Next Article in Journal
One-Pot Facile Synthesis of CuO–CdWO4 Nanocomposite for Photocatalytic Hydrogen Production
Next Article in Special Issue
Osteoregeneration of Critical-Size Defects Using Hydroxyapatite–Chitosan and Silver–Chitosan Nanocomposites
Previous Article in Journal
Synthesis of New Zirconium Magnetic Nanocomposite as a Bioactive Agent and Green Catalyst in the Four-Component Synthesis of a Novel Multi-Ring Compound Containing Pyrazole Derivatives
Previous Article in Special Issue
Formation of Aggregate-Free Gold Nanoparticles in the Cyclodextrin-Tetrachloroaurate System Follows Finke–Watzky Kinetics
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Mechanisms of Antifungal Properties of Metal Nanoparticles

Department of Medicine, Division of Infectious Diseases, University of British Columbia, 410-2660 Oak St., Vancouver, BC V6H3Z6, Canada
*
Author to whom correspondence should be addressed.
Nanomaterials 2022, 12(24), 4470; https://doi.org/10.3390/nano12244470
Submission received: 16 November 2022 / Revised: 11 December 2022 / Accepted: 13 December 2022 / Published: 16 December 2022
(This article belongs to the Special Issue Development of Nanomaterials for Biomedical Applications)

Abstract

:

Highlights

What are the main findings?
  • Mechanisms by which nanoparticles exert their damage on fungi were described.
  • Transcriptomic studies regarding exposure to nanoparticles were analyzed.
What is the implication of the main finding?
  • Potential applications and future venues were discussed.
  • More studies are needed to determine the safety of future applications of metallic nanoparticles.

Abstract

The appearance of resistant species of fungi to the existent antimycotics is challenging for the scientific community. One emergent technology is the application of nanotechnology to develop novel antifungal agents. Metal nanoparticles (NPs) have shown promising results as an alternative to classical antimycotics. This review summarizes and discusses the antifungal mechanisms of metal NPs, including combinations with other antimycotics, covering the period from 2005 to 2022. These mechanisms include but are not limited to the generation of toxic oxygen species and their cellular target, the effect of the cell wall damage and the hyphae and spores, and the mechanisms of defense implied by the fungal cell. Lastly, a description of the impact of NPs on the transcriptomic and proteomic profiles is discussed.

1. Background

Fungi were initially included as a part of Kingdom Plantae but in 1969 were grouped into Kingdom Fungi, which comprises diverse groups with different morphologies, such as unicellular yeasts and multicellular organisms [1]. However, since then, there have been many changes in classification within and even at the Kingdom level. For example, in 1981, oomycetes were moved to a novel Kingdom termed Kingdom Chromista due to new phylogenetic studies [2].
The exact values of fungal diversity have yet to be discovered, with numbers of species ranging from 2.2–3.8 million [3], up to 5.1 million [4], and 6 million [5], among others. While the species diversity ranges in the millions, around 8000 species are plant pathogens, and only a fraction infects humans. Nonetheless, fungal diseases are responsible for more than 1.6 million deaths yearly, a number over 3-fold higher than malaria estimates and comparable to tuberculosis [6]. Furthermore, the high number of fungal plant pathogens introduces more issues such as (1) loss of biodiversity (e.g., chytridiomycosis as a dominant cause of amphibian extinction) [7], and (2) blights that destroy crops (e.g., rice blast fungus, causing up to 35% rice harvest loss across 85 countries) [8].
The rate of antifungal resistance development [9] has been called “unprecedented”. This is because immunocompromised individuals are at a higher risk of fungal infections than healthy individuals. Moreover, medical advancements over the past few decades (e.g., antibiotic discovery, cancer treatment progress, surgical transplants) and the HIV epidemic have increased the number of immunocompromised people, which has, in turn, shifted fungal infections from being an infrequent cause of disease to being an important contributor to human morbidity and mortality worldwide [10]. There are six antifungal drug classes (see below), and this scarcity, combined with the increasing resistance, has led to the need for novel treatments.
One promising solution as an antifungal agent is the use of nanoparticles (NPs). Many studies investigate this field of nanotechnology, but the mechanisms of action involved in NPs use as an antifungal remain widely unknown. This review summarizes and discusses proposed mechanisms of action, including reactive oxygen species (ROS) production, cell wall architecture, plasma membrane deformation, interaction with fungal structures, inhibition of spore germination, and gene and protein regulation. This review also discusses the formulation of NPs using primary literature focused on NP synthesis for antifungal testing.

2. Fungi Groups

Generally, all fungi may be split into two groups: yeasts or molds (filamentous fungi). Dimorphic fungi, such as Candida albicans, can grow in either form depending on environmental conditions, such as undergoing morphogenesis to engage in pathogenic behavior. Yeasts are unicellular organisms that typically reproduce asexually through binary fission or budding but uncommonly reproduce sexually, usually in high-stress environments such as that void of nutrients. Molds are multicellular organisms comprised of hyphae filaments that branch out as the fungus grows, responsible for enzyme release and nutrient absorption. Usually, hyphae that are composed of cells separated by walls called septum (septa in plural) are termed “septate hyphae” but can also lack walls and cell membranes between the compartments, termed “coenocytic hyphae” or “non-septate hyphae”. Collectively, a network of hyphae is called mycelium. If a mycelium grows in an aggregated form, this aggregation is called a pellet. During the budding process of yeast, when daughter cells stick to the parent cell, and a small structure is formed, this is termed “pseudohyphae” [11]. Molds reproduce sexually or asexually through the formation of spores such as conidia and sporangiospores, which are types of asexual spores. As the primary goal of spores is to be dispersed, they experience a period of dormancy and have different compounds coating them for adaptation to stressful environments [12,13].

3. Fungal Cell Wall Structure

The cell wall is critical for fungal growth, defense, morphogenesis, and biofilm formation. It functions to buffer fluctuations of osmotic pressure, sense external stimuli, and protect against detrimental conditions such as dryness, heat, toxic molecules, etc., while protecting from microbes. The cell wall is also crucial for pathogenicity and virulence in pathogenic fungi as it will aid in invasion while protecting the fungus from host defense mechanisms [14]. However, the distinctive structure makes it a good target for antifungal agents.
The cell wall (Figure 1) contains polysaccharides and proteins such as chitin, glucan, mannan, glucanases, and glycoproteins. The cell wall may appear to be a rigid structure, but it is dynamic, constantly being remodeled due to proliferation (expansion, sporulation, and branching) and environmental challenges [15]. Events such as binary fission or expansion of hyphae require the simultaneous activities of anabolic and catabolic enzymes to remodel the structure as a whole completely.
A brief description of the main components of the fungal cell wall is provided below.

4. Polysaccharides

4.1. Glucan

This polysaccharide (Figure 2) is the major component of the fungal cell wall, and it is synthesized by the polymerization of glucose monomers (approximately 1500 residues) through covalent β-1,3-linkages [16,17,18]. Alternate linkages occur at the C6 position of the linear glucan, where branching of the polysaccharide occurs through β-1,6 linkages of approximately 50 glucose monomers [19,20,21]. The glucan synthesis is mediated by membrane-bound glucan synthases that transfer a glucosyl residue to the chain through a β-1,3 bond in a vectorial synthesis [19,22,23]. Additionally, the branched glucans provide a point of cross-linking to other cell wall components, such as chitin and proteins [24].

4.2. Chitin

This polysaccharide (Figure 3) is synthesized by polymerizing N-acetylglucosamine monomers, which are linked as β-1,4-linkages. The chitin synthesis is linear, and the formation of microfibrils has been observed. Moreover, these microfibrils might condensate and crystallize extracellularly. It represents a total of 1–2% and up to 20% of the total dry weight of yeast and filamentous fungi, respectively, [25]. Chitin provides tensile strength to the cell wall; when disrupted, fungal cells undergo morphological changes accompanied by an osmotic sensitivity [26].

5. Proteins

The proteins on the fungal cell wall are mainly glycoproteins which are extensively modified with N-linked and O-linked oligosaccharides. Their linkage to different oligosaccharide chains depends on the fungal species. For instance, yeast such as Saccharomyces cerevisiae and C. albicans synthesize mannoproteins, which are the product of glycosylation of the proteins with mannose polymers or mannans [15]. Other filamentous fungi, such as Neurospora crassa and Aspergillus fumigatus, contain a hybrid of galactose and glucose, forming galactomannan structures [27,28,29].
A family of five proteins, called GAS, is anchored onto the membrane by a phosphoglyceride glycosylphosphatidylinositol (GPI) [30]. Interestingly, the deletion of the GAS1 gene caused morphological defects in S. cerevisiae [31] and is likely associated with its transglycosylase function [32].
The cell wall undergoes a dynamic rearrangement due to the activity of multiple enzymes. These enzymes are orchestrated during the remodeling of the cell wall, including chitinases and glucanases. This battery of proteins hydrolyzes and rearranges the skeleton of the cell wall before the enzymatic activity of glycosyltransferases, which will finish the synthesis and cross-linking of the polymers [33,34].

5.1. Glucanases

As a component of the cell wall, glucans change as a result of the constant remodeling of the cell wall. For instance, S. cerevisiae contains an array of approximately 15 endo- and exo-1,3-β-glucanases with crucial functions during proliferation and transglycosylation activity [35]. Interestingly, the deletion of endoglucanases caused cell clumping, supporting that these enzymes are involved in the digestion of septa [35]. Similar enzymes have been found in other yeasts, such as C. albicans and Schizosaccharomyces pombe, and similar phenotypes have been observed upon their deletion compared to S. cerevisiae [36].
Endoglucanases have also been reported in filamentous fungi, exhibiting similar glucanase and transglycosidase activities. For example, a dual function was attributed to glucanases in A. fumigatus, which cleave 1,3-β-linkages in glucan with a concomitant transfer to a non-reducing end of another 1,3-β-glucan molecule, leading to an elongation of the chain catalyzed by the enzyme glucan elongating glucanosyltransferase or GEL1. This enzyme is also anchored to the membrane by GPI molecules and showed homology to GAS1 in the yeasts [32].

5.2. Chitinases

Chitin synthases are responsible for the production of chitin. The genome of S. cerevisiae encodes three chitin synthases termed CHS1p, CHS2p, and CHS3p [37]. CHS1p is involved in cell repair during the proliferation of the fungi, CHS2p is involved in synthesizing the septa [38,39], whereas CHS3p generates the rest of the chitin found in the cell [40,41]. Chitin is crucial for cell survival, and deleting the three chitin synthases is lethal for S. cerevisiae [39,42].
Other groups of fungi possess a higher number of chitin synthases. For instance, A. fumigatus has seven genes responsible for chitin polymerization, and their deletions from the genome are not necessarily lethal for growth, but they exhibit morphological changes [43,44]. In the case of N. crassa, four chitin synthases are present in the genome but are also predicted to have three additional isoenzymes [45,46,47,48,49]. As in A. fumigatus, individual deletions of these chitin synthase genes showed different phenotypes, such as slowing growth accompanied by a change in morphology [45].
Under specific physiological events, endochitinases remove chitin, allowing cell wall remodeling. For example, during the duplication of yeast, the newly formed cells (mother and daughter) are separated from each other by hydrolysis, which is mediated by an endochitinase [50,51,52]. These chitinases are also GPI-anchored to the cell membrane. Deleting the CHIA gene encoding a chitinase in A. nidulans affected the spores and decreased the growth rate of the hyphae [53].

6. Nanoparticle Formulation as Antifungal Agents

There is a need for novel antifungal treatments as currently available options are lacking. It is worth noting that the idea of increased potential future resistance is worrying as fungi are eukaryotes, as are their most common hosts found in Kingdoms Animalia and Plantae. That is to say, the eukaryotic hosts of pathogenic fungi have similarities in metabolism and protein structure, and finding targets to differentiate organisms becomes more complicated [54].
Nanotechnology has rapidly progressed over the past few decades and using NPs as potential antifungals have been an expanding field of interest. The present study summarized the literature from 2005 until 2022 regarding the NP formulations that showed antifungal activity (Table 1, Table 2, Table 3, Table 4, Table 5 and Table 6). A summary of the different types of metallic NPs studied is shown in Figure 4.
NPs have unique physical and chemical properties compared to their larger counterparts because their small size leads to a high surface-to-volume ratio. NPs can be produced using the top-down technique, where small particles are formed from a more significant part, or the bottom-up technique, where small particles assemble from smaller pieces. The bottom-up approach is typically favored for the synthesis of NPs as it makes more homogeneous products with fewer defects and has parameters that are easier to control during fabrication. For example, altering compound surface charges can facilitate bottom-up processes such as interfacial electrostatic self-assembly [63]. As another example, the electrochemical synthesis of metal NPs can be utilized to avoid reducing chemicals and control size dispersion based on the cell electrode potential tuning [80]. This is interesting because usually, a bottom-up process requires a metal salt and a reducing agent, such as sodium borohydride or Sn(II) chloride [77,92].
There is no standardized approach for the fabrication of NPs, and they can be found in various shapes, sizes, compositions, and formulations. However, spherical AgNPs are the most popular. Ag has been known for centuries to possess antimicrobial properties [93]. We know today that AgNPs possess these same properties and unique optical, electrical, and other physical/chemical properties. It is proposed that a potential mechanism of toxicity of Ag comes from its ion release coupled with its catalytic oxidation abilities [94].
Combining different AgNPs can exhibit a synergistic effect [95]. For example, combinations of AgNPs with other oxides, such as maghemite (γ-Fe2O3) or magnetite (Fe3O4), can lead to NPs that possess the unique Ag features while incorporating the magnetic features included in the iron oxide compounds, such as controllable alterations through magnetic field manipulation [67]. Moreover, combining AgNPs with graphene-oxide (GO) produced nanosheets with a three- and a seven-fold increase in bacterial inhibition efficiency over their counterparts in one study [63]. However, in another, combining AgNPs with ineffective copper NPs (CuNPs), a synergistic effect was not measured but somewhat diminished the activity of the AgNPs [96].
Other types of oxides have been investigated as well. A study looking at the antifungal activity of metal(loid) oxides found that Al2O3, Mn3O4, SiO2, and SnO2 reduced cell viability in a dose-dependent manner, whereas In2O3 showed no toxicity, even at concentrations of 100 mg/L [97]. GO is also of interest and has been used in functionalization and investigated as it is a single-atom-thick and a 2D sp2-bonded carbon lattice with a large surface area [98]. This is important because a lattice will innately prevent aggregation as dispersion will be enhanced and remain during the usage of the particles. Thus, GO’s surfactant-like properties allow it to attach metal NPs to hyphal interfaces, forming nanosheets with concomitant mechanical damage due to their extremely sharp edges [63].
Other oxides, such as titanium dioxide (TiO2) NPs, have shown that the mechanism of antimicrobial ability requires visible light to affect [99,100]. The photocatalytic activity is based on hydroxyl radical generation. It is known that photocatalysis works through photons exciting electrons to the conduction band and forming electron-hole pairs. An alternative to increase the radical hydroxyl production is doping TiO2 with Ag, potentially by accepting the photoinduced electrons and holes with an increase in the degradation efficiency [100]. Another study found that undoped TiO2NPs did have some antifungal properties, but the combination with Ag enhanced their efficacy. Fungal species also play a role in the antifungal activity of NPs, e.g., Venturia inaequalis was more affected by undoped TiO2 than Fusarium solani. Still, the most effective nanomaterial was Ag-doped hollow TiO2NPs. However, this may have been because the hollow formulation was spherical, and the solid formulation shape was indiscernible in TEM imaging [100]. As shown in another study, TiO2NPs did not show activity unless doped with nitrogen (N) and fluorine (F) [99]. Lastly, zinc oxide (ZnO) NPs are a prevalent oxide formulation [101,102,103,104,105].
Green synthesis is a method of synthesizing NPs that is more biologically- and environmentally friendly. This is done by using biogenic sources such as plants, bacteria, or fungi and incorporating their naturally occurring innately-antimicrobial metabolites or compounds into or onto NPs, acting as stabilizers or removing the need for harsh chemical components [106,107]. Sometimes, the color of the NPs may change when using plant extracts to synthesize NPs. For instance, a change to red/brown may indicate the incorporation of saponins and phenolic compounds as stabilizing agents providing the reduction power during production [108]. In another study, encapsulated Cymbopogon martinii (ginger grass) essential oil (extracted through hydro distillation) rather than using plant extracts during formulation in chitosan NPs to exploit the oil’s antifungal properties against Fusarium graminearum [106].
NPs made with biogenic sources [61,62,71,83,105,107,108,109,110] can have an enhanced effect compared to NPs made only chemically [105]. The biogenic formulation can increase the antifungal effect via increased reactive oxygen species (ROS) production, cell membrane disintegration, spore reduction, gene expression changes, and mycelial destruction [61,71,72,103,105]. Upon altering parameters, biogenic sources may produce AgNPs less effective than their chemical-only counterparts. An important parameter is a biological source; for example, the use of the fungus Penicillium chrysogenum as part of the synthesis of AgNPs was more effective than A. oryzae, although still less effective than the chemical-only process [62]. The formulation process may differ from organism to organism; however, the metal ions are typically entrapped by or surrounded by compounds released from the microorganism and undergo reduction.
Since fungal culture is a lab-controlled environment, the parameters for the formulation of NPs are also controlled [111]. For example, He et al. hypothesized that during gold NPs (AuNPs) formation, the bacterium Rhodopseudomonas capsulate secrets NADH- and NADH-dependent enzymes. During the electron transfer from NADH by NADH-dependent reductase, Au (III) ions capture electrons and, as a result, are reduced to Au (0) [112]. This mechanism of NADH- and NADH-dependent enzymes is likely also present in fungi, as extracellular filtrate of F. oxysporum strains used for synthesis contained NADH-dependent nitrate reductase enzymes [113].
Fungal biomolecules can behave as stabilizing agents during the formulation process and help achieve spherical particles [62,107]. For instance, during AuNPs formulation, it was found that biomolecules >3 kDa were not able to reduce Au (III) to Au (0). Interestingly, when comparing the AuNPs formation ability of various fungal extractions, it was found that while extractions containing large biomolecules were not capable of forming NPs, others containing small components would make unstable NPs, likely due to the small components’ inability to act as stabilizing agents. Combining different fractions developed stable NPs with slightly increased sizes of ~30 nm, compared to previous sizes of ~8–30 nm. Thus, it was concluded that biomolecules <3 kDa, such as glucose or amino acids, are involved in reducing the metal, and biomolecules >3 kDa, such as proteins, are involved in the stabilization [107].

7. Effect of the Size and Shape on the NPs Activity

Alongside the formulation process, the resulting shape and size are crucial to the activity. The shape is difficult to compare as most formulated particles are spherical, and amongst other forms, there are a variety of sizes and compositions. One study comparing CuNPs vs. Cu nanowires found that minimum inhibitory concentration (MIC) depended on the fungal strain. Still, mechanisms of action were different between the two samples [66] and are discussed in further depth later in this review. One study directly compared NPs shaped as spheres, cubes, and wires. Cubes were the most effective antifungal, followed by spheres and wires. Ag formulations were more effective than Au formulations, and the material seemed to matter more between spheres and cubes, but no wires. For example, Ag spheres were more effective than Au cubes in most cases. However, Ag wires were not. This may also be due to sizing, as the wires were 250–500 nm, while all other NPs were 30–50 nm. However, the shape was crucial for recovery as resistant isolates of Candida, when exposed to the nanocomposites, began to develop after 48 h of exposure to wires or spheres, but not with cubes, regardless of composition [76].
Size is often one of the most important factors to consider. Typically smaller NPs are found to have more potent antifungal activity than larger NPs [66,69,77,80,86,114,115,116], potentially due to the higher surface-area-to-volume ratio, which can increase binding at various target sites, facilitates diffusion, and have a smaller tendency to aggregate [77,86]. NPs can decorate fungal spores whereas microparticles cannot, and when considering this decoration ability, the size factor is more important than surface functionalization, material, or charge [117]. In addition, combining NPs can result in an antagonistic effect, which is hypothesized to be due to increased particle size, potentially through aggregation [69]. Nonetheless, some experiments find NPs of different sizes to have comparable activity. However, it is worth noting that although outside the typical NPs range, NPs as large as 729 nm have exhibited antifungal activity [85,108].
Capping agents can be used in NPs formulation for increased consistency and stability while preventing undesired changes such as aggregation. This is not to be confused with reducing agents, which are necessary for synthesis through electron donation, not stability. Two types of reagents containing O and N atoms are used. These atoms should be able to interact with the Cu (II) ion and then prevent aggregation through steric hindrance.
Capping agents function by creating an interface between the NPs and the medium in which the NPs are synthesized [118]. Capping agents can also be used to prevent oxidation, such as the positive detergent cetrimonium or cetyl bromide (CTAB) in the formulation of CuNPs to prevent oxidizing damage from the air [80]. CTAB is a quaternary ammonium salt containing a polar group (NCH3)4+ and a C16 hydrocarbon chain, which enhances surface activity. The prevention of Cu oxidation follows the two-phase adsorption of the CTAB on the surface of the CuNPs. In the first phase, the bromide anions adsorb on the surface of the positively charged CuNPs, with a posterior interaction between the (NCH3)4+ cation to the bromide anion. When CTAB molecules adsorb on the CuNPs, the long chain n-cetyl can cover the more available surface [118].
Polymeric coatings are of interest as they can control the release of metal species due to a stabilizing shell. Because NPs loading and release are linked with antimicrobial activity, these coatings can manipulate antimicrobial efficacy [92]. For example, GO capped with polyvinylpyrrolidone (PVP) was functionalized with polydiallyldimethylammonium chloride (polyDDA), rendering GO positive at 33 mV. This positive charge facilitated assembly with negatively charged AgNPs, forming a GO-AgNPs nanocomposite [92].
PVP is commonly used in the synthesis of AgNPs [62,65,115,118]. The lone pairs from the oxygen and nitrogen atoms are donated to the sp orbitals of the Ag ions, which form a complex that ultimately facilitates reduction and nucleation due to the increased negative charge [119]. Colloidal Ag solutions without stabilizing agents lead to low Ag content, and the addition of PVP usually increases Ag concentration, except at higher concentrations (6 g/L), where particle aggregation is induced through deterring diffusion and decreasing electrostatic repulsion [120].
The combinations of the capping agent PVP mixed with stabilizers such as sodium lauryl sulfate (Na-LS or SDS), sodium naphthalene sulfonate (Na-NS), or sodium dodecyl benzenesulfonate (SDBS) were tested in NPs [120]. Absolute zeta potential was >30 mV in SDS NPs and SDBS NPs, indicative of stability in the colloidal suspension [121]. At the same time, Na-NS NPs had a zeta potential of −18.2 mV and varying size results indicative of sedimentation. SDS NPs were most effective at fungal growth inhibition when tested against a mold colony [122]. On the other hand, AgNPs with SDS showed a significantly increased MIC and minimum fungicidal concentration (MFC) against yeasts when compared to no stabilizers, or AgNPs stabilized with the surfactants Tween-80, Brij, or PVP 360, although all surfactants increased the activity of AgNPs likely due to their ability to disrupt cell walls by themselves [59]. It is worth noting that SDS has antifungal properties alone as well due to bonds to the plasma membrane through lipid and protein interactions that lyse the cell [123]. This accounts for the synergistic effect with NPs as this would allow the penetration of NPs, increasing fungal sensitivity. SDS has also been shown to inhibit the ATPase activity by interfering with P-glycoprotein function [124], an antifungal mechanism used by select antifungal drugs such as itraconazole, hydroxyitraconazole, posaconazole, isavuconazole, anidulafungin, caspofungin, and micafungin [125].

8. Antifungal Classes and Combination with NPs

Currently, there are six antifungal drug categories, including four main antifungal drug classes: allylamines, azoles, echinocandins, and polyenes. Some literature will include the antimetabolite class [125,126] and, more recently, triterpenoids, such as ibrexafungerp, the first drug in the triterpenoid class [125,127].
Ergosterol and β-(1,3)-D-glucan are favored targets in antifungals, as these molecules are crucial for the survival of pathogenic fungi. These compounds are attractive because they are not produced by human cells (Table 7) [125]. Ergosterol is essentially the cholesterol equivalent in fungi and protozoa. Found in the cellular membrane, it is responsible for the membrane’s integrity and flexibility. Ergosterol is structurally similar to cholesterol aside from containing a double bond and additional methyl group in the alkyl side chain; this trans double bond means it is not saturated like cholesterol. There is also a second double bond at 7,8-position, alongside the 5,6-positioned double bond found in the cholesterol [128].
As mentioned earlier, the second compound, β-(1,3)-D-glucan, is a fungal cell wall component.
Antifungal resistance is a problem that is increasing worldwide. This can arise due to many mechanisms, including gene upregulation [129] for cell wall component synthesis [130] or efflux pump synthesis [131], modification of target site [132], or development of biofilm [133], amongst others. For example, the overexpression of ERG11 in yeast or CYP51 in mold confers resistance through the overproduction of lanosterol 14-α demethylase, aiding cell wall building and maintenance [130,134]. Because NPs have numerous mechanisms of action, fungi would have to evolve in multiple ways to acquire resistance while maintaining homeostasis and survival. As it is difficult to combat the simultaneous antifungal mechanisms of NPs, even though some are similar to antifungals (e.g., gene regulation), it is unlikely that fungi would be able to become resistant, at the very least, not at the same rate as the current antifungals used today.
NPs can be synthesized and combined with antifungals to increase one or both of their antimycotic capabilities [61,73,81,109,110,135]. One study decorated SiNPs with amphotericin B (AmB) to create particles with antifungal ability that could adhere to surfaces and be reused up to 5 times. The NPs alone was 3–33 times less effective than AmB on their own, but the addition of the antifungal gave the NPs an antifungal ability stronger than that of 10 nm colloidal Ag [73]. This, paired with the ability to coat surfaces, makes the particles worth further investigation, particularly in coating medical devices. While many studies have shown the synergism of NPs combined with antifungal medications, the effect can vary based on which fungal species are being treated [109,110].
The addition of antifungals to NPs can also benefit efficacy in unexpected ways. It can increase the roughness of NPs surfaces, which may account for mechanical damage or increased surface areas [81]. When antifungals are enclosed in NPs rather than coated, this can reduce their toxicity, such as in the case of AmB, where hemolytic activity in mammalian red blood cells was reduced from approximately 66% to 30% [136]. Therefore, combining NPs with antifungals can enhance activity, alter the morphology of NPs, and reduce cytotoxicity in human cells. Also, NPs on their own can have stronger antimycotic abilities than traditional antifungals [65].

9. Mechanisms of Fungal Cell Damage

9.1. Membrane Damage

Exposure to NPs causes changes in the fungal cell wall, including surface shrinkage, cell aggregation, pit and pore formation, and general deformation [55,60,65,70,95,104,137]. Through microscopy, it has been detected that NPs may have direct contact and embedment within fungal cell walls during adsorption, which induces morphological change. Inner membranes also suffer distortion, with altered organelle disposition (such as increased intracellular vesicle and vacuole count) and decreased cytoplasmic content [65,86,114]. Severe damage causing pits or cell wall rupture is not guaranteed upon exposure to AgNPs. Still, even without this damage, AgNPs can be found surrounding and adsorbing onto cell walls. Ag dots in the cytoplasm can be detected using energy-dispersive X-ray spectroscopy (EDS). This could be smaller NPs, but also Ag ion penetration that ends in the formation of NPs intracellularly through the organic compound reduction [138]. Difficulty differentiating the cytoplasm/cell wall/plasma membrane borders after exposure to NPs is expected as the NPs disrupt these structures and blur boundaries [70,86,138]. The loss of intracellular structure fills the cell with cytoplasm and a seeming lack of organelles.
Observations of cell wall components show changes upon exposure to NPs. For instance, chitosan NPs encapsulating C. martinii essential oil and ZnO NPs caused a dose-dependent decrease in ergosterol and chitin, respectively [104,106]. AgNPs caused alterations of phosphatidylcholine-to-phosphatidylethanolamine ratios in treated cells, causing a loss of membrane integrity and cell function. Additionally, fewer C18:2 phospholipids were present in treated cells and inhibited desaturation ability [61]. Other nanocomposites, such as MgO/CuO sodium alginate microspheres with antifungal nystatin, showed that exposure to C. albicans damaged the cell membrane and leaked intracellular components. After 24 h, 94% of cells were in the late apoptotic stage, and proteins and DNA had seeped through the broken cell wall [81]. AgNPs also produce this effect, where fungal cell-free filtrate contains DNA after exposure [105].
The cell leakage in C. albicans cells leads to increased intra- and extracellular levels of glucose and trehalose, which can have protective functions during stress conditions and function as a carbon source [139]. Similar effects were found compared to AmB [55]. Through Raman spectroscopy, ZnO NPs were also found to increase the levels of carbohydrates and nucleic acids in the pathogenic fungus Botrytis cinerea, potentially indicating the need for the protective mechanism that carbohydrates (e.g., trehalose) can provide for the cell [102,139]. Moreover, when fungal growth was inhibited entirely, carbohydrate and nucleic acid levels decreased [102]. Interestingly, the exposure to ZnO NPs did not change protein and lipid levels in fungi, indicating that perhaps the cell wall was not impacted, and the particles may have been internalized, causing intracellular effects more severe than the cell wall/extracellular effects, or the cell wall was affected. Still, mechanisms for cell wall synthesis were altered to compensate for any loss.
Increased glucose levels were also seen in C. albicans and S. cerevisiae after the exposure to AgNPs; however, the decrease in glucose uptake was likely associated with a G1 cell cycle delay or ROS production. When treated with N-acetylcysteine (NAC), a ROS scavenger that increases antioxidant-reduced glutathione levels (GSH), glucose uptake was restored in C. albicans only. Therefore ROS production might play a role in reducing glucose uptake by C. albicans [115]. Combining the treatment of AgNPs with a glycolysis inhibitor such as bromopyruvate exhibits a synergistic effect on cell death in C. albicans but not in S. cerevisiae [115]. These effects could also be because C. albicans is pathogenic, whereas S. cerevisiae is not. Then if any pathways for pathogenicity and virulence are being targeted, only one strain will be affected.
ROS partake in lipid peroxidation, which can induce cell wall damage. N, F co-doped TiO2 NPs capture light and generate ROS, mainly hydroxyl radicals. These radicals attack the monomers of the cell wall, cleaving the glycosidic linkage and creating pores, leading to fungal death. Ethidium bromide (EtBr) was used to confirm cell wall damage, as EtBr can penetrate the cell and intercalate with DNA to give off signals under excitation only if the cell wall has been damaged [99].
Similarly, CuNPs induced significant cell wall damage in yeast cells. The cytoplasm membrane was detached from the cell wall, and organelles appeared swollen. Along with this cellular collapse, fimbriae (hair-like protrusions on the cell wall) decreased in number. Other Cu nanocomposites, such as Cu nanowires, also decreased fimbriae quantity. While yeast cells experienced enlargement and some cytoplasm leakage with cell wall disruption, most cells conserved their shape and cell wall integrity [66]. This indicates that although the MIC of nanowires and NPs were comparable, the mode of antimycotic activity was not the same. One of the reasons can be attributed to the presence of sharp edges in the NPs, suggesting that the cell wall damage occurred due to mechanical damage. Thus, although cell wall damage is common amongst NPs, this happens in different ways, including ROS production, gene regulation resulting in changes in cell wall components, mechanical damage, internalization, and removal of protective compounds.

9.2. DNA Interactions

Smaller NPs may cause fluid-phase endocytosis, bypassing the need for any cell wall damage [65]. It was shown that AgNPs could damage the cell wall and enter the cell, but at a smaller size can also enter without much damage. The antimycotic injury occurs at an intracellular level, including mitochondrial fragmentation, ribosome depolymerization, and chromatin damage. This triggers the production of multivesicular bodies [60]. Indeed, cell wall damage can cause DNA leakage out of the cell, but once NPs are inside the cell, some can intercalate with nucleic acids intracellularly.
AgNPs caused DNA condensation and fragmentation in C. albicans nuclei. These damages were validated by adding thiourea and its protective abilities against oxidative stress [140]. This indicates that although Ag ions can bind to DNA (negatively charged), which may play a factor in DNA destruction, oxidative stress also has a role. Additionally, damage can occur through the contact of NPs with DNA. AuNPs caused DNA damage directly by nuclear condensation and DNA fragmentation in C. albicans, which may account for cellular dysfunction and apoptosis [140]. After ZnO treatment, the increased nucleic acid production in B. cinerea may indicate the mechanism of NPs known in bacteria, where NPs can interact with the gyrase-DNA complex. This would encourage the fungal cell to produce more nucleic acid as compensation [102].
On the other hand, CuO NPs induced dose-dependent DNA-strand breaks with an effect increasing after an additional week of incubation. Two factors enhanced this effect: (1) the exposure of Cu2+ rather than NPs, and (2) the effect increased in fungi isolated from streams without metal pollution vs. streams with metal pollution [141]. The efficacy of the Cu ion form may be due to size, and effectiveness based on fungal strain is likely due to metal adaptation.
In summary, the fragmentation of DNA, which is enclosed in the nucleus, implies that the NPs should cross the nuclear membrane or produce significant damage in the intracellular membranes that would allow them to be in contact with the DNA. Moreover, it might also be the results of the ions released by the NPs that can affect the nuclear membrane at a distance.

9.3. Ion Release

Ion release has increased antimicrobial activity with an extended-release over time, improving results [63]. Some studies have shown that ions are more toxic than their NPs counterparts [141,142]. This could be due to size, which may facilitate penetration into cells, or the ability to complex with other biomolecules, such as proteins, nucleic acids, negatively-charged lipids, etc. In arsenic-resistant Aspergillus ssp., thiol compounds were found to play a significant role in detoxification. Because Ag ions are complexed with thiols, ion release may interfere with a fungal detoxification process [143], as thiols in the cytoplasm are essential to keep reducing conditions in this compartment. Thus, ROS produced by the effect of ions released by the NPs exposed to an oxidative environment would exhaust the levels of thiols, altering the intracellular redox potential. As a result, cells might activate the apoptosis program [144]. Another study suggested that producing Cu2+ paired with cell wall destruction through mechanical means allows ions to behave as a biocide, interacting with cellular compounds [66]. However, a study found that using NaCl neutralized the efficacy of Ag, indicating that ions are partially responsible for its antifungal ability [145]. This neutralization can result from a change in the electrolyte balance with the activation of different ion pumps. But this phenomenon might not happen in applying NaCl as an accompanying therapy because it might affect the isotonic balance and cause indirect damage to other eukaryotic cells.
To account for ion activity, supernatants of NPs were tested, and it was found that SiO2 NPs toxicity was due to NPs alone, Al2O3 NPs and SnO2 NPs toxicity was due to a combination of NPs and ions, and Mn3O4 NPs toxicity was primarily due to ion activity [97]. Ultimately, fewer studies about ion release look at fungi rather than bacteria. Still, from the ones available, it is evident that ion release is sometimes an NP mechanism of action for antimycotic activity.

9.4. Damage to Hyphae and Spores

NPs can have severe impacts on fungal hyphae and spores. Fungi treated with GO-AgNPs, AgNPs, MgO NPs, ZnO NPs, or CuNPs showed hyphae deformation, appearing distorted and shrunken [63,69,95,96,105,114]. NPs change the growth patterns, clumping and thinning hyphal fibers [60,70,86].
Interestingly, even when CuNPs did not impact fungal growth, hyphae still appeared damaged [96]. As a result of hyphal damage, NPs can inhibit mycelial growth, often dose-dependent manner [65,86,95,112,146]. Sub-inhibitory AgNP treatment in C. albicans inhibited the growth of mycelia after morphogenesis was triggered. Observations of the mycelia showed that it did not extend nor form around the presence of AgNPs, but untreated control had a healthy mycelial formation [147]. SiO2 NPs could reduce mycelial radial growth up to 100% at 100 µg/mL exposure. They behaved dose-dependently, impacting mycelium fresh and dry weight [63]. ZnO NPs formed bulges on B. cinerea hyphae surfaces, inhibiting the growth [102]. Quantum dots (QD) and superparamagnetic (SP) NPs were found to interact with fungal hypha. However, the QDs were internalized, likely due to their small 13.5 nm diameter, as opposed to the 100–150 nm diameter SP NPs which seemed to adsorb onto hyphal cell surfaces. QDs were initially (3 h incubation) distributed evenly throughout the cell. Still, at a 6 h incubation time, they were seen to accumulate, indicating some intracellular processing of NPs, although there was also evidence of QD exocytosis [148].
The effect on spores and their germination contributes to the antifungal efficacy of NPs, alongside hyphal deformation. Penicillium expansum and its external mycelial colonies treated with ZnO NPs showed damaged conidia and distorted structure, inhibiting conidial germination and development and, ultimately, fungal growth [101]. Another study showed that spores incubated until logarithmic mycelial growth were observed with AgNPs and were shown to have a decreased mycelium growth rate upon germination, with a growth rate depending on the concentrations of AgNPs. However, this was significant in higher concentrations of 5–10 ppm Ag solutions, with a much smaller difference in the <2.5 ppm range [149]. N, F co-doped TiO2 NPs exposure inhibited spore germination in a dose-dependent manner. It resulted in damaged mycelia showing ROS production around its surface, unlike the NP-treated samples in the dark or non-NP-treated samples in the light [99]. F. graminearum spores suffered damage upon exposure to NPs that collapsed, wrinkled, and distorted structures, giving them a rough surface. The MFC treatment showed more excessive damage than the MIC treatment, but both had apparent morphological destruction of the two types of spores produced by the fungus (macro- and microconidia) [106]. Other NPs, such as AgNPs, reduced the size and septation number of micro- and macroconidia [146].
AgNPs exposure decreased mycelial growth in five fungal species but not in Mortierella alpina, which had its growth significantly stimulated through exposure at all tested concentrations. The authors suggest that the known production of polyunsaturated fatty acids by M. alpina may act in a chelating fashion, as toxicity from metals often involves a mechanism of lipid peroxidation through the production of free radicals [150,151]. Conidiophore (a branch that produces spores) survival varied amongst the remaining five species and depended on the concentration of NPs and conidiophore location in the mycelium [150]. Despite how this shows that spore survival and germination can be based on certain conditions, generally, spores suffer a change in morphology upon NPs exposure, appearing damaged and shriveled, often coated in NPs [63,69,96]. NPs treatment typically reduces the germination rate, and asexual reproduction can be halted altogether [88,95,114].
There are various theories on how NPs affect spores and what matters in the formulation. Spores possess an overall negative charge on their surface due to the presence of hydrophobins, or low molecular mass proteins secreted by fungi with the ability to self-assemble into amphipathic layers [152]. Nonetheless, negatively charged polymer NPs of comparable sizes were bound more efficiently than positively charged ones, and the increasing positive charge did not enhance attachment or adsorption. This behavior persisted in the ion-free buffer, removing the potential impact of ions; however, the composition of the NPs matters, e.g., SiO2 NPs readily coat the hydrophobin layer, whereas CeO2 NPs agglomerated in buffer and adsorbed as clusters [116].
Size is necessary, with smaller NPs coating spores more readily. The small NPs experience enhanced binding as there are more surface points since they fit better into conidial surfaces [116]. GO-Ag nanosheets were found to wrap around spore cell walls inhibiting transport and nutrient absorption. Fungal spores mediated GO reduction in oxygen-containing functional groups, likely responsible for GO-AgNPs efficacy. Increasing GO-AgNPs or AgNPs concentration decreased the germination rate of spores [63]. Other studies using metal oxides also found that exposure inhibited sporulation [69,101]. For example, AgNPs reduced spore germination, and germination was halted when formulated biogenically, involving the incorporation of NPs of Trichoderma viride metabolites. It is hypothesized that direct interaction with fungi can inhibit conidiospore germination and sporulation [114]. NPs associate with spores immediately upon incubation but do not affect colloidal stability. Kinetic analysis has revealed that NPs coating occurs rapidly (under 30 s), and pH and temperature ranges from 4–8 and 4–55 °C, respectively, did not impact the coating [116].
Contrarily, one study found that NPs coating does not impact spore vitality and reduces its sensitivity to defensins such as human neutrophil peptide 1 and human β-defensin 3, meaning that fungal survival was facilitated by circumventing components of the innate immune system [116]. This could be due to activated defense mechanisms. For example, increased melanin levels have been found in hyphae after exposure to NPs [105]. Melanins are dark polymers that function in fungi to protect the organism from external stresses. Melanin is associated with virulence, protection, pathogenicity, and survival. It can bind antifungal drugs, reduce their effects, and trigger the host complement system [153].

9.5. Effect on Biofilm Formation

Fungal biofilms are colonies of sessile cells embedded on a surface. Fungal cells capable of biofilm formation exhibit an increased resistance when in their biofilm morphology. The exudation of extracellular polymeric substances (EPS), increased expression of resistance genes, and altered growth rate contributes to virulence and drug tolerance. Consequently, biofilms are a major clinical concern, particularly in C. albicans [154]. Some fungal species, such as C. albicans, can form more mature biofilms containing extracellular matrix (ECM) with improved structural integrity than other species, such as S. cerevisiae. However, it is not morphology specific, and both yeast and filamentous cells are capable of biofilm development [133].
NPs exposure has been shown to inhibit the formation of biofilm [81,137,155]. Hyphae development is crucial for biofilm formation and adherence, which is required for pathogenesis and colonization. NPs inhibit filamentation, leading to an absence of hyphae and prevention of biofilm development. Cell wall disruption is the driving force behind this inhibition. NPs can also impact pre-formed biofilms and deposit onto EPS, which are essential for structural integrity [137]. It is not unexpected that biofilms would interact with NPs, as it has been shown that Candida biofilms may adsorb cations of various metal species onto the ECM, which facilitates resistance. Biofilms are 65 times more tolerant to metal exposure than planktonic cultures. Heavy metals Cu2+ and Ni2+ form metal-chelator precipitate on ECM, proving that adsorption occurs for the tolerance [156].
Using Ag-doped hydroxyapatite NPs against C. albicans biofilms decreased biomass values in the biofilms, showing an impact on ECM production. However, this impact did not remain in the colony-forming units (CFU) count, indicating recovery. Additionally, sub-inhibitory NPs concentrations increased biofilm formation, perhaps due to gene upregulation related to virulence factors [68]. AgNPs exposure was able to reduce C. albicans biofilm biomass and CFU number at both intermediate (24 h) and mature (48 h) development phases [157] but demonstrated higher efficacy in biomass reduction when applied to adhered cells vs. pre-formed biofilms, where the incubation time is approximately 2 h, and 48 h, respectively [57].
A study encasing AmB in poly(lactic-co-glycolic acid) NPs found significant C. albicans biofilm destruction when the treatment was combined with 15 min of 42 kHz of ultrasound irradiation compared to AmB or ultrasound alone. The treatment reduced biomass and activities of proteinases and phospholipases (known to be virulence factors) and significantly changed biofilm morphology. The same test was done on a catheter biofilm using a rat model, and after seven days of treatment, the biofilm on the catheter surface was eradicated. After treatment, biofilms appeared to have a reduced thickness, with channels and pores throughout. COMSTAT analysis showed that the different treatments had significantly altered thickness, diffusion distance, textural entropy, and areal porosity. Regardless of the use of ultrasound, NPs treatment shortened mycelia and reduced ECM on catheters. Still, after applying ultrasound which causes biofilm dispersal, there was a complete biofilm inhibition [136].
Contrarily, another study found that AgNPs treatment induced EPS-rich biofilm formation and hypothesized that this might be due to the known release of sugar which leads to an environment appropriate for the biofilm formation [55,65,133]. However, AgNPs would adhere to the biofilm and cellular surfaces, ultimately leading to cell damage and growth arrest [65].

9.6. ROS Generation

ROS are a family of oxidants often formed as a metabolic by-product and include compounds such as hydrogen peroxide, superoxide anions, and hydroxyl radicals [158]. Mammalian cells in non-viable environmental conditions produce excessive levels of ROS, which cause cellular injury and ultimately end in the arrest of cell growth or cell death [159]. GSH is a thiol in yeast and filamentous fungi crucial for cell function and stress response. GSH has anti-oxidizing properties and reacts with ROS to protect the cell. GSH oligomers termed “phytochelatins” are involved in heavy metal detoxification in fungi through sequestering metal ions or the production of ROS [160].
One of the main antimicrobial characteristics of NPs is attributed to the generation of ROS. For instance, ZnO NPs generate ROS when UV light radiates and oxidizes GSH to form a disulfide from the thiol groups, making glutathione disulfide (GSSG) [103]. Many other NPs induce ROS generation, and activity is often tested by adding a ROS scavenger and observing if NPs effects can be reversed [63,70,103,105,114,140]. SiO2 NPs, Al2O3 NPs, Mn3O4 NPs, and SnO2 NPs generated intracellular ROS that reduced metabolic activity and cell viability. This effect was reversible primarily by adding a free radical scavenger, the ascorbic acid [97]. Another study used the nitroblue tetrazolium reduction assay to detect superoxide radicals in Alternaria brassicicola. It showed an accumulation of formazan in AgNP-treated fungal mycelia, indicating that the exhibited structural abnormalities are due to the activity of superoxide radicals [105]. It is hypothesized that structural defects create something termed “disordered sites,” which ultimately result in a change in electronic distribution that, through electron-hole reactions, can generate hydroxyl radicals and superoxide anions [68]. ZnO NPs decreased C. albicans viability dose-dependently. When histidine, a known quencher of ROS such as singlet oxygen and hydroxyl molecules [161], was added to the incubation at a 5 mM concentration, the antifungal effect was almost entirely undone. Electron paramagnetic resonance provided further evidence of a ROS-mediated decrease in cell survival. It showed that a reduced C. albicans produced ROS upon histidine addition [162].
Decreases in membrane permeability lead to intracellular ROS generation [61]. The increase in lipid peroxidation is hypothesized to be due to ROS production, as this is a known mechanism [105,106,163]. However, a decrease in intracellular ROS also impacts. AgNPs cause a reduction in the secretion of aflatoxin B1 (AFB1), a poisonous carcinogen and mutagen, in Aspergillus flavus. When this fungus is cultured, AFB1 is increasingly released, but a sub-inhibitory treatment of AgNPs reduced AFB1 from 66.23 ng/mL when untreated to 28.96 ng/mL with treatment. Simultaneously, superoxide is released from fungal mycelia. It is proposed that the released superoxide causes a reduction in intracellular ROS levels, which triggers suppression of the aflatoxin biosynthesis [89].
ROS release can play a role in antifungal effects, but this is not always the case. One study found that exposure of AuNPs to C. albicans did not alter ROS production and hypothesized that the NPs damage and cell death occurring post-exposure must not be from ROS influence. Furthermore, antioxidant treatment did not protect cells from apoptotic processes [164]. Another study found that C. albicans experienced an increase in ROS production during the start of culture, but S. cerevisiae did not. AgNPs were exposed to S. cerevisiae wild type and mutants containing deletions of stress response genes. The AgNPs inhibitory effect was comparable amongst all strains, meaning ROS did not strongly impact the lack of growth [115]. A third study found that the level of superoxide in S. cerevisiae cells decreases upon AgNPs exposure, potentially because Ag ions are quenching the ROS [165,166].
There may also be only certain types of radicals whose production is induced. QDs and SP NPs significantly increased superoxides after incubation. Still, QD incubation showed a significant decrease in H2O2 production, and the H2DCFDA signal (representative of ROS accumulation) was not increased in either sample. Therefore only the superoxide anion was raised, and other ROS were unaffected or decreased [148]. This may be due to the materials, as the QDs were made of a CdSe/ZnS core, and the SP NPs were made of SiO2, which has been shown in plants to behave protectively and lower ROS [91]. N, F co-doped TiO2 NPs generate ROS under visible light around the fungal cell, damaging the cell wall and leading to death. When F. oxysporum inoculated with TiO2 NPs is grown under visible light, fungal colony growth is halted, whereas it usually progresses in darkness. Under the same conditions with p-benzoquinone, a superoxide radical scavenger, and EDTA, an electron-hole scavenger, growth inhibition still occurred, meaning that superoxides and electron holes were not involved in the process [99].

9.7. ROS Impact on Mitochondria

One of the most critical ways ROS can affect the fungal cell is through the mitochondria. AgNPs treatment decreased mitochondrial inner membrane potential (ΔΨm) in C. albicans, comparable to the positive control using H2O2. However, the decrease was minimal when cells were treated with AgNPs and thiourea (a hydroxyl radical scavenger). This was tested using DiOC6(3) and JC-1 dyes, and both confirmed that AgNPs decreased ΔΨm in a magnitude comparable to H2O2. At the same time, adding thiourea removed this effect, indicating that preventing hydroxyl radical formation maintains mitochondrial membrane balance, potentially containing an apoptosis [140]. AuNPs caused a mitochondrial increase in Ca2+ levels and mitochondrial dysfunction, and as a result, mitochondrial mass in C. albicans cells was increased [164]. Ca2+ excess in mitochondria plays a part in apoptosis as it can open channels, changing permeability and thus the mitochondrial mass. The homeostasis disturbance leads to reduced mitochondrial membrane potential and cell death by apoptosis. AuNPs exposure also increased cytosolic cytochrome c levels and decreased cytochrome c levels in C. albicans mitochondria, making cells no longer able to partake in the electron transport [164].
Cytochrome c resides in the mitochondrial membrane, transferring electrons in the electron transport chain. When the outer mitochondrial membrane experiences a change in permeability (for example, if damage causes swelling to the high surface area of inner cristae, this can damage the outer membrane), cytochrome c will release into the cytosol during early apoptotic phases, triggering a signaling cascade. This has been proven in mammalian cells, and the process is similar in fungi, although it has been found that in yeast, loss of cytochrome c is not necessarily required for programmed cell death [167]. Additionally, the signaling cascade contains caspases that are not in fungi. However, similar compounds termed “metacaspases” exist and were found to be increased upon AgNPs exposure. The addition of thiourea decreased metacaspase levels indicating that AgNPs induce apoptosis with an increase in metacaspases. However, oxidative stress is involved, and once removed, apoptosis is reduced [140]. AuNPs have also increased the metacaspase activation [164].
Flow cytometry has demonstrated mitochondrial membrane potential reduction along with the mitochondrial release of cytochrome C upon AgNPs treatment. Fluorescence microscopy showed apoptosis through nuclear fragmentation, metacaspase activation, and phosphatidylserine externalization. Additionally, ROS levels were increased. The AgNPs accumulate intracellularly in C. albicans, increasing radical hydroxyl levels, a factor in yeast apoptosis. Adding thiourea decreases apoptosis mediated by mitochondrial dysfunction, suggesting that the hydroxyl radical is critical for apoptosis [140].

9.8. Gene Regulation and Protein Levels

Gene expression and protein levels can be altered upon NPs exposure. As discussed previously, gene up- and downregulation can confer resistance to antifungals in fungal species. So, these changes would be expected upon interaction with a compound with antimycotic activity.
Multiple studies investigated the effect of NPs on the S. cerevisiae genome. Horstmann et al. found that exposure to AgNPs altered gene expression in 1845 genes: 1077 upregulated and 768 downregulated [165]. Of 1077 genes, 651 were involved in nitrogen compound metabolism. Other upregulated categories included ribosome biogenesis, rRNA processing, translation, and translational initiation. DBP2 and FAF1 genes involved in rRNA processing were upregulated 7.9-fold and 7.2-fold, respectively, and the most upregulated gene was found to be CTR1, responsible for copper uptake in low copper environments. S. cerevisiae does contain metallothioneins that uptake Cu in the form of Cu(I); thus, it is possible that Ag can trigger Cu-related mechanisms [168]. The majority (25.3%) of downregulated genes were involved in the single-organism metabolic process. While many of the most downregulated genes had an unknown function, other top categories included lipid metabolic processes (such as genes involved in ergosterol synthesis: ERG3, ERG5, ERG6, ERG11, ERG25, ERG28), transmembrane transporters, and mitochondrial cellular respiration [168]. AAC3 and MPC3, involved in the transport of ADP/ATP and pyruvate in the mitochondria, were downregulated, along with ACO1, CIT3, and ISF1, which are involved in mitochondrial respiration as well. Other downregulated genes included those regulating oxidative stress responses (GRX6, TSA2, VHR1, ZTA1), regulating osmotic stress response (CIN5, NRG2), and involved in spore cell wall formation (GIP1, GSC2, OSW2, SPO73), and involved in cell wall mannoprotein production (TIP1, TIR1, TIR2, TIR3, TIR4, DAN1), which included a protein that experienced a 174-fold downregulation (DAN1). In conclusion, the authors hypothesized that sublethal AgNPs levels stimulate ribosome biogenesis in the nucleolus and inhibit mannoprotein and glycan turnover, negatively impacting cell wall integrity. The decrease in sugar transporters and plasma membrane destabilization may also impact cytosol sugar levels. Finally, AgNPs damaged mitochondria, halting proper respiratory chain function, generating ROS, and inducing apoptosis [165].
Babele et al. investigated gene expression in S. cerevisiae upon exposure to the ZnO NPs treatment [104]. They found upregulation of the chitin synthesis genes CHS1, CHS3, and CHS5, indicating damage to the cell wall. Further analysis was done on genes involved in cell wall integrity, redox homeostasis, and the HOG signaling pathway (High Osmolarity Glycerol, involved in cell response to hyperosmotic stress). The HOG pathway consists of a MAP kinase cascade and works in conjunction with the cell wall integrity (CWI) pathway to respond to the cell stress [169,170]. When treated with ZnO NPs, growth was inhibited in yeast with the following null mutations: ROM1, MEK MKK1, MEK MKK2, MEKK BCK1, MAPK SLT2, RLM1, and GAS1, suggesting that the CWI pathway is essential for ZnO NPs tolerance and that there is involvement with MAP kinase [104].
To investigate lipid metabolism and homeostasis, fatty acid synthase genes FAS1 and FAS2 were examined, and it was found that their proteins accumulated in cytosol upon ZnO treatment, unlike the untreated control, where the proteins were evenly dispersed. Similar trends of accumulation in foci were observed when studying mitochondria and HSP1, a heat-shock protein necessary for stress management. Mitochondrial PSD1 was found localized in the cytosol, supporting the finding of mitochondrial dysfunction while also indicating lipid mechanism defect. So, ZnO NP-treated cells experienced damage to both the mitochondria and endoplasmic reticulum (ER), further evidenced by the upregulation of the IRE1 gene and an accumulation of activated HAC1 forms. IRE1 is a sensor for the unfolded protein response (UPR), an ER stress response related to faulty protein production, and HAC1 is concomitant with UPR activation. The mitochondrial/ER protein complex, ERMES, appeared damaged as both components were affected. The ER is necessary for protein folding, lipid synthesis, and redox homeostasis [171,172,173,174]. Because it has been hypothesized that the ER and cell wall stress responses are intertwined with the UPR and CWI pathway [172], it is not surprising that both pathways are impacted by NP exposure. Finally, treated cells showed an increase in cellular lipid droplets and the presence of ATG8 in vacuoles, indicating the start of autophagy. Ultimately, this study found that ZnO NPs treatment impacted stress mechanisms involving the cell wall, the ER, and the mitochondria, leading to cell death.
There is a possibility that specific genes are regulated not because of the NPs but rather by the released ions. A transcriptomic analysis using AgNPs and Ag ions treatments in S. cerevisiae found a ~3% and ~5% change in genome expression, respectively [142]. Most up- and down-regulation in known gene functions at earlier incubation stages (120 min) included heat shock, cell wall proteins, oxidative stress, and metabolism. Later (210 min), incubation affected genes like Ag ion exposure, suggesting AgNPs toxicity evolves to be similar to Ag ion toxicity over time. Targeted genes involved iron retention, cell division, stress, and transporter proteins. The most upregulated genes for both AgNPs and Ag ions were CUP1-1 and CUP1-2 (responsible for metallothionein production involved in Cu and Cd resistance, known to be upregulated in response to cytosolic Cu increase), PHO89 (Na+/Pi cotransporter), and heat shock genes (e.g., HSP12, HSP26) [142].
A. brassicicola was also tested with AgNPs, and its stress enzyme activities (verified by calculating the content of thiobarbituric acid reactive substances) of superoxide dismutase (SOD), catalase, ascorbate peroxidase (APX) and glutathione peroxidase (GPX) decreased [105]. All of these enzymes play a role in protecting the cell from oxidative damage. The same study used reverse transcription polymerase chain reaction (RT-PCR) to examine the different gene expressions between AgNPs made chemically (CSNPs) and AgNPs made using T. viride (BSNPs). Both NPs heavily downregulated ATM, responsible for transmembrane protein synthesis as a part of intracellular redox regulation, indicating issues with this homeostasis. Both NPs also heavily upregulated SLT2, a gene involved in a signal cascade for cell wall maintenance, indicating cell wall damage. Interestingly, genes NPS2, NIK, and AMK experienced downregulation upon BSNPs exposure and upregulation upon CSNPs exposure [105]. The varied expression of these genes, involved in various tasks such as spore survival, virulence, and cell wall and osmotic homeostasis, show that the fungi can attempt recovery when exposed to CSNPs, while BSNPs exhibit fungicidal behavior.
The different enzymatic levels in five fungal strains isolated from streams without metal pollution vs. with metal pollution upon CuO NPs exposure were studied to investigate how metal pollution can alter the gene expression regulation [141]. Glutathione reductase (GR), responsible for catalyzing GSSG to GSH, and GPX, responsible for catalyzing GSH to GSSG, increased with the CuO NPs exposure [141,173]. Fungi isolated from non-polluted streams had higher GPX and lowered GR activity, especially after prolonged exposure. It is possible that peroxide radicals were not scavenged as thoroughly by fungi from non-polluted streams, so they experienced lower levels of GSH. In comparison, fungi from polluted streams maintained higher GSH levels. SOD activity also increased concomitantly with CuO NPs in all fungi, more severely than other antioxidant enzymes. The level was higher in fungi isolated from metal-polluted streams [141]. It has been previously shown that aquatic fungi exposed to metals including Cu produce compounds such as thiols, small peptides (likely glutathione and phytochelatins), and metallothioneins, the latter of which likely behave as ROS scavengers [174]. Investigating the difference in metal pollution showed that the ionic Cu form is more impactful than its NPs form, perhaps due to bioavailability. Ultimately, metal adaptation in fungi can lead to mechanisms of metal tolerance, mainly dealing with oxidative stress, that can transfer to tolerating NPs exposure [141].
Other tolerance mechanisms include efflux pump production, melanin synthesis, and biofilm formation. This is also reflected in gene regulation. AgNPs exposure downregulated the expression of genes polyketide synthase (PKS1) and scytalone dehydratase (SCD1), both involved in melanin biosynthesis, by 6.47-fold and 1.8-fold, respectively [88]. CDR1, CYP51A, HSP90, and ERG11 were significantly downregulated at antifungal concentrations in both fluconazole-resistant and –susceptible C. albicans strains when exposed to SeNPs [71,72]. These genes have been linked to resistance mechanisms [175,176,177]. CDR genes are responsible for efflux pumps, and CDR1 is often overexpressed in resistant strains [178], ALS (agglutinin-like sequence), SAP (secreted aspartyl protease), PLB (phospholipase B), and LIP (lipase) gene families have all shown involvement in C. albicans biofilm formation, along with HWP1 (hyphal wall protein) [178]. HWP1 is a GPI-linked protein also associated explicitly with hyphae. It has been shown to confer adherence and virulence [179]. ZnO NPs decreased HWP1 gene expression in C. albicans isolates [155]. The ALS gene family encodes GPI-linked cell surface glycoproteins. This is important for host cell invasion. ALS3, in particular, is upregulated during infection [179]. Sub-inhibitory ZnO NPs exposure downregulated the expression of ALS1 and ALS3 in fluconazole-resistant C. albicans strains significantly more than fluconazole treatment [74]. Secreted aspartyl proteinases (SAP) are extracellular hydrolytic enzymes coded by 10 SAP gene families, known to be produced by many pathogenic Candida species, including C. albicans. SAP is involved in virulence and is responsible for evading immune response, acquiring cell nutrition, and facilitating invasion. SAP is produced in higher concentrations in more resistant strains [180]. C. albicans treated with subinhibitory levels of ZnO NP decreased SAP1, SAP2, and SAP3 gene expression 2.44-fold, 5.55-fold, and 2.6-fold, respectively. Nystatin showed similar decreases at 4.16-fold, 3.12-fold, and 4.8-fold, respectively [76].
The global damages caused as a result of the exposure of the fungal cell to NPs is pictured in Figure 5.

10. Experimental Conditions

The parameters and conditions under which an experiment is done can alter results. Specific NPs formulations may require special considerations during experimentation, such as light exposure or an applied magnetic field. Still, factors such as the media used also change the results and can do so significantly. For example, an experiment tested the effects of AgNPs on Fusarium culmorum spores in three different media: nutrient-rich potato dextrose agar (PDA), nutrient-poor PDA, and agar. The nutrient-rich PDA required an incubation time with AgNPs of 24 h for antimicrobial activity and found AgNPs concentration variation to be insignificant. Nutrient-poor PDA found AgNPs significantly reduced mycelial growth regardless of incubation time or concentration. In agar, a low 0.12 ppm concentration increased mycelial growth, while 1.25 ppm was comparable to the control, and only 2.5 ppm slightly decreased. Additionally, mycelium in the agar media did not form spores [149]. Compared to malt extract agar (MEA) or corn meal agar (CMA), PDA found more promising results, with 90–100% inhibition of mycelial growth rate across 18 fungal species [87].
This effect of media is unsurprising as it is known that fungi will change metabolism depending on external stresses. If fungi are carbon deprived, this triggers GSH activity and adapts the organism to the oxidative stress [159]. C. albicans reacts to environmental conditions, which can trigger morphogenesis. The addition of serum to the culture, change in pH, or change in incubation temperature can all promote the transition of C. albicans yeast cells into filamentous forms [161]. Growing yeast cells in <0.1% glucose changes cell wall structure, decreasing chitin and β-glucan count and decreasing mannan chain length [181]. Compounds in complex media (e.g., glucose) can also behave as reducing agents, creating NPs from metal salts. As a result, during fungal use for NPs formation, the media within which the fungi are cultured can affect NPs characteristics [107].
Incubation time is another factor that changes results. Fungal growth during the logarithmic growth phase post-inoculation is impacted by both the growth medium used and incubation time (with and without NPs) [149]. It is also worth noting that there is no standardized incubation time to compare all experimental assays, and almost every study testing any mechanism (e.g., cell wall interaction or mycelial growth inhibition) across multiple time points will report varying results, usually either increased efficacy overtime or recovery [81,114,135,141,142,145].
Fungal species used in experiments can also produce different results for various reasons. Some strains acquire resistance and thus survive using mechanisms other than their non-resistant counterparts, many of which are discussed in this review. If NPs target any of these mechanisms, the outcome differs based on strain. There is also a difference between pathogenic and non-pathogenic fungi, and increased AgNPs efficacy against pathogenic fungi indicates that some AgNPs formulations function by targeting a pathogenic pathway [115]. Lastly, fungal growth patterns can also alter results; fungi that grow in a more dense matter may be more sensitive to NPs due to the potential of increased exposure [102].

11. Potential Applications of the Use of Metallic NPs with Antifungal Activity

11.1. Agriculture

One of the main challenges of farmers is to control the production of their produces. Phytopathogenic fungi decimate significant farming areas, and their control burden agriculture. Although fungi are controlled with synthetic chemical products, their indiscriminate use creates environmental problems, such as pollution, water contamination, reduction in pollinators, the disappearance of species, toxicity to humans and animals, etc. More challenging is the development of fungal resistance against these chemical products.
Moreover, fungal diseases threaten biodiversity and agricultural practices. Fungal infection can decrease root and shoot lengths and root and shoots dry weight [88]. NPs could be a potential alternative to pesticides, allowing the control of fungal pathogens in a comparable or superior way to current chemical methods. Many studies have shown NP’s antimycotic activity against fungi that cause crop diseases [79,86,91,99,100,140,145,146,148,182]. The application of AgNPs in a greenhouse experiment using Triticum aestivum infected with Bipolaris sorokiniana increased root length, shoot length, and shoot dry weight by 11.55%, 30.67%, and 60.83%, respectively, when compared to untreated plants [88].
Interestingly, this study also found that applying AgNPs increased lignification, a known factor in plant development and disease resistance [183], to an intensity similar to healthy control. Infected and untreated plant lignification was barely visible through histochemical staining. SiO2 NPs used to treat Rhizoctonia solani were able to alleviate symptoms of infection in seedlings. The NPs exposure increased levels of photosynthetic pigments (chlorophylls and carotenoids), reduced lipid peroxidation, enhanced root and shoot lengths and fresh and dry weight of seedlings, increased salicylic acid levels (leads to the defense of lignification), induced phenolic and flavonoid antioxidant activity, and upregulated genes POD, SOD, APX, CAT, and PPO to manage oxidative stress. Therefore SiO2 NPs were able to prevent the harm of R. solani infection and help seedlings maintain homeostasis [91].

11.2. Dentistry

One of the applications of metallic NPs with antifungal activity is in oral diseases. One of the most studied fungi affecting the oral cavity is the opportunistic pathogen C. albicans. This yeast is relevant in different patients with specific treatments that weaken the immune response, e.g., leukemia, malnutrition, immunosuppression, radiation, and chemotherapy [184]. These patients have shown growth of C. albicans in denture prosthetics with a concomitant resistance against conventional antifungal agents.
One study showed the development of a polymeric agent consisting of a mixture of AgBr-NPs, quaternary ammonium salt polymer 4-vinyl pyridinium (NPVP), mixed with a denture resin such as polymethylmethacrylate (PMMA) showed antifungal activity against C. albicans. The AgBr-NPs used had an average size of 30 nm with a MIC of 250 µg/mL [185].

11.3. Biomedical

Using cotton fibers covered by metallic NPs is an attractive application for topical infections or burns. For example, anti-Candida activity was achieved when the fibers were covered with CuNPs. The antifungal activity remained even after washing or ironing for 6 months [186].
Another composite that showed antifungal activity against C. albicans was ZnONPs with a 30 nm diameter coated with chitosan-linoleic acid. This nanocomposite showed MICs of 32 mg/mL vs. 8 mg/mL measured for fluconazole [187].
Lastly, metallic prostheses fabricated with titanium are susceptible to being covered by the synthesis of C. albicans biofilms [188]. When titanium was coated with poly(dimethyl siloxane) (PDMS), a strong antifungal activity was measured with a decrease in the C. albicans biofilm biosynthesis [189].

11.4. Industrial

The food industry is another attractive field for developing metallic NPs for fungal control. For instance, aflatoxins are produced by filamentous fungi such as the genus Aspergillus. These compounds are toxins with known mutagenic and allergenic activities with severe clinical implications [190]. The use of poly(lactic acid) as a matrix covered with ZnO NPs showed an inhibitory effect against A. flavus and A. parasiticus [191].
Another potential application of ZnO NPs is preventing spore germination on air filters. In this way, the dissemination of fungi could be restricted. This is critical in food manufacturing facilities where the air is filtered for trapping microorganisms, dust, and other particles.
A summary of the potential applications of the NPS is presented in Figure 6.

12. Conclusions

There is a lack of knowledge pertaining to the mechanism of action when it comes to the antifungal activity of NPs. NP-mediated fungal cell toxicity occurs through many mechanisms, including cell wall damage and change (such as ergosterol synthesis inhibition), gene regulation, interaction with reproductive structures and hyphae, and ROS production, causing cascades of damages such as lipid peroxidation and mitochondrial damage. This multi-faceted attack makes NPs an avenue worth investigating as potential future antifungal treatments. Due to the multitude of factors involved in NPs formation, such as shape, size, and composition, combined with the variety of fungal strains that undergo testing, it is challenging to compare and contrast experiments and pinpoint what factors are more crucial than others. However, it is apparent in the literature that NPs exhibit a promising antifungal effect that can be enhanced or customized through formulation. These NPs could be helpful in a clinical setting, either in medical applications such as functionalized with known antifungals or on medical devices and surfaces.
Further studies are necessary to develop metallic NPs for different applications as their use so far is very limited. In the case of their applications as antifungal in packaging should consider more toxic studies because of their contact with meals, fruits, and vegetables. Additionally, the available techniques used to determine the level of metallic NPs are deficient when studying NPs that interact with food.
Also, the impact of metallic NPs on the environment should be considered, including their effects on aquatic life and the health risks associated with their use.

Author Contributions

Y.N.S. and H.B. conceptualized, wrote the original manuscript, and edited the final version. All authors have read and agreed to the published version of the manuscript.

Funding

This study was supported by the Antibody Engineering Facility, Immunity and Infection Research Centre, University of British Columbia, Vancouver, BC, Canada.

Institutional Review Board Statement

Not applicable for studies not involving humans or animals.

Informed Consent Statement

Not applicable.

Data Availability Statement

All data is available upon request.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Whittaker, R.H. New concepts of kingdoms of organisms: Evolutionary relations are better represented by new classifications than by the traditional two kingdoms. Science 1969, 163, 150–160. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  2. Cavalier-Smith, T. Kingdom Chromista and its eight phyla: A new synthesis emphasising periplastid protein targeting, cytoskeletal and periplastid evolution, and ancient divergences. Protoplasma 2018, 255, 297–357. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  3. Hawksworth, D.L.; Lücking, R. Fungal Diversity Revisited: 2.2 to 3.8 Million Species. Microbiol. Spectr. 2017, 5. [Google Scholar] [CrossRef] [PubMed]
  4. Blackwell, M. The Fungi: 1, 2, 3 … 5.1 million species? Am. J. Botany. 2011, 98, 426–438. [Google Scholar] [CrossRef] [PubMed]
  5. Taylor, D.L.; Hollingsworth, T.N.; McFarland, J.W.; Lennon, N.J.; Nusbaum, C.; Ruess, R.W. A first comprehensive census of fungi in soil reveals both hyperdiversity and fine-scale niche partitioning. Ecol. Monogr. 2014, 84, 3–20. [Google Scholar] [CrossRef]
  6. Bongomin, F.; Gago, S.; Oladele, R.; Denning, D. Global and multi-national prevalence of fungal diseases—Estimate precision. J. Fungi. 2017, 3, 57. [Google Scholar] [CrossRef]
  7. Van Rooij, P.; Martel, A.; Haesebrouck, F.; Pasmans, F. Amphibian chytridiomycosis: A review with focus on fungus-host interactions. Vet. Res. 2015, 46, 137. [Google Scholar] [CrossRef] [Green Version]
  8. Fisher, M.C.; Henk Daniel, A.; Briggs, C.J.; Brownstein, J.S.; Madoff, L.C.; McCraw, S.L. Emerging fungal threats to animal, plant and ecosystem health. Nature 2012, 484, 186–194. [Google Scholar] [CrossRef] [Green Version]
  9. Fisher, M.C.; Hawkins, N.J.; Sanglard, D.; Gurr, S.J. Worldwide emergence of resistance to antifungal drugs challenges human health and food security. Science 2018, 360, 739–742. [Google Scholar] [CrossRef] [Green Version]
  10. Robbins, N.; Caplan, T.; Cowen, L.E. Molecular evolution of antifungal drug resistance. Ann. Rev. Microbiol. 2017, 71, 753–775. [Google Scholar] [CrossRef]
  11. Cole, G.T. Chapter 73 Basic Biology of Fungi. In Medical Microbiology, 4th ed.; University of Texas Medical Branch at Galveston: Galveston, TX, USA, 1996. [Google Scholar]
  12. Wyatt, T.T.; Wösten, H.A.B.; Dijksterhuis, J. Fungal spores for dispersion in space and time. Adv. Appl. Microbiol. 2013, 85, 43–91. [Google Scholar]
  13. Wendland, J. Comparison of morphogenetic networks of filamentous fungi and yeast. Fungal Genet. Biol. 2001, 34, 63–82. [Google Scholar] [CrossRef]
  14. Free, S.J. Fungal cell wall organization and biosynthesis. Adv. Genet. 2013, 81, 33–82. [Google Scholar]
  15. Brown, J.A.; Catley, B.J. Monitoring polysaccharide synthesis in Candida albicans. Carbohyd. Res. 1992, 227, 195–202. [Google Scholar] [CrossRef]
  16. Bernard, M. Aspergillus fumigatus cell wall: Composition and biosynthesis. Med. Mycol. 2001, 39, 9–17. [Google Scholar] [CrossRef] [Green Version]
  17. Grun, C.H. The structure of cell wall-glucan from fission yeast. Glycobiology 2004, 15, 245–257. [Google Scholar] [CrossRef] [Green Version]
  18. Klis, F.M.; Groot, P.D.; Hellingwerf, K. Molecular organization of the cell wall of Candida albicans. Med. Mycol. 2001, 39 (Suppl. SI), 1–8. [Google Scholar] [CrossRef] [Green Version]
  19. Manners, D.J.; Masson, A.J.; Patterson, J.C. The structure of a β-(1→3)-D-glucan from yeast cell walls. Biochem. J. 1973, 135, 19–30. [Google Scholar] [CrossRef]
  20. Manners, D.J.; Masson, A.J.; Patterson, J.C.; Björndal, H.; Lindberg, B. The structure of a β-(1→6)-D-glucan from yeast cell walls. Biochem. J. 1973, 135, 31–36. [Google Scholar] [CrossRef]
  21. Fontaine, T.; Simenel, C.; Dubreucq, G.; Adam, O.; Delepierre, M.; Lemoine, J.; Vorgias, C.E.; Diaquin, M.; Latgé, J. Molecular organization of the alkali-insoluble fraction of Aspergillus fumigatus cell wall. J. Biol. Chem. 2000, 275, 27594–27607. [Google Scholar] [CrossRef] [Green Version]
  22. Douglas, C.M. Fungal β(1,3)-D-glucan synthesis. Med. Mycol. 2001, 39 (Suppl. I), 55–66. [Google Scholar] [CrossRef] [PubMed]
  23. Shematek, E.M.; Braatz, J.A.; Cabib, E. Biosynthesis of the yeast cell wall. I. Preparation and properties of β-(1→3)glucan synthetase. J. Biol. Chem. 1980, 255, 888–894. [Google Scholar] [CrossRef] [PubMed]
  24. Kollár, R.; Reinhold, B.B.; Petráková, E.; Yeh, H.J.; Ashwell, G.; Drgonová, J.; Kapteyn, J.C.; Klis, F.M.; Cabib, E. Architecture of the yeast cell wall. J. Biol. Chem. 1997, 272, 17762–17775. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  25. Cabib, E.; Bowers, B.; Sburlati, A.; Silverman, S.J. Fungal cell wall synthesis: The construction of a biological structure. Microbiol. Sci. 1988, 5, 370–375. [Google Scholar] [PubMed]
  26. Cabib, E.; Roberts, R.; Bowers, B. Synthesis of the yeast cell wall and its regulation. Ann. Rev. Biochem. 1982, 51, 763–793. [Google Scholar] [CrossRef]
  27. Latgé, J. The cell wall: A carbohydrate armour for the fungal cell. Mol. Microbiol. 2007, 66, 279–290. [Google Scholar] [CrossRef]
  28. Nakajima, T.; Yoshida, M.; Hiura, N.; Matsuda, K. Structure of the cell wall proteogalactomannan from Neurospora crassa. I. Purification of the proteoheteroglycan and characterization of alkali-labile oligosaccharides. J. Biol. Chem. 1984, 96, 1005–1011. [Google Scholar]
  29. Nakajima, T.; Yoshida, M.; Nakamura, M.; Hiura, N.; Matsuda, K. Structure of the cell wall proteogalactomannan from Neurospora crassa. II. Structural analysis of the polysaccharide part. J. Biol. Chem. 1984, 96, 1013–1020. [Google Scholar]
  30. Ragni, E.; Fontaine, T.; Gissi, C.; Latgè, J.P.; Popolo, L. The Gas family of proteins of Saccharomyces cerevisiae: Characterization and evolutionary analysis. Yeast. 2007, 24, 297–308. [Google Scholar] [CrossRef] [Green Version]
  31. Popolo, L.; Vai, M. The Gas1 glycoprotein, a putative wall polymer cross-linker. Biochim. Bioph. Acta (BBA) Gen. Subj. 1999, 1426, 385–400. [Google Scholar] [CrossRef]
  32. Mouyna, I.; Fontaine, T.; Vai, M.; Monod, M.; Fonzi, W.A.; Diaquin, M.; Popolo, L.; Hartland, R.P.; Latgé, J. Glycosylphosphatidylinositol-anchored glucanosyltransferases play an active role in the biosynthesis of the fungal cell wall. J. Biol. Chem. 2000, 275, 14882–14889. [Google Scholar] [CrossRef] [Green Version]
  33. Georgopapadakou, N. The fungal cell wall as a drug target. Trends Microbiol. 1995, 3, 98–104. [Google Scholar] [CrossRef]
  34. Klutts, J.S.; Yoneda, A.; Reilly, M.C.; Bose, I.; Doering, T.L. Glycosyltransferases and their products: Cryptococcal variations on fungal themes. FEMS Yeast Res. 2006, 6, 499–512. [Google Scholar] [CrossRef] [Green Version]
  35. Baladrón, V.; Ufano, S.; Dueñas, E.; Martín-Cuadrado, A.B.; del Rey, F.; Vázquez de Aldana, C.R. Eng1p, an endo-1,3-β-glucanase localized at the daughter side of the septum, is involved in cell separation in Saccharomyces cerevisiae. Eukaryot. Cell. 2002, 1, 774–786. [Google Scholar] [CrossRef]
  36. Martín-Cuadrado, A.B.; Dueñas, E.; Sipiczki, M.; de Aldana, C.R.V.; del Rey, F. The endo-β-1,3-glucanase eng1p is required for dissolution of the primary septum during cell separation in Schizosaccharomyces pombe. J. Cell Sci. 2003, 116, 1689–1698. [Google Scholar] [CrossRef] [Green Version]
  37. Roncero, C. The genetic complexity of chitin synthesis in fungi. Curr. Genet. 2002, 41, 367–378. [Google Scholar] [CrossRef]
  38. Silverman, S.J.; Sburlati, A.; Slater, M.L.; Cabib, E. Chitin synthase 2 is essential for septum formation and cell division in Saccharomyces cerevisiae. PNAS 1988, 85, 4735–4739. [Google Scholar] [CrossRef] [Green Version]
  39. Shaw, J.A.; Mol, P.C.; Bowers, B.; Silverman, S.J.; Valdivieso, M.H.; Durán, A.; Cabib, E. The function of chitin synthases 2 and 3 in the Saccharomyces cerevisiae cell cycle. J. Cell Biol. 1991, 114, 111–123. [Google Scholar] [CrossRef] [Green Version]
  40. Bulawa, C.E. CSD2, CSD3, and CSD4, genes required for chitin synthesis in Saccharomyces cerevisiae: The CSD2 gene product is related to chitin synthases and to developmentally regulated proteins in Rhizobium species and Xenopus laevis. Mol. Cell. Biol. 1992, 12, 1764–1773. [Google Scholar]
  41. Valdivieso, M.H.; Mol, P.C.; Shaw, J.A.; Cabib, E. CAL1, a gene required for activity of chitin synthase 3 in Saccharomyces cerevisiae. J. Cell Biol. 1991, 114, 101–109. [Google Scholar] [CrossRef] [Green Version]
  42. Bulawa, C.E. Genetics and molecular biology of chitin synthesis in fungi. Ann. Rev. Microbiol. 1993, 47, 505–534. [Google Scholar] [CrossRef] [PubMed]
  43. Mellado, E.; Aufauvre-Brown, A.; Gow, N.A.R.; Holden, D.W. The Aspergillus fumigatus chsC and chsG genes encode Class III chitin synthases with different functions. Mol. Microbiol. 1996, 20, 667–769. [Google Scholar] [CrossRef] [PubMed]
  44. Munro, C.A.; Gow, N.A.R. Chitin synthesis in human pathogenic fungi. Med. Mycol. 2001, 39, 41–53. [Google Scholar] [CrossRef] [PubMed]
  45. Yarden, O.; Yanofsky, C. Chitin synthase 1 plays a major role in cell wall biogenesis in Neurospora crassa. Genes Develop. 1991, 5, 2420–2430. [Google Scholar] [CrossRef] [PubMed]
  46. Din, A.B.; Specht, C.A.; Robbins, P.W.; Yarden, O. Chs-4, a class IV chitin synthase gene from Neurospora crassa. Molec. Gen. Genet. 1996, 250, 214–222. [Google Scholar]
  47. Borkovich, K.A.; Alex, L.A.; Yarden, O.; Freitag, M.; Turner, G.E.; Read, N.D.; Seiler, S.; Bell-Pedersen, D.; Paietta, J.; Plesofsky, N.; et al. Lessons from the genome sequence of Neurospora crassa: Tracing the path from genomic blueprint to multicellular organism. Microbiol. Mol. Biol. Rev. 2004, 68, 1–108. [Google Scholar] [CrossRef]
  48. Din, A.B.; Yarden, O. The Neurospora crassa chs-2 gene encodes a non-essential chitin synthase. Microbiology 1994, 140, 2189–2197. [Google Scholar] [CrossRef] [Green Version]
  49. Beth-Din, A.; Yarden, O. The Neurospora crassa chs3 gene encodes an essential class I chitin synthase. Mycologia 2000, 92, 65–73. [Google Scholar] [CrossRef]
  50. Kuranda, M.J.; Robbins, P.W. Chitinase is required for cell separation during growth of Saccharomyces cerevisiae. J. Biol. Chem. 1991, 266, 19758–19767. [Google Scholar] [CrossRef]
  51. Jaques, A.K.; Fukamizo, T.; Hall, D.; Barton, R.C.; Escott, G.M.; Parkinson, T.; Hitchcock, C.A.; Adams, D.J. Disruption of the gene encoding the ChiB1 chitinase of Aspergillus fumigatus and characterization of a recombinant gene product. Microbiology 2003, 149, 2931–2939. [Google Scholar] [CrossRef] [Green Version]
  52. Reichard, U.; Hung, C.-Y.; Thomas, P.W.; Cole, G.T. Disruption of the gene which encodes a serodiagnostic antigen and chitinase of the human fungal pathogen Coccidioides immitis. Infect. Immun. 2000, 68, 5830–5838. [Google Scholar] [CrossRef] [Green Version]
  53. Takaya, N.; Yamazaki, D.; Horiuchi, H.; Ohta, A.; Takagi, M. Cloning and characterization of a chitinase-encoding gene (chiA) from Aspergillus nidulans, disruption of which decreases germination frequency and hyphal growth. Biosci. Biotechnol. Biochem. 1998, 62, 60–65. [Google Scholar] [CrossRef] [Green Version]
  54. Lee, Y.; Puumala, E.; Robbins, N.; Cowen, L.E. Antifungal drug resistance: Molecular mechanisms in Candida albicans and beyond. Chem. Rev. 2021, 121, 3390–3411. [Google Scholar] [CrossRef]
  55. Kim, K.J.; Sung, W.S.; Suh, B.K.; Moon, S.K.; Choi, J.S.; Kim, J.G.; Lee, D.G. Antifungal activity and mode of action of silver nano-particles on Candida albicans. Biometals 2009, 22, 235–242. [Google Scholar] [CrossRef]
  56. Wang, D.; Xue, B.; Wang, L.; Zhang, Y.; Liu, L.; Zhou, Y. Fungus-mediated green synthesis of nano-silver using Aspergillus sydowii and its antifungal/antiproliferative activities. Sci. Rep. 2021, 11, 10356. [Google Scholar] [CrossRef]
  57. Monteiro, D.R.; Gorup, L.F.; Silva, S.; Negri, M.; De Camargo, E.R.; Oliveira, R.; Barbosa, D.D.B.; Henriques, M. Silver colloidal nanoparticles: Antifungal effect against adhered cells and biofilms of Candida albicans and Candida glabrata. Biofouling 2011, 27, 711–719. [Google Scholar] [CrossRef]
  58. Auyeung, A.; Casillas-Santana, M.A.; Martinez-Castanon, G.A.; Slavin, Y.N.; Zhao, W.; Asnis, J.; Häfeli, U.O.; Bach, H. Effective control of molds using a combination of nanoparticles. PLoS ONE 2017, 12, e0169940. [Google Scholar] [CrossRef] [Green Version]
  59. Panáček, A.; Kolář, M.; Večeřová, R.; Prucek, R.; Soukupová, J.; Kryštof, V.; Hamal, P.; Zbořil, R.; Kvítek, L. Antifungal activity of silver nanoparticles against Candida spp. Biomaterials 2009, 30, 6333–6340. [Google Scholar] [CrossRef]
  60. Xia, Z.K.; Ma, Q.H.; Li, S.Y.; Zhang, D.Q.; Cong, L.; Tian, Y.L.; Yang, R.Y. The antifungal effect of silver nanoparticles on Trichosporon asahii. J. Microbiol. Immunol. Infect. 2016, 49, 182–188. [Google Scholar] [CrossRef] [Green Version]
  61. Różalska, B.; Sadowska, B.; Budzyńska, A.; Bernat, P.; Różalska, S. Biogenic nanosilver synthesized in Metarhizium robertsii waste mycelium extract—As a modulator of Candida albicans morphogenesis, membrane lipidome and biofilm. PLoS ONE 2018, 13, e0194254. [Google Scholar] [CrossRef] [Green Version]
  62. Pereira, L.; Dias, N.; Carvalho, J.; Fernandes, S.; Santos, C.; Lima, N. Synthesis, characterization and antifungal activity of chemically and fungal-produced silver nanoparticles against Trichophyton rubrum. J. Appl. Microbiol. 2014, 117, 1601–1613. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  63. Chen, J.; Sun, L.; Cheng, Y.; Lu, Z.; Shao, K.; Li, T.; Hu, C.; Han, H. Graphene oxide-silver nanocomposite: Novel agricultural antifungal agent against Fusarium graminearum for crop disease prevention. ACS Appl. Mater. Interfaces 2016, 8, 24057–24070. [Google Scholar] [CrossRef] [PubMed]
  64. Martinez-Gutierrez, F.; Olive, P.L.; Banuelos, A.; Orrantia, E.; Nino, N.; Sanchez, E.M.; Ruiz, F.; Bach, H.; Av-Gay, Y. Synthesis, characterization, and evaluation of antimicrobial and cytotoxic effect of silver and titanium nanoparticles. Nanomedicine 2010, 6, 681–688. [Google Scholar] [CrossRef] [PubMed]
  65. Selvaraj, M.; Pandurangan, P.; Ramasami, N.; Rajendran, S.B.; Sangilimuthu, S.N.; Perumal, P. Highly potential antifungal activity of quantum-sized silver nanoparticles against Candida albicans. Appl. Biochem. Biotechnol. 2014, 173, 55–66. [Google Scholar] [CrossRef] [PubMed]
  66. Martínez, A.; Apip, C.; Meléndrez, M.F.; Domínguez, M.; Sánchez-Sanhueza, G.; Marzialetti, T.; Catalán, A. Dual antifungal activity against Candida albicans of copper metallic nanostructures and hierarchical copper oxide marigold-like nanostructures grown in situ in the culture medium. J. Appl. Microbiol. 2021, 130, 1883–1892. [Google Scholar] [CrossRef]
  67. Prucek, R.; Tuček, J.; Kilianová, M.; Panáček, A.; Kvítek, L.; Filip, J.; Kolář, M.; Tománková, K.; Zbořil, R. The targeted antibacterial and antifungal properties of magnetic nanocomposite of iron oxide and silver nanoparticles. Biomaterials 2011, 32, 4704–4713. [Google Scholar] [CrossRef]
  68. Zamperini, C.A.; André, R.S.; Longo, V.M.; Mima, E.G.; Vergani, C.E.; Machado, A.L.; Varela, J.A.; Longo, E. Antifungal applications of Ag-decorated hydroxyapatite nanoparticles. J. Nanomater. 2013, 2013, e174398. [Google Scholar] [CrossRef] [Green Version]
  69. De la Rosa-García, S.C.; Martínez-Torres, P.; Gómez-Cornelio, S.; Corral-Aguado, M.A.; Quintana, P.; Gómez-Ortíz, N.M. Antifungal activity of ZnO and MgO nanomaterials and their mixtures against Colletotrichum gloeosporioides strains from tropical fruit. J. Nanomater. 2018, 2018, e3498527. [Google Scholar] [CrossRef] [Green Version]
  70. Athie-García, M.S.; Piñón-Castillo, H.A.; Muñoz-Castellanos, L.N.; Ulloa-Ogaz, A.L.; Martínez-Varela, P.I.; Quintero-Ramos, A.; Duran, R.; Murillo-Ramirez, J.G.; Orrantia-Borunda, E. Cell wall damage and oxidative stress in Candida albicans ATCC 10231 and Aspergillus niger caused by palladium nanoparticles. Toxicol. Vitr. 2018, 48, 111–120. [Google Scholar] [CrossRef]
  71. Parsameher, N.; Rezaei, S.; Khodavasiy, S.; Salari, S.; Hadizade, S.; Kord, M.; Mousavi, S.A.A. Effect of biogenic selenium nanoparticles on ERG11 and CDR1 gene expression in both fluconazole-resistant and -susceptible Candida albicans isolates. Curr. Med. Mycol. 2017, 3, 16–20. [Google Scholar]
  72. Bafghi, M.H.; Nazari, R.; Darroudi, M.; Zargar, M.; Zarrinfar, H. The effect of biosynthesized selenium nanoparticles on the expression of CYP51A and HSP90 antifungal resistance genes in Aspergillus fumigatus and Aspergillus flavus. Biotechnol. Prog. 2022, 38, e3206. [Google Scholar] [CrossRef]
  73. Paulo, C.S.O.; Vidal, M.; Ferreira, L.S. Antifungal nanoparticles and surfaces. Biomacromolecules 2010, 11, 2810–2817. [Google Scholar] [CrossRef]
  74. Hosseini, S.S.; Ghaemi, E.; Noroozi, A.; Niknejad, F. Zinc oxide nanoparticles inhibition of initial adhesion and ALS1 and ALS3 gene expression in Candida albicans strains from urinary tract infections. Mycopathologia 2019, 184, 261–271. [Google Scholar] [CrossRef]
  75. Hosseini, S.S.; Joshaghani, H.; Shokohi, T.; Ahmadi, A.; Mehrbakhsh, Z. Antifungal activity of ZnO nanoparticles and nystatin and downregulation of SAP1-3 genes expression in fluconazole-resistant Candida albicans isolates from vulvovaginal candidiasis. Infect. Drug Resist. 2020, 13, 385–394. [Google Scholar] [CrossRef] [Green Version]
  76. Jebali, A.; Hajjar, F.H.E.; Pourdanesh, F.; Hekmatimoghaddam, S.; Kazemi, B.; Masoudi, A.; Daliri, K.; Sedighi, N. Silver and gold nanostructures: Antifungal property of different shapes of these nanostructures on Candida species. Med. Mycol. 2014, 52, 65–72. [Google Scholar]
  77. Ahmad, T.; Wani, I.A.; Lone, I.H.; Ganguly, A.; Manzoor, N.; Ahmad, A.; Ahmed, J.; Al-Shihri, A.S. Antifungal activity of gold nanoparticles prepared by solvothermal method. Mater. Res. Bull. 2013, 48, 12–20. [Google Scholar] [CrossRef]
  78. Ing, L.Y.; Zin, N.M.; Sarwar, A.; Katas, H. Antifungal activity of chitosan nanoparticles and correlation with their physical properties. Int. J. Biomat. 2012, 2012, 632698. [Google Scholar] [CrossRef] [Green Version]
  79. Bramhanwade, K.; Shende, S.; Bonde, S.; Gade, A.; Rai, M. Fungicidal activity of Cu nanoparticles against Fusarium causing crop diseases. Environ. Chem. Lett. 2016, 14, 229–235. [Google Scholar] [CrossRef]
  80. Osonga, F.J.; Kalra, S.; Miller, R.M.; Isika, D.; Sadik, O.A. Synthesis, characterization and antifungal activities of eco-friendly palladium nanoparticles. RSC Adv. 2020, 10, 5894–5904. [Google Scholar] [CrossRef]
  81. Abid, S.; Uzair, B.; Niazi, M.B.K.; Fasim, F.; Bano, S.A.; Jamil, N.; Batool, R.; Sajjad, S. Bursting the virulence traits of MDR strain of Candida albicans using sodium alginate-based microspheres containing nystatin-loaded MgO/CuO nanocomposites. Int. J. Nanomed. 2021, 16, 1157–1174. [Google Scholar] [CrossRef]
  82. Chougale, R.; Kasai, D.; Nayak, S.; Masti, S.; Nasalapure, A.; Raghu, A.V. Design of eco-friendly PVA/TiO2 based nanocomposites and their antifungal activity study. Green Mater. 2020, 8, 40–48. [Google Scholar] [CrossRef]
  83. Pulit, J.; Banach, M.; Szczygłowska, R.; Bryk, M. Nanosilver against fungi. Silver nanoparticles as an effective biocidal factor. Acta Biochim. Pol. 2013, 60, 795–798. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  84. Guerra, J.D.; Sandoval, G.; Avalos-Borja, M.; Pestryakov, A.; Garibo, D.; Susarrey-Arce, A.; Bogdanchikova, N. Selective antifungal activity of silver nanoparticles: A comparative study between Candida tropicalis and Saccharomyces boulardii. Colloids Interface Sci. Commun. 2020, 37, 100280. [Google Scholar] [CrossRef]
  85. Aguilar-Méndez, M.A.; San Martín-Martínez, E.; Ortega-Arroyo, L.; Cobián-Portillo, G.; Sánchez-Espíndola, E. Synthesis and characterization of silver nanoparticles: Effect on phytopathogen Colletotrichum gloesporioides. J. Nanopart. Res. 2011, 13, 2525–2532. [Google Scholar] [CrossRef]
  86. Arciniegas-Grijalba, P.A.; Patiño-Portela, M.C.; Mosquera-Sánchez, L.P.; Guerrero-Vargas, J.A.; Rodríguez-Páez, J.E. ZnO nanoparticles (ZnO-NPs) and their antifungal activity against coffee fungus Erythricium salmonicolor. Appl. Nanosci. 2017, 7, 225–241. [Google Scholar] [CrossRef] [Green Version]
  87. Kim, S.W.; Jung, J.H.; Lamsal, K.; Kim, Y.S.; Min, J.S.; Lee, Y.S. Antifungal effects of silver nanoparticles (AgNPs) against various plant pathogenic fungi. Mycobiology 2012, 40, 53–58. [Google Scholar] [CrossRef] [Green Version]
  88. Mishra, S.; Singh, B.R.; Singh, A.; Keswani, C.; Naqvi, A.H.; Singh, H.B. Biofabricated silver nanoparticles act as a strong fungicide against Bipolaris sorokiniana causing spot blotch disease in wheat. PLoS ONE 2014, 9, e97881. [Google Scholar] [CrossRef]
  89. Zhao, J.; Wang, L.; Xu, D.; Lu, Z. Involvement of ROS in nanosilver-caused suppression of aflatoxin production from Aspergillus flavus. RSC Adv. 2017, 7, 23021–23026. [Google Scholar] [CrossRef]
  90. Tiwari, A.K.; Gupta, M.K.; Pandey, G.; Tilak, R.; Narayan, R.J.; Pandey, P.C. Size and zeta potential clicked germination attenuation and anti-sporangiospores activity of PEI-functionalized silver nanoparticles against COVID-19 associated Mucorales (Rhizopus arrhizus). Nanomaterials 2022, 12, 2235. [Google Scholar] [CrossRef]
  91. Abdelrhim, A.; Mazrou, Y.; Nehela, Y.; Atallah, O.; El-Ashmony, R.; Dawood, M. Silicon dioxide nanoparticles induce innate immune responses and activate antioxidant machinery in wheat against Rhizoctonia solani. Plants 2021, 10, 2758. [Google Scholar] [CrossRef]
  92. Cioffi, N.; Torsi, L.; Ditaranto, N.; Tantillo, G.; Ghibelli, L.; Sabbatini, L.; Bleve-Zacheo, T.; D’Alessio, M.; Zambonin, P.G.; Traversa, E. Copper nanoparticle/polymer composites with antifungal and bacteriostatic properties. Chem. Mater. 2005, 17, 5255–5262. [Google Scholar] [CrossRef]
  93. Alexander, J.W. History of the medical use of silver. Surg. Infect. 2009, 10, 289–292. [Google Scholar] [CrossRef] [Green Version]
  94. Xiong, Y.; Brunson, M.; Huh, J.; Huang, A.; Coster, A.; Wendt, K.; Fay, J.; Qin, D. The role of surface chemistry on the toxicity of Ag nanoparticles. Small 2013, 9, 2628–2638. [Google Scholar] [CrossRef]
  95. Kim, S.W.; Kim, K.S.; Lamsal, K.; Kim, Y.J.; Kim, S.B.; Jung, M.Y.; Sim, S.J.; Kim, H.S.; Chang, S.J.; Kim, J.K.; et al. An in vitro study of the antifungal effect of silver nanoparticles on oak wilt pathogen Raffaelea sp. J. Microb. Microbiol. 2009, 19, 760–764. [Google Scholar]
  96. Ouda, S.M. Antifungal activity of silver and copper nanoparticles on two plant pathogens, Alternaria alternata and Botrytis cinerea. Res. J. Microbiol. 2014, 9, 34–42. [Google Scholar] [CrossRef] [Green Version]
  97. Sousa, C.A.; Soares, H.M.V.M.; Soares, E.V. Metal(loid) oxide (Al2O3, Mn3O4, SiO2 and SnO2) nanoparticles cause cytotoxicity in yeast via intracellular generation of reactive oxygen species. Appl. Microbiol. Biotechnol. 2019, 103, 6257–6269. [Google Scholar] [CrossRef] [Green Version]
  98. Eda, G.; Fanchini, G.; Chhowalla, M. Large-area ultrathin films of reduced graphene oxide as a transparent and flexible electronic material. Nat. Nanotech. 2008, 3, 270–274. [Google Scholar] [CrossRef]
  99. Mukherjee, K.; Acharya, K.; Biswas, A.; Jana, N.R. TiO2 nanoparticles co-doped with nitrogen and fluorine as visible-light-activated antifungal agents. ACS Appl. Nano Mater. 2020, 3, 2016–2025. [Google Scholar] [CrossRef]
  100. Boxi, S.S.; Mukherjee, K.; Paria, S. Ag doped hollow TiO2 nanoparticles as an effective green fungicide against Fusarium solani and Venturia inaequalis phytopathogens. Nanotechnology 2016, 27, 085103. [Google Scholar] [CrossRef]
  101. Wani, A.H.; Shah, M.A. A unique and profound effect of MgO and ZnO nanoparticles on some plant pathogenic fungi. J. Appl. Pharm. Sci. 2012, 2, 40–44. [Google Scholar]
  102. He, L.; Liu, Y.; Mustapha, A.; Lin, M. Antifungal activity of zinc oxide nanoparticles against Botrytis cinerea and Penicillium expansum. Microbiol. Res. 2011, 166, 207–215. [Google Scholar] [CrossRef] [PubMed]
  103. Navale, G.R.; Shinde, S.S. Antimicrobial activity of ZnO nanoparticles against pathogenic bacteria and fungi. JSM Nanotechnol. Nanomed. 2015, 3, 1033. [Google Scholar]
  104. Babele, P.K.; Thakre, P.K.; Kumawat, R.; Tomar, R.S. Zinc oxide nanoparticles induce toxicity by affecting cell wall integrity pathway, mitochondrial function and lipid homeostasis in Saccharomyces cerevisiae. Chemosphere 2018, 213, 65–75. [Google Scholar] [CrossRef] [PubMed]
  105. Kumari, M.; Giri, V.P.; Pandey, S.; Kumar, M.; Katiyar, R.; Nautiyal, C.S.; Mishra, A. An insight into the mechanism of antifungal activity of biogenic nanoparticles than their chemical counterparts. Pestic. Biochem. Physiol. 2019, 157, 45–52. [Google Scholar] [CrossRef] [PubMed]
  106. Kalagatur, N.K.; Nirmal Ghosh, O.S.; Sundararaj, N.; Mudili, V. Antifungal activity of chitosan nanoparticles encapsulated with Cymbopogon martinii essential oil on plant pathogenic fungi Fusarium graminearum. Front. Pharmacol. 2018, 9, 610. [Google Scholar] [CrossRef] [Green Version]
  107. Molnár, Z.; Bódai, V.; Szakacs, G.; Erdélyi, B.; Fogarassy, Z.; Sáfrán, G.; Varga, T.; Kónya, Z.; Tóth-Szeles, E.; Szűcs, R.; et al. Green synthesis of gold nanoparticles by thermophilic filamentous fungi. Sci. Rep. 2018, 8, 3943. [Google Scholar] [CrossRef] [Green Version]
  108. Ruiz-Romero, P.; Valdez-Salas, B.; González-Mendoza, D.; Mendez-Trujillo, V. Antifungal effects of silver phytonanoparticles from Yucca shilerifera against strawberry soil-borne pathogens: Fusarium solani and Macrophomina phaseolina. Mycobiology 2018, 46, 47–51. [Google Scholar] [CrossRef] [Green Version]
  109. Gajbhiye, M.; Kesharwani, J.; Ingle, A.; Gade, A.; Rai, M. Fungus-mediated synthesis of silver nanoparticles and their activity against pathogenic fungi in combination with fluconazole. Nanomedicine 2009, 5, 382–386. [Google Scholar] [CrossRef]
  110. Singh, M.; Kumar, M.; Kalaivani, R.; Manikandan, S.; Kumaraguru, A.K. Metallic silver nanoparticle: A therapeutic agent in combination with antifungal drug against human fungal pathogen. Bioprocess. Biosyst. Eng. 2013, 36, 407–415. [Google Scholar] [CrossRef]
  111. Li, X.; Xu, H.; Chen, Z.-S.; Chen, G. Biosynthesis of nanoparticles by microorganisms and their applications. J. Nanomater. 2011, 2011, 1–16. [Google Scholar] [CrossRef] [Green Version]
  112. He, S.; Guo, Z.; Zhang, Y.; Zhang, S.; Wang, J.; Gu, N. Biosynthesis of gold nanoparticles using the bacteria Rhodopseudomonas capsulata. Mater. Lett. 2007, 61, 3984–3987. [Google Scholar] [CrossRef]
  113. Durán, N.; Marcato, P.D.; Alves, O.L.; De Souza, G.I.; Esposito, E. Mechanistic aspects of biosynthesis of silver nanoparticles by several Fusarium oxysporum strains. J. Nanobiotechnol. 2005, 3, 8. [Google Scholar] [CrossRef] [Green Version]
  114. Chen, J.; Wu, L.; Lu, M.; Lu, S.; Li, Z.; Ding, W. Comparative study on the fungicidal activity of metallic MgO nanoparticles and macroscale MgO against soilborne fungal phytopathogens. Front. Microbiol. 2020, 11, 365. [Google Scholar] [CrossRef] [Green Version]
  115. Lee, B.; Lee, M.J.; Yun, S.J.; Kim, K.; Choi, I.-H.; Park, S. Silver nanoparticles induce reactive oxygen species-mediated cell cycle delay and synergistic cytotoxicity with 3-bromopyruvate in Candida albicans, but not in Saccharomyces cerevisiae. Int. J. Nanomed. 2019, 14, 801–816. [Google Scholar] [CrossRef] [Green Version]
  116. Westmeier, D.; Solouk-Saran, D.; Vallet, C.; Siemer, S.; Docter, D.; Götz, H.; Männ, L.; Hasenberg, A.; Hahlbrock, A.; Erler, K.; et al. Nanoparticle decoration impacts airborne fungal pathobiology. PNAS 2018, 115, 7087–7092. [Google Scholar] [CrossRef] [Green Version]
  117. Gulati, S.; Sachdeva, M.; Bhasin, K.K. Capping agents in nanoparticle synthesis: Surfactant and solvent system. AIP Conf. Proc. 2018, 1953, 030214. [Google Scholar]
  118. Ma, H.; Chen, S.; Zhao, S.; Liu, X.; Li, D. A study of corrosion behavior of copper in acidic solutions containing cetyltrimethylammonium bromide. J. Electrochem. Soc. 2001, 148, B482. [Google Scholar] [CrossRef]
  119. Zhang, Z.; Zhao, B.; Hu, L. PVP protective mechanism of ultrafine silver powder synthesized by chemical reduction processes. J. Solid State Chem. 1996, 121, 105–110. [Google Scholar] [CrossRef]
  120. Petica, A.; Gavriliu, S.; Lungu, M.; Buruntea, N.; Panzaru, C. Colloidal silver solutions with antimicrobial properties. Mater. Sci. Eng. B 2008, 152, 22–27. [Google Scholar] [CrossRef]
  121. Zhang, Y.; Yang, M.; Portney, N.G.; Cui, D.; Budak, G.; Ozbay, E.; Ozkan, M.; Ozkan, C.S. Zeta potential: A surface electrical characteristic to probe the interaction of nanoparticles with normal and cancer human breast epithelial cells. Biomed. Microdevices 2008, 10, 321–328. [Google Scholar] [CrossRef]
  122. Wang, H.X.; Douglas, L.M.; Aimanianda, V.; Latgé, J.-P.; Konopka, J.B. The Candida albicans Sur7 protein is needed for proper synthesis of the fibrillar component of the cell wall that confers strength. Eukaryot. Cell 2011, 10, 72–80. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  123. Doige, C.A.; Yu, X.; Sharom, F.J. The effects of lipids and detergents on ATPase-active P-glycoprotein. Biochim. Bioph. Acta (BBA) Biomembr. 1993, 1146, 65–72. [Google Scholar] [CrossRef]
  124. Lempers, V.J.; van den Heuvel, J.J.; Russel, F.G.; Aarnoutse, R.E.; Burger, D.M.; Brüggemann, R.J.; Koenderink, J.B. Inhibitory potential of antifungal drugs on ATP-binding cassette transporters P-Glycoprotein, MRP1 to MRP5, BCRP, and BSEP. Antimicrob. Agents Chemother. 2016, 60, 3372–3379. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  125. Gintjee, T.J.; Donnelley, M.A.; Thompson, G.R. Aspiring antifungals: Review of current antifungal pipeline developments. J. Fungi. 2020, 6, 28. [Google Scholar] [CrossRef] [Green Version]
  126. Dixon, D.M.; Walsh, T.J. Chapter 76 Antifungal agents. In Medical Microbiology, 4th ed.; University of Texas Medical Branch at Galveston: Galveston, TX, USA, 1996. [Google Scholar]
  127. Jallow, S.; Govender, N.P. Ibrexafungerp: A first-in-class oral triterpenoid glucan synthase inhibitor. J. Fungi. 2021, 7, 163. [Google Scholar] [CrossRef]
  128. Cournia, Z.; Ullmann, G.M.; Smith, J.C. Differential effects of cholesterol, ergosterol and lanosterol on a dipalmitoyl phosphatidylcholine membrane: A molecular dynamics simulation study. J. Phys, Chem. B 2007, 111, 1786–1801. [Google Scholar] [CrossRef]
  129. Vermitsky, J.-P.; Earhart, K.D.; Smith, W.L.; Homayouni, R.; Edlind, T.D.; Rogers, P.D. Pdr1 regulates multidrug resistance in Candida glabrata: Gene disruption and genome-wide expression studies. Mol. Microbiol. 2006, 61, 704–722. [Google Scholar] [CrossRef]
  130. Flowers, S.A.; Barker, K.S.; Berkow, E.L.; Toner, G.; Chadwick, S.G.; Gygax, S.E.; Morschhäuser, J.; Rogers, P.D. Gain-of-function mutations in UPC2 are a frequent cause of ERG11 upregulation in azole-resistant clinical isolates of Candida albicans. Eukaryot. Cell 2012, 11, 1289–1299. [Google Scholar] [CrossRef] [Green Version]
  131. Morschhäuser, J.; Barker, K.S.; Liu, T.T.; Blaß-Warmuth, J.; Homayouni, R.; Rogers, P.D. The transcription factor Mrr1p controls expression of the MDR1 efflux pump and mediates multidrug resistance in Candida albicans. PLoS Pathog. 2007, 3, e164. [Google Scholar] [CrossRef] [Green Version]
  132. Sagatova, A.A.; Keniya, M.V.; Wilson, R.K.; Monk, B.C.; Tyndall, J.D.A. Structural insights into binding of the antifungal drug fluconazole to Saccharomyces cerevisiae lanosterol 14α-demethylase. Antimicrob. Agents Chemother. 2015, 59, 4982–4989. [Google Scholar] [CrossRef] [Green Version]
  133. Chandra, J.; Kuhn, D.M.; Mukherjee, P.K.; Hoyer, L.L.; McCormick, T.; Ghannoum, M.A. Biofilm formation by the fungal pathogen Candida albicans: Development, architecture, and drug resistance. J. Bacteriol. 2001, 183, 5385–5394. [Google Scholar] [CrossRef] [Green Version]
  134. Marichal, P.; Koymans, L.; Willemsens, S.; Bellens, D.; Verhasselt, P.; Luyten, W.; Borgers, M.; Ramaekers, F.C.; Odds, F.C.; Bossche, H.V. Contribution of mutations in the cytochrome P450 14α-demethylase (Erg11p, Cyp51p) to azole resistance in Candida albicans. Microbiology 1999, 145, 2701–2713. [Google Scholar] [CrossRef]
  135. Liu, Y.; Cui, X.; Zhao, L.; Zhang, W.; Zhu, S.; Ma, J. Chitosan nanoparticles to enhance the inhibitory effect of natamycin on Candida albicans. J. Nanomater. 2021, 2021, 6644567. [Google Scholar] [CrossRef]
  136. Yang, M.; Du, K.; Hou, Y.; Xie, S.; Dong, Y.; Li, D.; Du, Y. Synergistic antifungal effect of amphotericin B-loaded poly(lactic-co-glycolic acid) nanoparticles and ultrasound against Candida albicans biofilms. Antimicrob. Agents Chemother. 2019, 63, e02022-18. [Google Scholar] [CrossRef] [Green Version]
  137. Lara, H.H.; Romero-Urbina, D.G.; Pierce, C.; Lopez-Ribot, J.L.; Arellano-Jiménez, M.J.; Jose-Yacaman, M. Effect of silver nanoparticles on Candida albicans biofilms: An ultrastructural study. J. Nanobiotechnol. 2015, 13, 91. [Google Scholar] [CrossRef] [Green Version]
  138. Vazquez-Muñoz, R.; Avalos-Borja, M.; Castro-Longoria, E. Ultrastructural analysis of Candida albicans when exposed to silver nanoparticles. PLoS ONE 2014, 9, e108876. [Google Scholar] [CrossRef]
  139. Alvarez-Peral, F.J.; Zaragoza, O.; Pedreno, Y.; Argüelles, J.-C. Protective role of trehalose during severe oxidative stress caused by hydrogen peroxide and the adaptive oxidative stress response in Candida albicans. Microbiology 2002, 148, 2599–2606. [Google Scholar] [CrossRef] [Green Version]
  140. Hwang, I.; Lee, J.; Hwang, J.H.; Kim, K.-J.; Lee, D.G. Silver nanoparticles induce apoptotic cell death in Candida albicans through the increase of hydroxyl radicals: Silver nanoparticles induce apoptotic cell death. FEBS J. 2012, 279, 1327–1338. [Google Scholar] [CrossRef]
  141. Pradhan, A.; Seena, S.; Schlosser, D.; Gerth, K.; Helm, S.; Dobritzsch, M.; Krauss, G.J.; Dobritzsch, D.; Pascoal, C.; Cássio, F. Fungi from metal-polluted streams may have high ability to cope with the oxidative stress induced by copper oxide nanoparticles: CuO-NPs induce oxidative stress in aquatic fungi. Environ. Toxicol. Chem. 2015, 34, 923–930. [Google Scholar] [CrossRef]
  142. Niazi, J.H.; Sang, B.-I.; Kim, Y.S.; Gu, M.B. Global gene response in Saccharomyces cerevisiae exposed to silver nanoparticles. Appl. Biochem. Biotechnol. 2011, 164, 1278–1291. [Google Scholar] [CrossRef]
  143. Cánovas, D.; Vooijs, R.; Schat, H.; de Lorenzo, V. The role of thiol species in the hypertolerance of Aspergillus sp. P37 to arsenic. J. Biol. Chem. 2004, 279, 51234–51240. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  144. Lushchak, V.I. Adaptive response to oxidative stress: Bacteria, fungi, plants and animals. Comp. Biochem. Physiol. Part C Toxicol. Pharmacol. 2011, 153, 175–190. [Google Scholar] [CrossRef]
  145. Jo, Y.-K.; Kim, B.H.; Jung, G. Antifungal activity of silver ions and nanoparticles on phytopathogenic fungi. Plant Dis. 2009, 93, 1037–1043. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  146. Akpinar, I.; Unal, M.; Sar, T. Potential antifungal effects of silver nanoparticles (AgNPs) of different sizes against phytopathogenic Fusarium oxysporum f. sp. radicis-lycopersici (FORL) strains. SN Appl. Sci. 2021, 3, 506. [Google Scholar] [CrossRef]
  147. Kim, K.-J.; Sung, W.S.; Moon, S.-K.; Choi, J.-S.; Kim, J.G.; Lee, D.G. Antifungal effect of silver nanoparticles on dermatophytes. J. Microbiol. Biotechnol. 2008, 18, 1482–1484. [Google Scholar] [PubMed]
  148. Rispail, N.; De Matteis, L.; Santos, R.; Miguel, A.S.; Custardoy, L.; Testillano, P.S.; Risuenño, M.C.; Peérez-de-Luque, A.; Maycock, C.; Fevereiro, P.; et al. Quantum dot and superparamagnetic nanoparticle interaction with pathogenic fungi: Internalization and toxicity profile. ACS Appl. Mater. Interfaces 2014, 6, 9100–9110. [Google Scholar] [CrossRef] [Green Version]
  149. Kasprowicz, M.J.; Kozioł, M.; Gorczyca, A. The effect of silver nanoparticles on phytopathogenic spores of Fusarium culmorum. Can. J. Microbiol. 2010, 56, 247–253. [Google Scholar] [CrossRef]
  150. Ogar, A.; Tylko, G.; Turnau, K. Antifungal properties of silver nanoparticles against indoor mould growth. Sci. Total Environm. 2015, 521–522, 305–314. [Google Scholar] [CrossRef]
  151. Ayala, A.; Muñoz, M.F.; Argüelles, S. Lipid peroxidation: Production, metabolism, and signaling mechanisms of malondialdehyde and 4-hydroxy-2-nonenal. Oxid. Med. Cell. Longe. 2014, 2014, 360438. [Google Scholar] [CrossRef] [Green Version]
  152. Bayry, J.; Aimanianda, V.; Guijarro, J.I.; Sunde, M.; Latgé, J.-P. Hydrophobins—Unique fungal proteins. PLoS Pathog. 2012, 8, e1002700. [Google Scholar] [CrossRef] [Green Version]
  153. Gómez, B.L.; Nosanchuk, J.D. Melanin and fungi. Curr. Opin. Infect. Dis. 2003, 16, 91–96. [Google Scholar] [CrossRef]
  154. Ramage, G.; Mowat, E.; Jones, B.; Williams, C.; Lopez-Ribot, J. Our current understanding of fungal biofilms. Crit. Rev. Microbiol. 2009, 35, 340–355. [Google Scholar] [CrossRef]
  155. Aslani, P.; Roudbar, M.S.; Roudbary, M. Novel formulated zinc oxide nanoparticles reduce Hwp1 gene expression involved in biofilm formation in Candida albicans with minimum cytotoxicity effect on human cells. Jundishapur J. Microbiol. 2018, 11, e79562. [Google Scholar] [CrossRef] [Green Version]
  156. Harrison, J.J.; Rabiei, M.; Turner, R.J.; Badry, E.A.; Sproule, K.M.; Ceri, H. Metal resistance in Candida biofilms. FEMS Microbiol. Ecol. 2006, 55, 479–491. [Google Scholar] [CrossRef]
  157. Monteiro, D.R.; Takamiya, A.S.; Feresin, L.P.; Gorup, L.F.; de Camargo, E.R.; Delbem, A.C.B.; Henriques, M.; Barbosa, D.B. Susceptibility of Candida albicans and Candida glabrata biofilms to silver nanoparticles in intermediate and mature development phases. J. Prosthodont. Res. 2015, 59, 42–48. [Google Scholar] [CrossRef] [Green Version]
  158. Das, K.; Roychoudhury, A. Reactive oxygen species (ROS) and response of antioxidants as ROS-scavengers during environmental stress in plants. Front. Environ. Sci. 2014, 2, 53. [Google Scholar] [CrossRef] [Green Version]
  159. Lee, Y.H.; Cheng, F.Y.; Chiu, H.W.; Tsai, J.C.; Fang, C.Y.; Chen, C.W.; Wang, Y.J. Cytotoxicity, oxidative stress, apoptosis and the autophagic effects of silver nanoparticles in mouse embryonic fibroblasts. Biomaterials 2014, 35, 4706–4715. [Google Scholar] [CrossRef]
  160. Pócsi, I.; Prade, R.A.; Penninckx, M.J. Glutathione, altruistic metabolite in fungi. Adv. Microbial. Physiol. 2004, 49, 1–76. [Google Scholar]
  161. Wade, A.M.; Tucker, H.N. Antioxidant characteristics of L-histidine. J. Nutrit. Biochem. 1998, 9, 308–315. [Google Scholar] [CrossRef]
  162. Lipovsky, A.; Nitzan, Y.; Gedanken, A.; Lubart, R. Antifungal activity of ZnO nanoparticles—The role of ROS mediated cell injury. Nanotechnology 2011, 2, 105101. [Google Scholar] [CrossRef]
  163. Waldeck, A.R.; Stocker, R. Radical-initiated lipid peroxidation in low density lipoproteins: Insights obtained from kinetic modeling. Chem. Res. Toxicol. 1996, 9, 954–964. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  164. Seong, M.; Lee, D.G. Reactive oxygen species-independent apoptotic pathway by gold nanoparticles in Candida albicans. Microbiol. Res. 2018, 207, 33–40. [Google Scholar] [CrossRef] [PubMed]
  165. Horstmann, C.; Campbell, C.; Kim, D.S.; Kim, K. Transcriptome profile with 20 nm silver nanoparticles in yeast. FEMS Yeast Res. 2019, 19, foz003. [Google Scholar] [CrossRef] [PubMed]
  166. Jones, A.M.; Garg, S.; He, D.; Pham, A.N.; Waite, T.D. Superoxide-mediated formation and charging of silver nanoparticles. Environ. Sci. Technol. 2011, 45, 1428–1434. [Google Scholar] [CrossRef] [PubMed]
  167. Toth, A.; Aufschnaiter, A.; Fedotovskaya, O.; Dawitz, H.; Ädelroth, P.; Büttner, S.; Ott, M. Membrane-tethering of cytochrome c accelerates regulated cell death in yeast. Cell Death Dis. 2020, 11, 722. [Google Scholar] [CrossRef] [PubMed]
  168. Mehra, R.K.; Winge, D.R. Metal ion resistance in fungi: Molecular mechanisms and their regulated expression. J. Cell Biochem. 1991, 45, 30–40. [Google Scholar] [CrossRef] [PubMed]
  169. Hohmann, S. Control of high osmolarity signalling in the yeast Saccharomyces cerevisiae. FEBS Lett. 2009, 583, 4025–4029. [Google Scholar] [CrossRef] [Green Version]
  170. García, R.; Rodríguez-Peña, J.M.; Bermejo, C.; Nombela, C.; Arroyo, J. The high osmotic response and cell wall integrity pathways cooperate to regulate transcriptional responses to zymolyase-induced cell wall stress in Saccharomyces cerevisiae. J. Biol. Chem. 2009, 284, 10901–10911. [Google Scholar] [CrossRef] [Green Version]
  171. Back, S.H.; Schröder, M.; Lee, K.; Zhang, K.; Kaufman, R.J. ER stress signaling by regulated splicing: IRE1/HAC1/XBP. Methods 2005, 35, 395–416. [Google Scholar] [CrossRef]
  172. Krysan, D. The cell wall and endoplasmic reticulum stress responses are coordinately regulated in Saccharomyces cerevisiae. Comm. Integr. Biol. 2009, 2, 233–235. [Google Scholar] [CrossRef]
  173. Sato, I.; Shimatani, K.; Fujita, K.; Abe, T.; Shimizu, M.; Fujii, T.; Hoshino, T.; Takaya, N. Glutathione reductase/glutathione is responsible for cytotoxic elemental sulfur tolerance via polysulfide shuttle in fungi. J. Biol. Chem. 2011, 286, 20283–20291. [Google Scholar] [CrossRef] [Green Version]
  174. Guimarães-Soares, L.; Felícia, H.; João Bebianno, M.; Cássio, F. Metal-binding proteins and peptides in the aquatic fungi Fontanospora fusiramosa and Flagellospora curta exposed to severe metal stress. Sci. Total Environ. 2006, 372, 148–156. [Google Scholar] [CrossRef]
  175. White, T.C.; Holleman, S.; Dy, F.; Mirels, L.F.; Stevens, D.A. Resistance mechanisms in clinical isolates of Candida albicans. Antimicrob. Agents Chemother. 2002, 46, 1704–1713. [Google Scholar] [CrossRef]
  176. Wang, D.Y.; Gricourt, M.; Arné, P.; Thierry, S.; Seguin, D.; Chermette, R.; Huang, W.Y.; Dannaoui, E.; Botterel, F.; Guillot, J. Mutations in the Cyp51A gene and susceptibility to itraconazole in Aspergillus fumigatus isolated from avian farms in France and China. Poult. Sci. 2014, 93, 12–15. [Google Scholar] [CrossRef]
  177. Gong, Y.; Li, T.; Yu, C.; Sun, S. Candida albicans heat shock proteins and Hsps-associated signaling pathways as potential antifungal targets. Front. Cell. Infect. Microbiol. 2017, 7, 520. [Google Scholar] [CrossRef] [Green Version]
  178. Nailis, H.; Kucharíková, S.; Řičicová, M.; Van Dijck, P.; Deforce, D.; Nelis, H.; Coenye, T. Real-time PCR expression profiling of genes encoding potential virulence factors in Candida albicans biofilms: Identification of model-dependent and -independent gene expression. BMC Microbiol. 2010, 10, 114. [Google Scholar] [CrossRef] [Green Version]
  179. Mayer, F.L.; Wilson, D.; Hube, B. Candida albicans pathogenicity mechanisms. Virulence 2013, 4, 119–128. [Google Scholar] [CrossRef] [Green Version]
  180. Naglik, J.R.; Challacombe, S.J.; Hube, B. Candida albicans secreted aspartyl proteinases in virulence and pathogenesis. Microbiol. Mol. Biol. Rev. 2003, 67, 400–428. [Google Scholar] [CrossRef] [Green Version]
  181. Lima, S.L.; Colombo, A.L.; de Almeida Junior, J.N. Fungal cell wall: Emerging antifungals and drug resistance. Front. Microbiol. 2019, 10, 2573. [Google Scholar] [CrossRef] [Green Version]
  182. Mahdizadeh, V.; Safaie, N.; Khelghatibana, F. Evaluation of antifungal activity of silver nanoparticles against some phytopathogenic fungi and Trichoderma harzianum. J. Crop Prot. 2015, 4, 291–300. [Google Scholar]
  183. Mandal, S.; Mitra, A. Reinforcement of cell wall in roots of Lycopersicon esculentum through induction of phenolic compounds and lignin by elicitors. Physiol. Molec. Plant Pathol. 2007, 71, 201–209. [Google Scholar] [CrossRef]
  184. Millsop, J.W.; Fazel, N. Oral candidiasis. Clin. Dermatol. 2016, 34, 487–494. [Google Scholar] [CrossRef] [PubMed]
  185. Zhang, Y.U.; Chen, Y.; Huang, L.; Chai, Z.; Shen, L.; Xiao, Y. The antifungal effects and mechanical properties of silver bromide/cationic polymer nano-composite-modified poly-methyl 2 methacrylate-based dental resin. Sci. Rep. 2017, 7, 1547. [Google Scholar] [CrossRef] [PubMed]
  186. Eremenko, A.M.; Petrik, I.S.; Smirnova, N.P.; Rudenko, A.V.; Marikvas, Y.S. Antibacterial and antimycotic activity of cotton fabrics, impregnated with silver and binary silver/copper nanoparticles. Nanoscale Res. Lett. 2016, 11, 28. [Google Scholar] [CrossRef] [PubMed]
  187. Barad, S.; Roudbary, M.; Nasrollahi Omran, A.; Porgham Daryasari, M. Preparation and characterization of ZnO nanoparticles coated by chitosan-linoleic acid; fungal growth and biofilm assay. Bratisl. Med. J. 2017, 118, 169–174. [Google Scholar] [CrossRef]
  188. Dhir, S. Biofilm and dental implant: The microbial link. J. Indian Soc. Periodontol. 2013, 17, 5–11. [Google Scholar] [CrossRef]
  189. Groza, A.; Ciobanu, C.S.; Popa, C.L.; Iconaru, S.L.; Chapon, P.; Luculescu, C.; Ganciu, M.; Predoi, D. Structural properties and antifungal activity against Candida albicans biofilm of different composite layers based on Ag/Zn doped hydroxyapatite-polydimethylsiloxanes. Polymers 2016, 8, 131. [Google Scholar] [CrossRef] [Green Version]
  190. Sarma, U.P.; Bhetaria, P.J.; Devi, P.; Varma, A. Aflatoxins: Implications on health. Indian J. Clin. Biochem. 2017, 32, 124–133. [Google Scholar] [CrossRef]
  191. Nasab, M.S.; Tabari, M.; Bidarigh, S. Antifungal activity of nano-composite films-based poly lactic acid. Nanomed. Res. J. 2019, 4, 186–192. [Google Scholar]
Figure 1. Fungal cell wall organization and membrane.
Figure 1. Fungal cell wall organization and membrane.
Nanomaterials 12 04470 g001
Figure 2. Glucan structure shows the (A) β-1,6 linkages and (B) β-1,3 linkages.
Figure 2. Glucan structure shows the (A) β-1,6 linkages and (B) β-1,3 linkages.
Nanomaterials 12 04470 g002
Figure 3. Chitin structure of polymerized N-acetylglucosamine monomers with (A) β-1,4 linkages.
Figure 3. Chitin structure of polymerized N-acetylglucosamine monomers with (A) β-1,4 linkages.
Nanomaterials 12 04470 g003
Figure 4. Different types of metallic NPs developed against fungi.
Figure 4. Different types of metallic NPs developed against fungi.
Nanomaterials 12 04470 g004
Figure 5. NPs mechanisms at the cellular level that lead to fungal cell damage include (A) ROS-inducing lipid peroxidation, (B) adsorption embedment and breakage of cell wall and membrane, (C) pit and pore formation, (D) leakage, releasing DNA and organelles from the cell, (E) ion release, (F) DNA intercalation, causing condensation and fragmentation, (G) gene expression changes, (H) ROS generation, (I) Mitochondrial release of cytochrome C into the cytosol, increasing metacaspase levels, leading to apoptosis cascade, (J) ribosome depolymerization, and (K) adsorption onto EPS, inhibiting biofilm formation.
Figure 5. NPs mechanisms at the cellular level that lead to fungal cell damage include (A) ROS-inducing lipid peroxidation, (B) adsorption embedment and breakage of cell wall and membrane, (C) pit and pore formation, (D) leakage, releasing DNA and organelles from the cell, (E) ion release, (F) DNA intercalation, causing condensation and fragmentation, (G) gene expression changes, (H) ROS generation, (I) Mitochondrial release of cytochrome C into the cytosol, increasing metacaspase levels, leading to apoptosis cascade, (J) ribosome depolymerization, and (K) adsorption onto EPS, inhibiting biofilm formation.
Nanomaterials 12 04470 g005
Figure 6. Potential application of NPs with antifungal activity to different industries.
Figure 6. Potential application of NPs with antifungal activity to different industries.
Nanomaterials 12 04470 g006
Table 1. Activities of NPs tested on different fungal species expressed in MIC100 (µg/mL).
Table 1. Activities of NPs tested on different fungal species expressed in MIC100 (µg/mL).
Type of NPShapeSize (nm)Organism(s) TestedMIC100 (µg/mL) Reference
AgSpherical3Saccharomyces cerevisiae (KCTC 7296)
Trichosporon beigelii (KCTC 7707)
Candida albicans (ATCC 90028)
2 [55]
Spherical1–21Candida albicans
Candida glabrata
Candida parapsilosis
Candida tropicalis
Fusarium solani
Fusarium moniliforme
Fusarium oxysporum
Aspergillus flavus
Aspergillus fumigatus
Aspergillus terrus
Sporothrix schenckii
Cryptococcus neoformans
0.25
0.125
0.25
0.125
1
2
4
1
2
2
0.25
0.25
[56]
Spherical5Candida albicans (324LA/84)
Candida glabrata (ATCC 90030)
Candida albicans (ATCC 10321)
Candida glabrata (D1)
0.4–0.8
0.4–0.8
0.8–1.6
1.6–3.3
[57]
Spherical7Aspergillus flavus
Aspergillus fumigatus
100
100
[58]
Spherical25Candida albicans (I)
Candida albicans (II)
Candida parapsilosis
Candida tropicalis
0.42
0.21
1.690
0.840
[59]
Spherical5–20Trichosporon asahii (CBS2479)
Trichosporon asahii (BZ701)
Trichosporon asahii (BZ702)
Trichosporon asahii (BZ703)
Trichosporon asahii (BZ704)
Trichosporon asahii (BZ705)
Trichosporon asahii (BZ705R)
Trichosporon asahii (BZ901)
Trichosporon asahii (BZ902)
Trichosporon asahii (BZ121)
Trichosporon asahii (BZ122)
Trichosporon asahii (BZ123)
Trichosporon asahii (BZ124)
Trichosporon asahii (BZ125)
Trichosporon asahii (CBS8904)
Trichosporon asahii (CBS7137)
Trichosporon asahii (CBS8520)
0.50
0.67
0.50
1.00
0.67
0.50
0.50
0.67
1.00
0.83
0.67
0.50
0.67
0.83
0.50
0.67
0.67
[60]
Spherical15–25Candida albicans (ATCC 10231)
Candida albicans (ATCC 90028)
Candida glabrata (ATCC 90030)
Candida parapsilosis (ATCC 22019)
1.56
6.25
3.12
6.25
[61]
Spherical25Candida albicans (I)
Candida albicans (II)
Candida parapsilosis
Candida tropicalis
27
27
27
27
[59]
Spherical to polyhedric73.72
Trichophyton rubrum (n = 8)
Trichophyton rubrum (ATCC MYA 4438)
0.5–2.5
<0.25
[62]
Spherical76.14Trichophyton rubrum (n = 8)
Trichophyton rubrum (ATCC MYA 4438)
>7.5
>7.5
[62]
Spherical100.6Trichophyton rubrum (n = 8)
Trichophyton rubrum (ATCC MYA 4438)
0.5–5
0.5
[62]
NRNRFusarium graminearum4.68 [63]
NRNRCandida albicans (I)
Candida albicans (II)
Candida parapsilosis
Candida tropicalis
0.42
0.21
1.69
0.84
[59]
NRNRCandida albicans (I)
Candida albicans (II)
Candida parapsilosis
Candida tropicalis
0.052
0.1
0.84
0.42
[59]
NRNRCandida albicans (I)
Candida albicans (II)
Candida parapsilosis
Candida tropicalis
3.38
3.38
3.38
3.38
[59]
NR20–25Aspergillus niger
Candida albicans
Cryptococcus neoformans
25
6
3
[64]
NRNRFusarium graminearum12.5[63]
Ag/ZnO Spherical7/477Aspergillus flavus
Aspergillus fumigatus
50/10
50/10
[58]
qAgSpherical2–3Candida albicans0.07[65]
CuSpherical10–40Candida albicans (ATCC 10231)
Candida albicans (Clinical strain C)
Candida albicans (Clinical strain E)
129.7
1037.5
518.8
[66]
Wires20–30 µm, 30–60 nm diameterCandida albicans (ATCC 10231)
Candida albicans (Clinical strain C)
Candida albicans (Clinical strain E)
260.3
260.3
260.3
[66]
γ-Fe2O3/AgNR20–40 (Ag) + 5 (γ-Fe2O3)Candida albicans (I)
Candida albicans (II)
Candida tropicalis (5)
Candida parapsilosis (6)
1.9
1.9
31.3
31.3
[67]
Fe3O4/AgNR~5 (Ag) + ~70 (Fe3O4)Candida albicans (I)
Candida albicans (II)
Candida tropicalis (5)
Candida parapsilosis (6)
1.9
1.9
3.9
7.8
[67]
GO/AgSpherical10–35Fusarium graminearum9.37[63]
HA/AgRod/Spherical12–27Candida albicans62.5[68]
MgO
500 °C calcination
Flaked layers52 ± 18Colletotrichum gloeosporioides (from papaya)
Colletotrichum gloeosporioides (from avocado)
156
312
[69]
MgO
1000 °C calcination
Flaked layers96 ± 33Colletotrichum gloeosporioides (from papaya)
Colletotrichum gloeosporioides (from avocado)
312
312
[69]
PdSpherical9 ± 3.9Candida albicans (ATCC 10231)
Aspergillus niger
212.5
200
[70]
SeSpherical80–200Candida albicans (ATCC 76615)
Candida albicans (ATCC 10231)
100
70
[71]
SeSpherical37–46Aspergillus fumigatus (TIMML-025)
Aspergillus fumigatus (TIMML-050)
Aspergillus fumigatus (TIMML-379)
0.5
0.5
1
[72]
Silica/AmBNR5Candida albicans
Candida krusei
Candida parapsilosis
Candida glabrata
Candida tropicalis
100
1000
1000–2000
300
100
[73]
TiO2/Ag + NaHB4NR250–300Aspergillus niger
Candida albicans
Cryptococcus neoformans
25
12.5
12.5
[64]
TiO2/Ag + UVNR250–300Aspergillus niger
Candida albicans
Cryptococcus neoformans
12.5
6
3
[64]
ZnO Spherical20–40Candida albicans (n = 125)0.2–296[74]
Spherical477Aspergillus flavus
Aspergillus fumigatus
20
20
[58]
ZnO
500 °C calcination
Spherical + cylindrical51 ± 13Colletotrichum gloeosporioides (from papaya)
Colletotrichum gloeosporioides (from avocado)
156
312
[69]
ZnO
1000 °C calcination
Hexagonal bars53 ± 17Colletotrichum gloeosporioides (from papaya)
Colletotrichum gloeosporioides (from avocado)
156
312
[69]
ZnO/Mg(OH)2
25 °C synthesis
Flakes + bars54 ± 17Colletotrichum gloeosporioides (from papaya)
Colletotrichum gloeosporioides (from avocado)
156
312
[69]
ZnO NP
25 °C synthesis
Hexagonal bars + ovoidal63 ± 18Colletotrichum gloeosporioides (from papaya)
Colletotrichum gloeosporioides (from avocado)
312
312
[69]
ZnO/Mg(OH)2
70 °C synthesis
Flakes + bars71 ± 22Colletotrichum gloeosporioides (from papaya)
Colletotrichum gloeosporioides (from avocado)
312
312
[69]
ZnO
25 °C synthesis, hydrothermal
Prisms with pyramidal ends77 ± 31Colletotrichum gloeosporioides (from papaya)
Colletotrichum gloeosporioides (from avocado)
312
312
[69]
ZnO/Mg(OH)2 NP
25 °C synthesis, hydrothermal 160 °C
Flakes + bars88 ± 30Colletotrichum gloeosporioides (from papaya)
Colletotrichum gloeosporioides (from avocado)
312
312
[69]
ZnO/Mg(OH)2
70 °C synthesis, hydrothermal 160 °C
Flakes + bars98 ± 41Colletotrichum gloeosporioides (from papaya)
Colletotrichum gloeosporioides (from avocado)
312
312
[69]
ZnO/MgO
25 °C synthesis, 500 °C calcination
Flakes139 ± 49Colletotrichum gloeosporioides (from papaya)
Colletotrichum gloeosporioides (from avocado)
312

312
[69]
ZnO/MgO
70 °C synthesis, 500 °C calcination
Flakes161 ± 44Colletotrichum gloeosporioides (from papaya)
Colletotrichum gloeosporioides (from avocado)
312

312
[69]
ZnO/MgO
70 °C synthesis, 1000 °C calcination
Flakes219 ± 39Colletotrichum gloeosporioides (from papaya)
Colletotrichum gloeosporioides (from avocado)
625
625
[69]
ZnONRNRCandida albicans (n = 10)0.02–269[75]
MIC, minimal inhibitory concentration; qAg, quantum silver; NR, Not reported; HA, hydroxyapatite.
Table 2. Activities of NPs tested on different fungal species expressed as MIC50 (µg/mL).
Table 2. Activities of NPs tested on different fungal species expressed as MIC50 (µg/mL).
Type of NPShapeSize (nm)Organism(s) TestedMIC50 (µg/mL) Reference
AgSpherical30–50Candida albicans (n = 30)
Candida glabrata (n = 30)
Candida tropicalis (n = 30)
4
9
9
[76]
AgCubical40–50Candida albicans (n = 30)
Candida glabrata (n = 30)
Candida tropicalis (n = 30)
1
7
7
[76]
AgWires250–300Candida albicans (n = 30)
Candida glabrata (n = 30)
Candida tropicalis (n = 30)
5
12
11
[76]
AuCubical30–50Candida albicans (n = 30)
Candida glabrata (n = 30)
Candida tropicalis (n = 30)
3
10
11
[76]
AuSpherical35–50Candida albicans (n = 30)
Candida glabrata (n = 30)
Candida tropicalis (n = 30)
8
13
12
[76]
AuWires300–500Candida albicans (n = 30)
Candida glabrata (n = 30)
Candida tropicalis (n = 30)
7
15
15
[76]
ZnO Spherical20–40Candida albicans (n = 125)8.2[74]
ZnONRNRCandida albicans (n = 10)5[75]
MIC, minimal inhibitory concentration; NR, Not reported.
Table 3. Activities of NPs tested on different fungal species expressed as MIC80 (µg/mL).
Table 3. Activities of NPs tested on different fungal species expressed as MIC80 (µg/mL).
Type of NPShapeSize (nm)Organism(s) TestedMIC80 (µg/mL) Reference
AuNP + SnCl2 as reducing agentPolygonal, almost spherical5–50Candida albicans
Candida tropicalis
Candida glabrata
16
16
16
[77]
AuNP + NaBH4 as reducing agentSpherical3–20Candida albicans
Candida tropicalis
Candida glabrata
4
4
4
[77]
MIC, minimal inhibitory concentration.
Table 4. Activities of NPs tested on different fungal species expressed as MIC90 (µg/mL).
Table 4. Activities of NPs tested on different fungal species expressed as MIC90 (µg/mL).
Type of NPShapeSize (nm)Organism(s) TestedMIC90 (µg/mL) Reference
LMW Chitosan NP ± 1 mg/mL chitosan Spherical174 ± 38.47 Candida albicans
Fusarium solani
Aspergillus niger
250
1000
NR
[78]
HMW Chitosan NP ± 1 mg/mL chitosan Spherical210 ± 24.54 Candida albicans
Fusarium solani
Aspergillus niger
1000
500
NR
[78]
LMW Chitosan NP ± 2 mg/mL chitosan Spherical233 ± 41.38Candida albicans
Fusarium solani
Aspergillus niger
857.2
857.2
NR
[78]
HMW Chitosan NP ± 2 mg/mL chitosan Spherical263 ± 86.44 Candida albicans
Fusarium solani
Aspergillus niger
857.2
857.2
1714
[78]
LMW Chitosan NP ± 3 mg/mL chitosan Spherical255 ± 42.81Candida albicans
Fusarium solani
Aspergillus niger
607.2
1214
NR
[78]
HMW Chitosan NP ± 3 mg/mL chitosan Spherical301 ± 72.85 Candida albicans
Fusarium solani
Aspergillus niger
607.2
1214.3
2428.6
[78]
AgCubical40–50Candida albicans (n = 30)
Candida glabrata (n = 30)
Candida tropicalis (n = 30)
8
30
31
[76]
AgSpherical30–50Candida albicans (n = 30)
Candida glabrata (n = 30)
Candida tropicalis (n = 30)
11
37
35
[76]
AgWires250–300 Candida albicans (n = 30)
Candida glabrata (n = 30)
Candida tropicalis (n = 30)
21
51
48
[76]
AuCubical30–50Candida albicans (n = 30)
Candida glabrata (n = 30)
Candida tropicalis (n = 30)
10
45
46
[76]
AuSpherical35–50Candida albicans (n = 30)
Candida glabrata (n = 30)
Candida tropicalis (n = 30)
15
47
48
[76]
AuWires300–500Candida albicans (n = 30)
Candida glabrata (n = 30)
Candida tropicalis (n = 30)
30
70
73
[76]
ZnOSpherical20–40Candida albicans (n = 125)17.76 [74]
ZnONRNRCandida albicans (n = 10)11.3
MIC, minimal inhibitory concentration; NR, Not reported; LMW, Low-molecular-weight; HMW, High-molecular-weight.
Table 5. Activities of NPs tested on different fungal species expressed as the zone of inhibition (mm).
Table 5. Activities of NPs tested on different fungal species expressed as the zone of inhibition (mm).
Type of NPShapeSize (nm)Organism(s) TestedActivity (mm) Reference
CuOSpherical3–30Fusarium equiseti
Fusarium oxysporum
Fusarium culmorum
25
20
19
[79]
PdSpherical200Colletotrichum gloeosporioidesDay 2: 3.6
Day 4: 1.6
[80]
Fusarium oxysporumDay 2: 12.2
Day 4: 10.9
PdSpherical220Colletotrichum gloeosporioidesDay 2: 7.9
Day 4: 6.3
[80]
Fusarium oxysporumDay 2: 5.1
Day 4: 4.7
PdSpherical250Colletotrichum gloeosporioidesDay 2: 2.4
Day 4: 0.7
[80]
Fusarium oxysporumDay 2: 10.4
Day 4: 9.6
PdSpherical350Fusarium oxysporumDay 2: 1.5
Day 4: 1.3
[80]
PdSpherical550Fusarium oxysporumDay 2: 3.8
Day 4: 3.3
[80]
Nystatin/MgO/CuO Spherical8000–10000Candida albicans (AH201)
Candida albicans (AH267)
24.5 ± 1.7
14.3 ± 1.2
[81]
Nystatin Spherical8000–10000Candida albicans (AH201)
Candida albicans (AH267)
0.41 ± 0.23
0.5 ± 0.21
[81]
MgO/CuO Spherical8000–10000Candida albicans (AH201)
Candida albicans (AH267)
19.2 ± 1.6
1.3 ± 0.61
[81]
TiO2/BPE BNRNRCandida albicans (ATCC 14053)11.2 ± 0.02[82]
TiO2/BPE CNRNRCandida albicans (ATCC 14053)15.9 ± 0.04[82]
TiO2/BPE DNRNRCandida albicans (ATCC 14053)13.5 ± 0.04[82]
TiO2/BPE ENRNRCandida albicans (ATCC 14053)14.6 ± 0.01[82]
TiO2/BPE BNRNRPenicillum chrysogenum (MTCC 5108)10.2 ± 0.05[82]
TiO2/BPE CNRNRPenicillum chrysogenum (MTCC 5108)18.0 ± 0.03[82]
TiO2/BPE DNRNRPenicillum chrysogenum (MTCC 5108)15.0 ± 0.04[82]
TiO2/BPE ENRNRPenicillum chrysogenum (MTCC 5108)13.5 ± 0.08[82]
BPE, Brahmi plant extract; NR, Not reported.
Table 6. Activities of NPs tested on different fungal species expressed as a percentage.
Table 6. Activities of NPs tested on different fungal species expressed as a percentage.
Type of NPShapeSize (nm)Organism(s) TestedActivity (%)Reference
AgNR20–100Cladosporium cladosporioides
Aspergillus niger
50 µg/mL → 90
50 µg/mL → 70
[83]
AgPolygonal35 ± 15Candida tropicalis
Saccharomyces boulardii
25 µg/mL → >95
50 µg/mL → >95
25 µg/mL → <50
50 µg/mL → >95
[84]
AgSpherical5Colletotrichum gloeosporioides13 µg/mL → 73
26 µg/mL → 82
56 µg/mL → 89
[85]
AgSpherical24Colletotrichum gloeosporioides13 µg/mL → 74
26 µg/mL → 82
56 µg/mL → 89
[85]
ZnOSpherical30–45Erythricium salmonicolor12 mmol/L, day 7 → 71.3
12 mmol/L, day 10 → 51.1
9 mmol/L, day 7 → 58.3
9 mmol/L, day 10 → 44.6
6 mmol/L, day 7 → 48.9
6 mmol/L, day 10 → 23.6
3 mmol/L, day 7 → 36.9
3 mmol/L, day 10 → 14.1
[86]
AgNRNRRhizoctonia solani (AG1)





Rhizoctonia solani (AG4)





Macrophomina phaseolina





Sclerotinia sclerotiorum





Trichoderma harzianum





Pythium aphanidermatum
6 µg/mL → 75
8 µg/mL → 80
10 µg/mL → 90
12 µg/mL → 90
14 µg/mL → 90
16 µg/mL → 100
6 µg/mL → ≥90
8 µg/mL → ≥90
10 µg/mL → ≥90
12 µg/mL → 100
14 µg/mL → 100
16 µg/mL → 100
6 µg/mL → 100
8 µg/mL → 100
10 µg/mL → 100
12 µg/mL → 100
14 µg/mL → 100
16 µg/mL → 100
6 µg/mL → ≥95
8 µg/mL → 100
10 µg/mL → 100
12 µg/mL → 100
14 µg/mL → 100
16 µg/mL → 100
6 µg/mL → 80
8 µg/mL → 84
10 µg/mL → 90
12 µg/mL → 100
14 µg/mL → 100
16 µg/mL → 100
6 µg/mL → 100
8 µg/mL → 100
10 µg/mL → 100
12 µg/mL → 100
14 µg/mL → 100
16 µg/mL → 100
[87]
AgSpherical10–20 Bipolaris sorokiniana≥2 µg/mL → 100[88]
AgSpherical1–9Aspergillus flavus5 µg/mL → 0
15 µg/mL → 30
25 µg/mL → 58
35 µg/mL → 85
45 µg/mL → 98
60 µg/mL → 100
[89]
PEI1/AgSpherical20.6Rhizopus arrhizus1.6 µg/mL → 97[90]
PEI2/AgSpherical4.24 6.5 µg/mL → 94
SiO2Spherical9.92–19.8Rhizoctonia solani100 µg/mL → 93–100[91]
NR, Not reported; PEI1, polyethyleneimine 700,000 Da; PEI2, polyethyleneimine 65,000 Da.
Table 7. Current antifungal classes and their mechanisms of action.
Table 7. Current antifungal classes and their mechanisms of action.
ClassExamplesMechanism of Action
AllylaminesTerbinafine
Naftifine
Squalene epoxidase inhibition, responsible for conversion of squalene to ergosterol
AzolesClotrimazole
Miconazole
Ketoconazole
Fluconazole
Itrazonacole
C14-α demethylation inhibition of lanosterol, inhibiting ergosterol synthesis
EchinocandinsCaspofungin
Micafungin
Anidulafungin
β-(1,3)-D-glucan synthase inhibition, interfering with cell wall synthesis
PolyenesAmphotericin B
Nystatin
Candicidin
Ergosterol binding, forming pores and causing leakage, inhibiting proper transport mechanisms
AntimetabolitesFlucytosinePyrimidine analogue, interfering with nucleic acid synthesis
TriterpenoidsIbrexafungerpβ-(1,3)-D-glucan synthase inhibition, interfering with cell wall synthesis
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Slavin, Y.N.; Bach, H. Mechanisms of Antifungal Properties of Metal Nanoparticles. Nanomaterials 2022, 12, 4470. https://doi.org/10.3390/nano12244470

AMA Style

Slavin YN, Bach H. Mechanisms of Antifungal Properties of Metal Nanoparticles. Nanomaterials. 2022; 12(24):4470. https://doi.org/10.3390/nano12244470

Chicago/Turabian Style

Slavin, Yael N., and Horacio Bach. 2022. "Mechanisms of Antifungal Properties of Metal Nanoparticles" Nanomaterials 12, no. 24: 4470. https://doi.org/10.3390/nano12244470

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop