] [CrossRef]
  178. Lee, H.; Lee, D.G. Gold nanoparticles induce a reactive oxygen species-independent apoptotic pathway in Escherichia coli. Colloids Surf. B Biointerfaces 2018, 167, 1–7. [Google Scholar] [CrossRef]
  179. Serrano-Aroca, Á.; Takayama, K.; Tuñón-Molina, A.; Seyran, M.; Hassan, S.S.; Pal Choudhury, P.; Uversky, V.N.; Lundstrom, K.; Adadi, P.; Palù, G.; et al. Carbon-Based Nanomaterials: Promising Antiviral Agents to Combat COVID-19 in the Microbial-Resistant Era. ACS Nano 2021, 15, 8069–8086. [Google Scholar] [CrossRef]
  180. Lara, H.H.; Garza-Treviño, E.N.; Ixtepan-Turrent, L.; Singh, D.K. Silver nanoparticles are broad-spectrum bactericidal and virucidal compounds. J. Nanobiotechnol. 2011, 9, 30. [Google Scholar] [CrossRef]
  181. Lara, H.H.; Ayala-Nuñez, N.V.; Ixtepan-Turrent, L.; Rodriguez-Padilla, C. Mode of antiviral action of silver nanoparticles against HIV-1. J. Nanobiotechnol. 2010, 8, 1. [Google Scholar] [CrossRef]
  182. Park, S.; Park, H.H.; Kim, S.Y.; Kim, S.J.; Woo, K.; Ko, G. Antiviral properties of silver nanoparticles on a magnetic hybrid colloid. Appl. Environ. Microbiol. 2014, 80, 2343–2350. [Google Scholar] [CrossRef]
  183. Kumar, R.; Nayak, M.; Sahoo, G.C.; Pandey, K.; Sarkar, M.C.; Ansari, Y.; Das, V.N.R.; Topno, R.K.; Bhawna; Madhukar, M.; et al. Iron oxide nanoparticles based antiviral activity of H1N1 influenza A virus. J. Infect. Chemother. 2019, 25, 325–329. [Google Scholar] [CrossRef]
  184. Abo-Zeid, Y.; Ismail, N.S.M.; McLean, G.R.; Hamdy, N.M. A molecular docking study repurposes FDA approved iron oxide nanoparticles to treat and control COVID-19 infection. Eur. J. Pharm. Sci. 2020, 153, 105465. [Google Scholar] [CrossRef] [PubMed]
  185. Sarkar, J.; Das, S.; Aich, S.; Bhattacharyya, P.; Acharya, K. Antiviral potential of nanoparticles for the treatment of Coronavirus infections. J. Trace Elem. Med. Biol. Organ Soc. Miner. Trace Elem. (GMS) 2022, 72, 126977. [Google Scholar] [CrossRef]
  186. Bryaskova, R.; Pencheva, D.; Nikolov, S.; Kantardjiev, T. Synthesis and comparative study on the antimicrobial activity of hybrid materials based on silver nanoparticles (AgNps) stabilized by polyvinylpyrrolidone (PVP). J. Chem. Biol. 2011, 4, 185–191. [Google Scholar] [CrossRef] [PubMed]
  187. Ye, S.; Shao, K.; Li, Z.; Guo, N.; Zuo, Y.; Li, Q.; Lu, Z.; Chen, L.; He, Q.; Han, H. Antiviral Activity of Graphene Oxide: How Sharp Edged Structure and Charge Matter. ACS Appl. Mater. Interfaces 2015, 7, 21571–21579. [Google Scholar] [CrossRef] [PubMed]
  188. Broglie, J.J.; Alston, B.; Yang, C.; Ma, L.; Adcock, A.F.; Chen, W.; Yang, L. Antiviral Activity of Gold/Copper Sulfide Core/Shell Nanoparticles against Human Norovirus Virus-Like Particles. PLoS ONE 2015, 10, e0141050. [Google Scholar] [CrossRef]
  189. Ramos-Zúñiga, J.; Bruna, N.; Pérez-Donoso, J.M. Toxicity Mechanisms of Copper Nanoparticles and Copper Surfaces on Bacterial Cells and Viruses. Int. J. Mol. Sci. 2023, 24, 10503. [Google Scholar] [CrossRef]
  190. Li, Y.; Zhang, D.; Gao, X.; Wang, X.; Zhang, L. 2′- and 3′-Ribose Modifications of Nucleotide Analogues Establish the Structural Basis to Inhibit the Viral Replication of SARS-CoV-2. J. Phys. Chem. Lett. 2022, 13, 4111–4118. [Google Scholar] [CrossRef]
  191. Dimitrov, D.S. Virus entry: Molecular mechanisms and biomedical applications. Nat. Rev. Microbiol. 2004, 2, 109–122. [Google Scholar] [CrossRef]
  192. Azharuddin, M.; Zhu, G.H.; Sengupta, A.; Hinkula, J.; Slater, N.K.H.; Patra, H.K. Nano toolbox in immune modulation and nanovaccines. Trends Biotechnol. 2022, 40, 1195–1212. [Google Scholar] [CrossRef]
  193. Mahajan, S.; Choudhary, S.; Kumar, P.; Tomar, S. Antiviral strategies targeting host factors and mechanisms obliging +ssRNA viral pathogens. Bioorganic Med. Chem. 2021, 46, 116356. [Google Scholar] [CrossRef]
  194. Dungdung, R.; Bayal, M.; Valliyott, L.; Unniyampurath, U.; Nair, S.S.; Pilankatta, R. A slow, efficient and safe nanoplatform of tailored ZnS QD-mycophenolic acid conjugates for in vitro drug delivery against dengue virus 2 genome replication. Nanoscale Adv. 2020, 2, 5777–5789. [Google Scholar] [CrossRef] [PubMed]
  195. Nair, M.; Guduru, R.; Liang, P.; Hong, J.; Sagar, V.; Khizroev, S. Externally-controlled on-demand release of anti-HIV drug AZTTP using magneto-electric nanoparticles as carriers. Nat. Commun. 2013, 4, 1707. [Google Scholar] [CrossRef] [PubMed]
  196. Makvandi, P.; Wang, C.-Y.; Zare, E.N.; Borzacchiello, A.; Niu, L.-N.; Tay, F.R. Metal-Based Nanomaterials in Biomedical Applications: Antimicrobial Activity and Cytotoxicity Aspects. Adv. Funct. Mater. 2020, 30, 1910021. [Google Scholar] [CrossRef]
  197. Farooq, M.; Ilyas, N.; Ilyas, N.; Khan, I.; Saboor, A.; Khan, S.M.; Khan, M.N.; Qayum, A.; Bakhtiar, M.T. Antifungal activity of plant extracts and Silver nano particles against Citrus brown spot pathogen (Alternaria citri). Int. J. Environ. Agric. Res. 2018, 4, 118–125. [Google Scholar]
  198. Salah, I.; Parkin, I.P.; Allan, E. Copper as an antimicrobial agent: Recent advances. RSC Adv. 2021, 11, 18179–18186. [Google Scholar] [CrossRef]
  199. Matras, E.; Gorczyca, A.; Przemieniecki, S.W.; Oćwieja, M. Surface properties-dependent antifungal activity of silver nanoparticles. Sci. Rep. 2022, 12, 18046. [Google Scholar] [CrossRef]
  200. Munir, S.; Asghar, F.; Younis, F.; Tabassum, S.; Shah, A.; Khan, S.B. Assessing the potential biological activities of TiO2 and Cu, Ni and Cr doped TiO2 nanoparticles. RSC Adv. 2022, 12, 3856–3861. [Google Scholar] [CrossRef]
  201. Morsy, E.A.; Hussien, A.M.; Ibrahim, M.A.; Farroh, K.Y.; Hassanen, E.I. Cytotoxicity and Genotoxicity of Copper oxide Nanoparticles in chickens. Biol. Trace Elem. Res. 2021, 199, 4731–4745. [Google Scholar] [CrossRef]
  202. Khatami, M.; Alijani, H.Q.; Mousazadeh, F.; Hashemi, N.; Mahmoudi, Z.; Darijani, S.; Bamorovat, M.; Keyhani, A.; Abdollahpour-Alitappeh, M.; Borhani, F. Calcium carbonate nanowires: Greener biosynthesis and their leishmanicidal activity. RSC Adv. 2020, 10, 38063–38068. [Google Scholar] [CrossRef]
  203. de Villiers, K.A.; Egan, T.J. Heme Detoxification in the Malaria Parasite: A Target for Antimalarial Drug Development. Acc. Chem. Res. 2021, 54, 2649–2659. [Google Scholar] [CrossRef]
  204. El-khadragy, M.; Alolayan, E.M.; Metwally, D.M.; El-Din, M.F.S.; Alobud, S.S.; Alsultan, N.I.; Alsaif, S.S.; Awad, M.A.; Abdel Moneim, A.E. Clinical Efficacy Associated with Enhanced Antioxidant Enzyme Activities of Silver Nanoparticles Biosynthesized Using Moringa oleifera Leaf Extract, Against Cutaneous Leishmaniasis in a Murine Model of Leishmania major. Int. J. Environ. Res. Public Health 2018, 15, 1037. [Google Scholar] [CrossRef] [PubMed]
  205. Velásquez-Torres, M.; Trujillo-Ferrara, J.G.; Godínez-Victoria, M.; Jarillo-Luna, R.A.; Tsutsumi, V.; Sánchez-Monroy, V.; Posadas-Mondragón, A.; Cuevas-Hernández, R.I.; Santiago-Cruz, J.A.; Pacheco-Yépez, J. Riluzole, a Derivative of Benzothiazole as a Potential Anti-Amoebic Agent against Entamoeba histolytica. Pharmaceuticals 2023, 16, 896. [Google Scholar] [CrossRef] [PubMed]
  206. 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] [PubMed]
  207. Ghobashy, M.M.; Elkodous, M.A.; Shabaka, S.H.; Younis, S.A.; Alshangiti, D.M.; Madani, M.; Al-Gahtany, S.A.; Elkhatib, W.F.; Noreddin, A.M.; Nady, N.; et al. An overview of methods for production and detection of silver nanoparticles, with emphasis on their fate and toxicological effects on human, soil, and aquatic environment. Nanotechnol. Rev. 2021, 10, 954–977. [Google Scholar] [CrossRef]
  208. Setyawati, M.I.; Tay, C.Y.; Docter, D.; Stauber, R.H.; Leong, D.T. Understanding and exploiting nanoparticles intimacy with the blood vessel and blood. Chem. Soc. Rev. 2015, 44, 8174–8199. [Google Scholar] [CrossRef] [PubMed]
  209. Sinha, S.; Kumar, R.; Anand, J.; Gupta, R.; Gupta, A.; Pant, K.; Dohare, S.; Tiwari, P.; Kesari, K.K.; Krishnan, S.; et al. Nanotechnology-Based Solutions for Antibiofouling Applications: An Overview. ACS Appl. Nano Mater. 2023, 6, 12828–12848. [Google Scholar] [CrossRef]
  210. Jayakumar, A.; Mathew, S.; Radoor, S.; Kim, J.T.; Rhim, J.-W.; Siengchin, S. Recent advances in two-dimensional nanomaterials: Properties, antimicrobial, and drug delivery application of nanocomposites. Mater. Today Chem. 2023, 30, 101492. [Google Scholar] [CrossRef]
  211. Zein, R.; Sharrouf, W.; Selting, K. Physical Properties of Nanoparticles That Result in Improved Cancer Targeting. J. Oncol. 2020, 2020, 5194780. [Google Scholar] [CrossRef]
  212. Gorshkov, K.; Susumu, K.; Chen, J.; Xu, M.; Pradhan, M.; Zhu, W.; Hu, X.; Breger, J.C.; Wolak, M.; Oh, E. Quantum Dot-Conjugated SARS-CoV-2 Spike Pseudo-Virions Enable Tracking of Angiotensin Converting Enzyme 2 Binding and Endocytosis. ACS Nano 2020, 14, 12234–12247. [Google Scholar] [CrossRef]
  213. Khan, I.; Sun, N.; Zhang, Z.; Zhijun, L.; Humayun, M.; Ali, S.; Qu, Y.; Jing, L. Improved visible-light photoactivities of porous LaFeO3 by coupling with nanosized alkaline earth metal oxides and its mechanism insight. Catal. Sci. Technol. 2019, 9, 3149–3157. [Google Scholar] [CrossRef]
  214. Sidhu, V.; Marchi, J.; Borges, R.; Ahmadi, E. Surface modification of metallic orthopedic implants for anti-pathogenic characteristics. J. Compos. Compd. 2022, 4, 51–60. [Google Scholar] [CrossRef]
  215. Khan, I.; Yuan, A.; Khan, A.; Khan, S.; Khan, S.; Shah, S.; Yaseen, W.; Cui, Y.; Shen, X.; Wang, X. Efficient Visible-Light Activities of TiO2 decorated and Cr3+ incorporated-porous SmFeO3 for CO2 conversion and 4-chlorophenol degradation. Surf. Interfaces 2022, 34, 102358. [Google Scholar] [CrossRef]
  216. Manuja, A.; Kumar, B.; Kumar, R.; Chhabra, D.; Ghosh, M.; Manuja, M.; Brar, B.; Pal, Y.; Tripathi, B.N.; Prasad, M. Metal/metal oxide nanoparticles: Toxicity concerns associated with their physical state and remediation for biomedical applications. Toxicol. Rep. 2021, 8, 1970–1978. [Google Scholar] [CrossRef] [PubMed]
  217. Khan, I.; Khan, S.; Chen, J.; Shah, S.; Yuan, A. Biological Inspired Green Synthesis of TiO2 Coupled g-C3N4 Nanocomposites and Its Improved Activities for Sulfadiazine and Bisphenol A Degradation. J. Clust. Sci. 2022, 34, 1453–1464. [Google Scholar] [CrossRef]
  218. Graham, U.M.; Jacobs, G.; Yokel, R.A.; Davis, B.H.; Dozier, A.K.; Birch, M.E.; Tseng, M.T.; Oberdörster, G.; Elder, A.; DeLouise, L. From Dose to Response: In Vivo Nanoparticle Processing and Potential Toxicity. Adv. Exp. Med. Biol. 2017, 947, 71–100. [Google Scholar] [CrossRef]
  219. Pazos-Ortiz, E.; Roque-Ruiz, J.H.; Hinojos-Márquez, E.A.; López-Esparza, J.; Donohué-Cornejo, A.; Cuevas-González, J.C.; Espinosa-Cristóbal, L.F.; Reyes-López, S.Y. Dose-Dependent Antimicrobial Activity of Silver Nanoparticles on Polycaprolactone Fibers against Gram-Positive and Gram-Negative Bacteria. J. Nanomater. 2017, 2017, 4752314. [Google Scholar] [CrossRef]
  220. Mahmoud, N.N.; Albasha, A.; Hikmat, S.; Hamadneh, L.; Zaza, R.; Shraideh, Z.; Khalil, E.A. Nanoparticle size and chemical modification play a crucial role in the interaction of nano gold with the brain: Extent of accumulation and toxicity. Biomater. Sci. 2020, 8, 1669–1682. [Google Scholar] [CrossRef]
  221. Hu, W.; Wang, C.; Gao, D.; Liang, Q. Toxicity of transition metal nanoparticles: A review of different experimental models in the gastrointestinal tract. J. Appl. Toxicol. 2023, 43, 32–46. [Google Scholar] [CrossRef]
  222. Kim, J.K.; Jo, M.S.; Kim, Y.; Kim, T.G.; Shin, J.H.; Kim, B.W.; Kim, H.P.; Lee, H.K.; Kim, H.S.; Ahn, K.; et al. 28-Day inhalation toxicity study with evaluation of lung deposition and retention of tangled multi-walled carbon nanotubes. Nanotoxicology 2020, 14, 250–262. [Google Scholar] [CrossRef]
  223. Gao, W.; Zhang, L. Nanomaterials arising amid antibiotic resistance. Nat. Rev. Microbiol. 2021, 19, 5–6. [Google Scholar] [CrossRef]
  224. Zhang, S.; Khan, I.; Qin, X.; Qi, K.; Liu, Y.; Bai, S. Construction of 1D Ag-AgBr/AlOOH Plasmonic Photocatalyst for Degradation of Tetracycline Hydrochloride. Front. Chem. 2020, 8, 117. [Google Scholar] [CrossRef] [PubMed]
  225. Franci, G.; Falanga, A.; Galdiero, S.; Palomba, L.; Rai, M.; Morelli, G.; Galdiero, M. Silver nanoparticles as potential antibacterial agents. Molecules 2015, 20, 8856–8874. [Google Scholar] [CrossRef] [PubMed]
  226. Amaro, F.; Morón, Á.; Díaz, S.; Martín-González, A.; Gutiérrez, J.C. Metallic Nanoparticles-Friends or Foes in the Battle against Antibiotic-Resistant Bacteria? Microorganisms 2021, 9, 364. [Google Scholar] [CrossRef] [PubMed]
  227. Panáček, A.; Kvítek, L.; Smékalová, M.; Večeřová, R.; Kolář, M.; Röderová, M.; Dyčka, F.; Šebela, M.; Prucek, R.; Tomanec, O.; et al. Bacterial resistance to silver nanoparticles and how to overcome it. Nat. Nanotechnol. 2018, 13, 65–71. [Google Scholar] [CrossRef]
  228. Zhang, S.; Wang, Y.; Song, H.; Lu, J.; Yuan, Z.; Guo, J. Copper nanoparticles and copper ions promote horizontal transfer of plasmid-mediated multi-antibiotic resistance genes across bacterial genera. Environ. Int. 2019, 129, 478–487. [Google Scholar] [CrossRef]
  229. Gupta, R.; Xie, H. Nanoparticles in Daily Life: Applications, Toxicity and Regulations. J. Environ. Pathol. Toxicol. Oncol. 2018, 37, 209–230. [Google Scholar] [CrossRef]
  230. Furxhi, I.; Costa, A.; Vázquez-Campos, S.; Fito-López, C.; Hristozov, D.; Tamayo Ramos, J.A.; Resch, S.; Cioffi, M.; Friedrichs, S.; Rocca, C.; et al. Status, implications and challenges of European safe and sustainable by design paradigms applicable to nanomaterials and advanced materials. RSC Sustain. 2023, 1, 234–250. [Google Scholar] [CrossRef]
  231. Khan, I.; Luo, M.s.; Guo, L.; Khan, S.; Wang, C.; Saeed, M.; Zaman, S.; Qi, K.; Liu, Q.; Khan, A. Enhanced visible-light photoactivities of porous LaFeO3 by synchronously doping Ni2+ and coupling TS-1 for CO2 reduction and 2,4,6-Trinitrophenol degradation. Catal. Sci. Technol. 2021, 11, 6793–6803. [Google Scholar] [CrossRef]
Figure 1. Diagrammatic illustration depicting synthesis, classification, characterization techniques, and applications of multifunctional nanomaterials.
Figure 1. Diagrammatic illustration depicting synthesis, classification, characterization techniques, and applications of multifunctional nanomaterials.
Molecules 28 07674 g001
Figure 2. Diagrammatic representation of the antibacterial properties of nanoparticles.
Figure 2. Diagrammatic representation of the antibacterial properties of nanoparticles.
Molecules 28 07674 g002
Figure 3. Schematic illustration showing the antiviral mechanism of nanomaterials.
Figure 3. Schematic illustration showing the antiviral mechanism of nanomaterials.
Molecules 28 07674 g003
Figure 4. Issues and challenges regarding nanomaterials as antipathogens.
Figure 4. Issues and challenges regarding nanomaterials as antipathogens.
Molecules 28 07674 g004
Table 1. Summary of nanomaterials for antimicrobial activities.
Table 1. Summary of nanomaterials for antimicrobial activities.
NMsNatureAntipathogensMode of ActionTherapeutic OutcomeRef.
Polymeric micellesOrganicS. aureusMembrane lipases breakdown.Multi-resistance drugs, biofilms[41]
Chitosan HCLOrganicGram-negative and Gram-positive bacteriaDepolarizing the cell membrane.Multi-resistance drug[42]
Chitosan NPsOrganicE. coliGenerate ROS production.Antibacterial activity, meat preservation[43]
LiposomeOrganicGram-negative and Gram-positive bacteriaBreak down cell membrane. Antimicrobial activity, meat preservatives[44]
LiposomeOrganicSalmonella entericaTargeting viral cells; modified liposomes impair cellular processes.Reduce microbial contaminants in poultry feed[45]
Se-NPsInorganicGram-negative and Gram-positive bacteriaIncrease ROS production.Antimicrobial activity[46]
Pd-NPsInorganicS. aureus, E. coliROS induction via NIR.Photothermal activity[47]
ZnO-NPsInorganicGram-negative and Gram-positive bacteriaInduce ROS to disrupt essential proteins.Multidrug-resistant bacteria in the poultry[48,49]
Ag-NPs, Cu-NPs, Au-NPs, Pt-NPs, and Fe-NPsInorganicTreponema bacteriaOxidative stress damages cellular components.Combating hoof disorders in cows[50,51]
Au-NPsInorganicBacillus anthracisInduce ROS to disrupt cell membrane.Diagnostic marker in poultry and livestock[52]
QDsInorganicGram positive and Gram-negative bacteriaThe biochemical process is disrupted by damage to the plasma membrane and the cell wall.Drug-resistant topical infections in livestock[53,54]
PEG-GO-AuNPsHybridE. coli, S. typhimuriumDisrupt vital biomolecules by inducing ROS.Biosensor, antibacterial agent[55]
Chitosan-AgIO3HybridP. aeruginosa, K. pneumoniae, S. saprophyticus, E. coli, S. aureusOxidative stress damages cellular components.Antibacterial activity[56]
Liposome-loaded chitosanHybridSalmonella spp.Activate reactive oxygen species, causing membrane breakdown when exposed to UV light.Livestock food production[57]
Betanin nanoliposomes (G/CH NF/ZnO NPs/B NLPs)HybridE. coliCellular components are damaged by oxidative stress.Meat preservation, antibacterial effects[58]
LiposomesOrganicHSV-1Modified liposomes target viral cells, disrupt cellular machinery.Multi-resistance drug/biofilms[59]
Dendrimers/PLLOrganicH1N1, HIV, SARS, Ebola, MERS-CoVDendrimers interact with spike protein to inhibit DNA synthesis.Antiviral drug delivery modulates the immune response[60]
Polymeric lipid NPs OrganicMDVModified polymeric lipids specifically target viral cells and interfere with biological processes.Eliminate viral re-emergence[61]
Polyanhydride-NPsOrganicSwIAVNPs enhance antigen adsorption, uptake, processing, maturation, immune response regulation, and are easily phagocytosed by APCs.Lymphocyte proliferation, vaccines for pigs[62]
Graphene, fullerenes, and CNTsInorganicHSV-1, HIV, RSVElectrostatic interactions with viral proteins to generate oxidative stress and immune responses.Inhibiting viral replication, photothermal activity[63]
Ag-NPsInorganicH1N1, H3N2, enterovirus 71, HSV-1/HSV-2, DENV, HIV poliovirusPlasma membrane rupturing and cell wall disruption, disturbs the biochemical process.Eradicate viral replication[64]
Cu, Ag, TiO2,grapheneInorganicSARS-CoV-2Release toxic ions and ROS and UV-induced membrane destruction.PDT, PTT, PPE, antiviral activity[65]
Ag-NPsInorganicASFVDamage to membranes due to free radicals and ROS.Disinfectant[66]
ZnO-NPs, Ag-NPsInorganicBoHV-1Cellular damage from oxidative stressAntiviral agents[67]
Mesoporous Si-NPs, Au-NPsInorganicFMDVROS from ions disrupt homeostasis and permeate cells.Vaccines[68,69]
GSH-ZnS NPs,HybridPRRSVOxidative stress damages cellular components due to glycosylation and immunodominant decoy epitopes.Antiviral activity[70]
MES-coated tellurium NPs (Te/BSA NPs)HybridPRRSVTe/BSA nanostars inhibit PRRSV proliferation and prophylactic effect.Antiviral activity[71]
Ca3(PO4)2 biomineralized core immunogen shell NPsHybridFMDVThe addition of polar amino acids to VLPs can enhance their stability in extreme environments, potentially improving their heat resistance.Vaccines[72]
LiposomesOrganicA. flavusInteract with the membrane, causing destabilization, cellular leakage.Drug delivery, antifungal agent[73]
Polymeric NPsOrganicStreptomyces hygroscopicusAntifungal activity involves cell membrane damage, causing cell death.Drug delivery. treating allergies, autoimmune diseases[74]
LiposomesOrganicA. fumigatusLiposome binding affinity for fungal cell walls ensuring stability and preventing toxicity.Antimycotic infections, drug delivery[75]
Si-NPsInorganicC. aurisIon’s release generates ROS disrupt homeostasis cause cell leakage.Drug delivery, MDR[76]
Fe2O3, Fe3O4, ZnO NPsInorganicA. flavusROS induces mitochondrial dysfunctional apoptosis.Antifungal activity[77,78]
ZnO-CaOHybridC. aurisZn2+ disrupts zinc-mediated protein activity, generates oxidative stress.MDR[79]
Chol-PEG-SH, PEG-Fluc-GNRHybridC. albicansOpsonization and phagocytosis inhibit DNA/RNA synthesis.Drug delivery[80]
TiO2-Cu2CuIHybridA. Niger, C. parapsilosisRestrict enzyme function, release of Cu2+, alter NADPH generation.MDR[81]
Iron oxide and chitosan NPsHybridCandida albicans and Candida glabrataROS generation occurs when antifungal NMs attach to antifungal effect cells, elaborating O2 and metal ions.Antifungal activity[82]
Ag@Cu-NPsHybridCandida albicansRelease ions cause oxidative stress, cell wall damage, enzymatic activity inhibition.Antifungal activity[83]
LiposomesOrganicPlasmodium spp.Liposomes interact with ligands or antibodies and release encapsulated drugs.Antiparasitic activity, drug delivery[84]
PEG-liposomesOrganicP. falciparumPreventing immune system recognition and eliminating parasites through drug cellular uptake.Conjugated therapy, drug delivery, MDR[85]
LiposomeOrganicToxoplasma gondiiDestabilizing membranes through acidic pH, disulfide bonding cleaving, and degradation.Vaccines[86]
ChitosanOrganicEimeria spp.Chitosan destabilizes hydrophobic scaffolds in tertiary amines and degrades in response to intracellular environment.Drug delivery[87]
ChitosanOrganicLeishmaniaChitosan destabilizes cellular membrane.Drug delivery, antiparasitic activity[88]
Ag-NPsInorganicP. falciparumInduce ROS causing cellular contents leakage.Antiprotozoal activity[89]
Au, Ag, Cu-NPsInorganicT. gondii, malaria, leishmaniasisRelease ions, generate oxidative stress to kill parasites.Biomarkers[90,91,92]
Au, Ag, Pt NPsInorganicT. gondiiAdsorption, permeation, and cytotoxicity of NPs with electrically charged substances.Antiparasitic activity[93]
Au-NPsInorganicEchinococcus granulosusAuNPs on hydatid cyst protoscoleces, assessing their effects on cell wall and caspase-3 activation.Diagnostic marker[94]
Ag-NPsInorganicHaemonchus contortus, LeishmaniaFree radicals induce oxidative stress.Antiprotozoal activity[95]
ZnO and FeO-NPsInorganicToxocara vitulorumOxidative stress and ROS generation increasing antioxidant enzyme activity.Antiprotozoal activity[96]
PLGA@chitosanHybridT. gondiiAcidic environment causes PLGA degradation, releasing drugs, and targeting parasites.Vaccines[97]
TiO2/Zn-HYHybridL. amazonensisOxidative stress inhibits DNA/RNA synthesis.PDT, photosensitizer, and cutaneous leishmaniasis therapy[98]
CNC/ZnO/CuO HybridAnopheles stephensiGeneration of hydroxyl ions and ROS leads to membrane disruption.Photodegradation and larvicidal activities[99]
ZnO-CuO nanocompositeHybridCulex quinquefasciatusGeneration of ROS antioxidant property of enzymes.Antiprotozoal activity[100]
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

Sadiq, S.; Khan, I.; Shen, Z.; Wang, M.; Xu, T.; Khan, S.; Zhou, X.; Bahadur, A.; Rafiq, M.; Sohail, S.; et al. Recent Updates on Multifunctional Nanomaterials as Antipathogens in Humans and Livestock: Classification, Application, Mode of Action, and Challenges. Molecules 2023, 28, 7674. https://doi.org/10.3390/molecules28227674

AMA Style

Sadiq S, Khan I, Shen Z, Wang M, Xu T, Khan S, Zhou X, Bahadur A, Rafiq M, Sohail S, et al. Recent Updates on Multifunctional Nanomaterials as Antipathogens in Humans and Livestock: Classification, Application, Mode of Action, and Challenges. Molecules. 2023; 28(22):7674. https://doi.org/10.3390/molecules28227674

Chicago/Turabian Style

Sadiq, Samreen, Iltaf Khan, Zhenyu Shen, Mengdong Wang, Tao Xu, Sohail Khan, Xuemin Zhou, Ali Bahadur, Madiha Rafiq, Sumreen Sohail, and et al. 2023. "Recent Updates on Multifunctional Nanomaterials as Antipathogens in Humans and Livestock: Classification, Application, Mode of Action, and Challenges" Molecules 28, no. 22: 7674. https://doi.org/10.3390/molecules28227674

Article Metrics

Back to TopTop