Next Article in Journal
The Destructive Tree Pathogen Phytophthora ramorum Originates from the Laurosilva Forests of East Asia
Next Article in Special Issue
Lichens—A Potential Source for Nanoparticles Fabrication: A Review on Nanoparticles Biosynthesis and Their Prospective Applications
Previous Article in Journal / Special Issue
Differential Antimycotic and Antioxidant Potentials of Chemically Synthesized Zinc-Based Nanoparticles Derived from Different Reducing/Complexing Agents against Pathogenic Fungi of Maize Crop
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Fungal–Metal Interactions: A Review of Toxicity and Homeostasis

by
Janelle R. Robinson
,
Omoanghe S. Isikhuemhen
* and
Felicia N. Anike
Department of Natural Resources and Environmental Design, North Carolina Agricultural and Technical State University, 1601 East Market Street, Greensboro, NC 27411, USA
*
Author to whom correspondence should be addressed.
J. Fungi 2021, 7(3), 225; https://doi.org/10.3390/jof7030225
Submission received: 5 March 2021 / Revised: 15 March 2021 / Accepted: 17 March 2021 / Published: 18 March 2021
(This article belongs to the Special Issue Fungal Nanotechnology)

Abstract

:
Metal nanoparticles used as antifungals have increased the occurrence of fungal–metal interactions. However, there is a lack of knowledge about how these interactions cause genomic and physiological changes, which can produce fungal superbugs. Despite interest in these interactions, there is limited understanding of resistance mechanisms in most fungi studied until now. We highlight the current knowledge of fungal homeostasis of zinc, copper, iron, manganese, and silver to comprehensively examine associated mechanisms of resistance. Such mechanisms have been widely studied in Saccharomyces cerevisiae, but limited reports exist in filamentous fungi, though they are frequently the subject of nanoparticle biosynthesis and targets of antifungal metals. In most cases, microarray analyses uncovered resistance mechanisms as a response to metal exposure. In yeast, metal resistance is mainly due to the down-regulation of metal ion importers, utilization of metallothionein and metallothionein-like structures, and ion sequestration to the vacuole. In contrast, metal resistance in filamentous fungi heavily relies upon cellular ion export. However, there are instances of resistance that utilized vacuole sequestration, ion metallothionein, and chelator binding, deleting a metal ion importer, and ion storage in hyphal cell walls. In general, resistance to zinc, copper, iron, and manganese is extensively reported in yeast and partially known in filamentous fungi; and silver resistance lacks comprehensive understanding in both.

1. Introduction

The increasing applications of fungal–metal interactions have led to the need for research on their contributions to fungal resistance [1,2]. In nature, metals serve as micronutrients required for fungal growth, however, in excess they can influence homeostatic systems. In agricultural and human medicine, there is an increasing occurrence of pathogen resistance to traditional antifungal agents which has expanded the incidence of fungal superbugs; this has led to increased research on metals as alternative fungistatic and fungicidal agents [3,4]. Fungi are also being employed in the green biosynthesis of nanoparticles due to their economic viability, high levels of natural metal resistance, and ease of mass production as antimicrobial agents [5,6,7]. Both instances highlight contributions to increased incidence of fungal-metal interactions, demonstrating the importance of further divulging the intricacies of their relationship.

1.1. Fungal–Metal Interactions

Metals can exist in various forms such as salts, oxides, sulfates, and nanoparticles. Fungi are able to utilize metal ions from these compounds after dissociation, which leaves unbound ions available for uptake and transport. For example, in the presence of water, copper sulfate (CuSO4) hydrates to copper (II) sulfate pentahydrate (CuSO4 5H2O) and then dissociates into Cu2+ + SO42−. Upon dissociation, Cu2+ can then be reduced by fungal proteins for uptake. More recently, metals in the form of nanoparticles have gained interest for use as antifungals, which has fueled the escalation of nanoparticle production [8,9,10]. Nanoparticles are particles that range from 1 to 100 nm in size and vary in shape, physiochemical, optical, and biological properties [11]. Ions dissociate from nanoparticles at a much lower rate, but are also available to interact with homeostatic systems [12,13].
In general, most ions have dedicated homeostatic systems to control import, export, storage, and transport within the cell (Table 1). Metal ion import and export often occurs through transmembrane channels, which are proteins that span the entirety of the membrane and protrude from both sides (e.g., transmembrane proteins Fet4, Zrt1, and Zrt2 in Figure 1) [14,15]. In some species, chelators, such as siderophores, also play a role in uptake. These organic, low molecular weight compounds have a binding capacity for certain metal ions, such as iron, and are imported into the cell through transmembrane channels [16,17]. As a mechanism of ion storage or detoxification, metallothioneins (MTs), cysteine-rich proteins that use metal ions as cofactors, bind free cytosolic ions which may be released back into the cellular environment in metal deficient conditions [18,19]. For the movement of ions to organelles for storage or as cofactors for protein functioning, intracellular transporters, such as Zrc1 (Figure 1) or Pic2 (Figure 2), are utilized [20,21]. If these systems are interfered with, homeostatic imbalance can cause toxicity.
Table 1. Fungal proteins involved in metal transport.
Table 1. Fungal proteins involved in metal transport.
MetalTransport TypeYeast
Transporters
ReferenceFilamentous Fungi TransportersReference
ZincImportZrt1, Zrt2[15,22]zrfA/B/C, UmZRT1/2, Zip1/2[23,24,25,26,27]
VacuolarCot1, Zrc1[20,28]--
Vacuole to CytosolZrt3[29]--
CopperImportCtr1, Ctr3, Fet4, Ctr4, Ctr5, Mfc1[30,31,32,33,34,35,36]CtrA2, CtrC, Ctr1, PaCtr2[37,38,39]
Cytosol to GolgiAtx1, Ccc2[34,40,41,42]--
MitochondrialPic2, Cox17[21,43,44]--
Cytosol to Sod1Lys7, Pccs[45,46]--
Mitochondrial Inner Membrane Space to Cytochrome c oxidaseSco1, Sco2, Cox11[42,47,48]--
Export--CrpA[49]
IronImportFet4, Smf1, Fet3/Ftr1, Fip1, Str3, Shu1, Str1, Str2, Str3[50,51,52,53,54,55,56,57,58,59]Fer2[60]
Within the NucleusNpb35, Nar1, Cfd1, Cia1[61,62]--
VacuolarPcl1, Ccc1[63,64]--
Mangan-eseImportSmf1, Smf2, Pho85[52,65,66,67]PcPho84, PcSmfs[68]
MitochondrialMtm1[69]PcMtm1[68]
Cytosol to Golgi LumenPmr1, Gdt1[70,71,72]--
Cytosol to Endoplasmic Reticulum LumenSpf1[73]--
VacuolarCcc1, Ypk9[64,74,75,76]PcCCC1[68]
ExportPmr1, Hip1[77,78,79]PcMnt[68]
SilverImportCtr1[80,81]--
MitochondrialPic2[21]--

1.2. Metal Toxicity and Resistance

Metal toxicity occurs via the oligodynamic effect, which was initially described in 1893 in algae Spirogyra nitida and Spirogyra dubia, as toxicity or death in organisms due to exposure to trace amounts of metals, such as copper, lead, iron, or zinc [82]. In fungi, this exposure can have effects ranging interference in ergosterol biosynthesis to reduced MT activity (Table 2) [83,84].
Table 2. Mechanisms of toxicity in yeast and filamentous fungi.
Table 2. Mechanisms of toxicity in yeast and filamentous fungi.
MetalMechanism of
Toxicity in Yeast
ReferenceMechanism of Toxicity in Filamentous FungiReference
ZincInterference of synthesis of iron-sulfur clusters[85,86]increased chitin deposition within the cell wall, preventing hyphal extension[87,88]
Interference in ergosterol biosynthesis[83]increased hyphal branching and apical swelling[88]
Cellular leakage, polarization, and increased membrane potential[83]interruption of conidia and conidiophore development (interference of reproduction)[87]
Reduced cell wall integrity[83]--
CopperReduced ergosterol biosynthesis[12,89]Generation of reactive oxygen species[90]
Reduced metallothionein activity[84]--
IronInterference of vacuolar transport encoding gene CCC1[91,92]Inability to acquire iron[60,93]
ManganeseDown-regulation of HTB2, HTA1, HTA1, HTBI, HHF[94,95]potentially associated to reduced functioning of manganese peroxidase[96,97,98]
SilverInterference in ergosterol biosynthesis[80,99,100]--
In an effort to counter metal toxicity, and toxicity in general, fungi develop methods of resistance which can include the alteration of the target protein to inhibit substrate binding, cellular antimicrobial efflux, antimicrobial inactivation or degradation, restricted uptake to prevent cellular interference, overproduction of targeted proteins to prevent the complete inhibition of biochemicals, and compensation for loss of function directly related to the antimicrobial [101]. Some of these resistance mechanisms are relevant to excessive metal exposure in fungi (Table 3). Presently, research utilizes yeast such as Saccharomyces cerevisiae to investigate cellular and molecular impacts of fungal–metal interactions, but thorough knowledge is lacking in filamentous fungi [102,103,104]. Due to the increase in fungal-metal interactions, we should ensure that metal resistance mechanisms in multiple types of fungi are well-understood. In this review, we summarize existing knowledge on fungal metal homeostasis of zinc, copper, iron, manganese, and silver. Conclusions and indications are presented to pave the way for further research.
Table 3. Mechanisms of metal resistance in yeast and filamentous fungi.
Table 3. Mechanisms of metal resistance in yeast and filamentous fungi.
MetalMechanism of Metal Resistance in YeastReferenceMechanism of Metal Resistance
in Filamentous Fungi
Reference
ZincUp-regulation of ZRC1 and COT1[83,105,106,107,108]storage of excess zinc in vacuoles and cell walls of spores and hyphae[109,110]
--zinc efflux[111]
--zinc metallothioneins[112]
CopperUp-regulation of CUP1 and CRS5[113]Up-regulation of crpA[81,114,115,116]
Down-regulation of FRE1 and FRE7, and CTR1[113]increased production of chelator copper oxalate[117,118,119]
IronUp-regulation of CCC1[64,120]Unknown, but could associated with reduction of siderophore biosynthesis[60,121]
Expression of plant ferritin genes[122,123,124]--
ManganeseUp-regulation of MNR1[65,67,125,126]Deletion of PcPHO84[68]
Down-regulation of PHO84, SMF1[67,125,126]Expression of PcMNT[68]
SilverExpression of CUP1-1, CUP1-2[81,115,116]Expression of crpA[90]
Down-regulation of PHO84[116]--

2. Fungal–Metal Interactions

Metals play critical roles in fungal homeostasis. They are required for various biochemical processes, usually as enzymatic cofactors. Metals most recognized for their importance in fungi are copper, iron, zinc, and manganese. Pertaining to zinc, approximately 5% of fungal proteomes correlate to zinc-binding proteins, and 8% of yeast genomes correlate to zinc-binding proteins. In the model yeast S. cerevisiae, large portions of these zinc-binding proteins are related to critical functions, including DNA binding (31% of zinc-binding proteins), the regulation of transcription (25%), transcription factor activity (19%), and response to chemical stimuli (15%) [105,107,108]. Fungal–copper interactions are necessary for the activation of metalloproteins involved in biochemical processes. This includes the activation of superoxide dismutase, which is responsible for cellular detoxification of reactive oxygen species (ROS), virulence in pathogenic species, and activation of cytochrome c oxidate, a catalyst within the electron transport chain [39,48]. Iron is also essential for fungal virulence in pathogenic species, most importantly as an integral component of iron-sulfur clusters which are required for the activation of nuclear proteins involved in DNA repair [61]. Manganese also plays a critical role in fungi, in particular, in filamentous species where it (or copper) is required for the activation of manganese peroxidase. Dependent on nutrient availability, white-rot fungi utilize manganese peroxidase as a secondary metabolite to depolymerize lignin for nutrients; others are manipulated for increased manganese peroxidase production and extraction for use in the degradation of organo-pollutants [96,98].
Very few metals that are not considered essential have also been identified in some fungal–metal interactions; these include magnesium and molybdenum. Magnesium is a well-known micronutrient in other eukaryotic organisms, however, its homeostasis in fungi is undetermined. Only in recent years has magnesium been identified as a requirement for virulence in the agriculturally relevant fungus Magnaporthe oryzae [127]. Molybdenum is a metal that is discussed significantly less in eukaryotic homeostasis. It has only been identified as a cofactor for four human proteins, and in fungi it has only been suggested that it plays an unidentified role as a nitrate reductase and a xanthine dehydrogenase [128,129]. Other metals such as silver, gold, lead, nickel, and cadmium have only been implicated in fungal–metal interactions related to toxicity, nanoparticle myco-synthesis, and heavy metal myco-remediation, but information pertaining to homeostasis is limited [103,130,131].

2.1. Zinc

Zinc is a transition metal required for fungal survival and is necessary for various functions, including the structuring of nucleic acids, physical growth and, most predominately, protein folding [132,133]. In its role in DNA binding, zinc presents itself in class III zinc finger proteins, also known as zinc cluster proteins (Zn(II)2Cys6), found only in Ascomycetes (with the singular exception of Lentinus edodes) [107,134,135,136]. This protein class binds DNA, which is critical for the transcriptional activation and regulation of gene products [105,134].
In agriculture, fungal infections threaten food security by increasing global crop loss [137,138]. Traditionally, antifungal azoles have been used to combat disease, but with the emergence of azole-resistant pathogens, scientists have begun to develop possible alternatives, such as zinc-containing compounds [138,139]. Reports have demonstrated that zinc oxide nanoparticles (ZnO NPs) can control postharvest mold, plant wilts, and grey mold disease caused by Aspergillus niger, Fusarium oxysporum, and Botrytis cinerea, respectively [7,140,141,142,143]. It has also been demonstrated that ZnO NPs can significantly reduce the production of the mycotoxin fusaric acid from F. oxysporum [144]. This is significant because mycotoxins are common secondary metabolites of fungal pathogens with high rates of toxicity against cereal crops that can result in crop loss, and if consumed can result in a wide array of diseases in livestock [145,146]. Fusaric acid, in particular, can inhibit the production of dopamine-beta-hydroxylase, which acts as a messenger of signals within the nervous system and is responsible for altering the enzyme tyrosine hydrolase, which is involved in a rate-limiting step in catecholamine synthesis [147,148,149]. Zinc perchlorate Zn(ClO4)2 and zinc sulfate (ZnSO4) also inhibit mycelial growth that produces mycotoxins and reduces the production of mycotoxins themselves [150,151].

2.1.1. Zinc Transport and Homeostasis

Many fungi have mechanisms of zinc transport similar to that of other eukaryotes, through the ZRT (zinc regulated transporter)-IRT (iron-regulated transporter)-like protein (ZIP) family and the cation diffusor facilitator (CDF) protein family [152,153]. In S. cerevisiae, zinc transport occurs through several protein groups; the ZIP protein family (via Zrt1, Zrt2, and Zrt3), the CDF protein family (via Zrc1, Cot1, and Msc2), the ferrous transport protein Fet4, and others (Figure 1) [105,107,108,133,154,155]. Zrt1 and Zrt2 are high and low-affinity plasma membrane transporters, respectively; both ZRT1 and ZRT2 are upregulated in zinc-deficient conditions and repressed when zinc conditions are favorable [15,22,156]. In an excess-zinc environment, Zrc1 and Cot1 mediate zinc transport from the cytosol into the vacuole to prevent toxicity [20,28]. In a zinc-limiting environment, zinc is released back into the cytosol from the vacuole via Zrt3 or is scavenged by zincophore Zps1 [106,157,158]. Zap1 regulates the transcription of ZPS1 and contains two activators, Ad1 and Ad2, either independently activated or inactivated by the direct binding of zinc ions [105,108,159,160]. These mechanisms effectively control intracellular zinc uptake and help prevent excess accumulation in S. cerevisiae.
In filamentous Ascomycota, such as Apergillus fumigatus, genes in the ZIP family (zrfA, zrfB, zrfC, zrfD, zrfE and zrfH) also regulate zinc transport [23,27,161]. zrfA and zrfB, orthologues of S. cerevisiae ZRT1 and ZRT2, respectively, encode zinc membrane transporters that operate in acidic, low-zinc environments and are activated by transcription factor ZafA [161,162]. Conversely, the zrfC gene product is an alkaline zinc transporter activated in high pH, zinc-limiting conditions [23,27]. zrfD/E/H are not restricted by pH and can function in either acidic or alkaline environments [23]. In F. oxysporum, zrfA and zrfB are also zinc importers regulated by transcription factor ZafA [163]. During infection, ZafA allows F. oxysporum to adapt to a zinc-limiting environment, such as if the host enacts nutritional immunity to deprive it of this essential metal [163]. Basidiomycetes have similar homology. Ustilago maydis UmZRT1 and UmZRT2 genes, and Cryptococcus neoformans Zip1 and Zip2 are homologous to S. cerevisiae ZRT1 and ZRT2, respectively, with similar transport function [20,24,26]. Similarities also exist in the prevention of zinc over-accumulation. C. neoformans Zrc1 is homologous to S. cerevisiae Zrc1 and mediates zinc transport into the vacuole to prevent toxicity and decrease zinc sensitivity [20].
Mechanisms of zinc uptake and transport in fungi are mostly conserved through S. cerevisiae ZIP proteins and homologs. The next section will discuss how negative homeostatic interventions can result in toxicity.

2.1.2. Zinc Toxicity

Zinc-based antifungal compounds have mechanisms of toxicity that vary between species. Zinc pyrithione (ZPT), is a zinc ionophore often used to treat fungal dandruff caused by Malassezia spp. and induces toxicity by increasing cellular zinc uptake [164,165]. ZPT also causes partial mitochondrial malfunction by inhibiting mitochondrial synthesis of iron-sulfur clusters, which are integral in electron transport, respiration, and DNA repair and replication [165,166]. In contrast to Malassezia spp., ZPT toxicity in S. cerevisiae is not a result of increased zinc uptake, rather of increased copper uptake which overloads homeostatic systems [164,167,168]. ZnO NPs are also being explored for their antifungal properties. In S. cerevisiae, ZnO NPs reduce ergosterol biosynthesis which, in turn, increases cellular leakage (up to 24%) and depolarization, reduces cell wall integrity, and increases the occurrence of ROS [83]. In filamentous fungi, mechanisms of toxicity are not well-studied. In ericoid fungi, zinc ions reduced hyphal growth by increasing chitin deposition within the cell wall, preventing hyphal extension; zinc also increases hyphal branching and apical swelling, resulting in atypical hyphal morphology [88]. In the molds, excessive zinc exposure reduces hyphal growth, alters hyphal morphology and interrupts conidia and conidiophore development, limiting reproductive capabilities [87]. Zinc sensitivity can aid in the reduction of fungal pests; however, the development of tolerance and resistance can be an impedance.

2.1.3. Zinc Tolerance and Resistance

High-zinc environments can be detrimental to fungi; therefore they must possess resistance mechanisms to overcome toxicity. In yeast, resistance relies on the upregulation of Zrc1 and Cot1, which sequester Zn2+ to the vacuole (up to 100 mM) in S. cerevisiae, or the endoplasmic reticulum in C. albicans (Zrc1) [105,106,107,108,169]. Khouja et al. also described a resistance mechanism via OmFET in S. cerevisiae, though it is not yet fully understood [170]. They suggest that OmFET plays a role in Zn2+ uptake, and in that role increases tolerance through interactions with Mg, where Mg competes with Zn2+ for uptake, increasing intracellular Mg and restricting Zn [170]. In filamentous fungi, zinc resistance is not only attributed to vacuolar sequestration, but also to storage in the cytoplasm, storage in cell walls of spores and hyphae, and cellular efflux; and in ectomycorrhizal fungi, the presence of metallothionein-like peptides confers Zn2+ resistance [110,171,172,173]. To further investigate how fungi cope with toxic levels of other micronutrient metals, this review also assessed cellular interactions with copper.

2.2. Copper

Copper is also a transition metal and presents itself in oxidation states copper(I), Cu+, and copper(II), Cu2+ [32,48]. It is essential to agriculture and human medicine where it can serve as a fungicidal or fungistatic agent, or be the determining factor for virulence [174,175]. Some fungal pathogens heavily rely on copper exporters to prevent host-enacted copper toxicity or import machinery to maintain virulence. In both clinical and agricultural settings, fungal exposure to excess copper can result in ionic imbalance. Therefore, homeostatic mechanisms to maintain healthy intracellular copper levels are critical.

2.2.1. Copper Transport and Homeostasis

Generally, copper cannot permeate the plasma membrane and requires membrane transporters for uptake [32,48]. Before internalization, copper must exist as Cu+ (cuprous oxide); however, in the environment, it often exists as Cu2+ (cupric oxide) and must undergo reduction. In S. cerevisiae, cupric reductase Fre1, transcribed by Mac1, reduces Cu2+ to Cu+, making it readily available for uptake via high–affinity membrane transporters of the copper transporter (Ctr) protein family, Ctr1 and Ctr3 or low-affinity copper transporter Fet4 (Figure 2) [30,31,32,33,176]. Transcription of CTR1 and CTR3 is also regulated by transcription factor Mac1, which regulates transcription based on copper availability; copper depletion results in the upregulation of CTR1/3, and copper repletion results in downregulation [30,177].
After uptake, Cu+ serves as enzymatic cofactors. Apoproteins within the secretory pathway require copper for proper functioning, such as the multicopper oxidase Fet3, which is necessary for ferrous iron, Fe(II), uptake, and oxidation [53,178,179,180]. FET3 is regulated by transcription factor Aft1 (activator of ferrous transport) in iron-deficient conditions and its gene product contains four Cu+ binding sites where copper serves as a cofactor for enzyme activation [53,178,181]. Unmetalated Fet3 reduces cell growth in iron-limiting conditions, demonstrating the importance of copper transport [44,182].
Another enzyme dependent on copper is the cytoplasmic Cu/Zn superoxide dismutase (Sod1). This is an antioxidant for superoxide anions (O2•−) [183,184]. O2•− are ROS that cause cellular damage and toxicity and must be effectively dismutated to prevent stress; therefore, delivery of copper to Sod1 is critical [45,185,186]. In S. cerevisiae, the cytosolic copper chaperone Lys7 acquires Cu+ and delivers it to Sod1, with high specificity [45]. Once Sod1 is metalated, it is then able to catalyze the dismutation reaction that results in O2•− being successfully detoxified to hydrogen peroxide (H2O2) and molecular oxygen (O2); H2O2 is now readily available for further detoxification to water via catalyst Cct1 [183,187,188]. Cu+ transport to MTs Cup1 (also known as Cup1-1 and Cup1-2) and Crs5 is also integral to cellular detoxification [18,189]. Both MTs are regulated by transcription factor Ace1 (also known as Cup2), which activates the transcription of CUP1 and CRS5 at elevated copper concentrations [167,189]. Cup1 and Crs5 contain 8 and 11-12 Cu+ binding sites, respectively, and are responsible for buffering cytosolic copper to maintain safe intracellular copper concentrations [189,190,191]. Though Crs5 has a greater copper binding capacity, it plays a much smaller role in detoxification due to its promoter region, which only has one recognition sequence, compared to four in CUP1 [189,190,191].
S. pombe follows a pattern of copper transport similar to S. cerevisiae. Extracellular Cu2+ is reduced to Cu+ by cell surface reductases before uptake [34,36]. Cu+ can then be transported across the cell membrane, depending on the current cell cycle [34,35,36]. During mitosis, an integral membrane complex composed of proteins Ctr4 and Ctr5 are responsible for Cu+ uptake, and during meiosis, Mfc1 (localized in the forespore membrane) is responsible [34,35,36]. Expression of ctr4+ and ctr5+ is regulated by transcription factor Cuf1, and expression of mfc1+ is regulated by transcription factor Mca1, both of which are activated or deactivated by the absence or presence of sufficient copper levels, respectively [34,36]. Once inside the cell, copper chaperones such as Cox17, Pccs, and Atx1 transport Cu+ to respective organelles [46]. Pccs is a four domain, cytosolic chaperone. The first three domains are responsible for transporting Cu+ to unmetalated Sod1 in a copper-limited environment, activating Sod1 [46]. In high copper environments, the fourth domain acts as a copper buffering system, sequestering Cu+ to prevent toxic cytosolic levels [46]. Atx1 in S. pombe plays a similar role to Atx1 in S. cerevisiae. In S. pombe, Atx1 is also located in the cytosol and responsible for carrying Cu+ to Ccc2 [34,42]. Peter et al. and Beaudoin et al. described how Atx1 was also used for copper transport to copper amine oxidases (CAOs), a group of catalysts not present in S. cerevisiae [34,42]. Atx1 shuttles Cu+ to an active site on the CAO, where copper (and another required cofactor, 2, 4, 5-trihydroxyphenylalanine quinone) activates it [34,42,192]. S. pombes Cox17 is an orthologue to S. cerevisiae Cox17, sharing 38% identity and is located in the mitochondrial intermembrane space [42,48]. Once Cox17 acquires Cu+ it is delivered to Sco1, Sco2, and Cox11 for copper loading to cytochrome c oxidase subunits [42,47,48].
Filamentous fungi are also important in assessing copper homeostasis, as these organisms depend on copper for growth and virulence in pathogenic species. In the pathogenic Ascomycete Aspergillus fumigatus, studies have shown similarities to S. cerevisiae and S. pombe in copper uptake. Cu2+ must also be reduced before uptake, however, there is some ambiguity regarding the reductases responsible [39]. This reductase has been referred to as unknown ferric reductase (“Fre?”), a general Fre reductase, and metallo-reductase Afu8g01310 (homolog of S. cerevisiae FRE or FRE3) [39,193,194]. After reduction, CtrA2 and CtrC (both homologs of S. cerevisiae Ctr1) transport Cu+ into the cytosol and serve as enzymatic cofactors [37,39]. CtrA2 and CtrC are regulated by transcription factor MacA (also referred to as AfMac1) which senses low copper concentrations and activates CtrA2 and CtrC [39,49,195,196]. Conversely, in high copper concentrations, transcription factor AceA activates P-type ATPase CrpA as a defense mechanism for copper export and is responsible for extended life and virulence [39,49,195,196].
Limited knowledge exists on copper homeostasis in Basidiomycetes. Studies in two Basidiomycetes, the brow-rot fungus Fibroporia radiculosa and the edible white-rot fungus Pleurotus ostreatus, reported some details. In F. radiculosa, only the regulation of intracellular Cu+ concentration has been unveiled, by three, unnamed copper ATPases and one gene of unknown function, CutC, [197]. In P. ostreatus, membrane protein Ctr1 is involved in copper uptake and shares homology with the low-affinity copper transporter PaCtr2 of the Ascomycete Podospora anserine (20%) and the high-affinity S. cerevisiae copper transporter, Ctr1 (20%) [38]. This review shows that copper homeostasis is well-studied in S. cerevisiae and S. pombe; however, more research is needed in other Ascomycetes and Basidiomycetes.

2.2.2. Copper Toxicity

Copper contains antifungals that have been investigated against various fungi. In S. cerevisiae, cupric sulfate (CuSO4) and copper oxide nanoparticles (CuO NPs) significantly reduce growth in a dose dependent manner, with the toxicity of both potentially related to Cup2 [113,198,199]. Deletion of CUP2 increases copper sensitivity, suggesting that a mechanism of toxicity could be reducing or inactivating its regulation, resulting in decreased Cu+/MT binding and increased cytosolic Cu+ [200,201]. Giannousi et al. found that CuO NPs cause DNA damage that interferes with replication and increases lipid peroxidation, reducing membrane lipid content, resulting in porous cells [202]. In Candida spp., CuO NPs have also shown toxic capabilities by inducing porous cell membranes [12]. Copper(II) complexes which have been shown to exhibit fungicidal and fungistatic activity in species that have a history of azole resistance appear to have a similar mechanism by reducing ergosterol content [203,204,205,206,207]. In filamentous fungi, copper also has dose-dependent toxicity. In the agricultural pathogen Rhizoctonia solani, a copper (II)–lignin hybrid had high efficacy and significantly reduced the number of plants attacked by R. solani [84,208]. In some instances, fungi can overcome toxicity by increasing their tolerance, which may be beneficial in the case of nanoparticle production, but can become a nuisance in pathogenic species.

2.2.3. Copper Tolerance and Resistance

Since copper is implicated as an antifungal agent, its ability to evade copper toxicity must be continuously evaluated. In S. cerevisiae, short-term exposure to CuSO4 causes significant regulation of open reading frames (ORFs) responsible for cellular detoxification and Cu+ uptake [113]. Exposure results in the upregulation of CUP1 (~20-fold,) and CRS5 (~8-fold) and the downregulation of FRE1, FRE7, and CTR1 (0.07, 0.08, and 0.10-fold, respectively) [113]. This fold change, and increased CuSO4 sensitivity in cup2Δ mutants indicates MTs, coupled with decreased Cu2+ reduction and decreased Cu+ uptake, are likely to be employed as mechanisms of copper resistance [113,200].
Less is known about copper resistance in filamentous fungi. In Aspergillus spp., P-type ATPase CrpA has Cu+ exporting activity that aids in cellular detoxification, increasing Cu+ resistance [90,193,209]. High-affinity copper importers, CtrA2 and CtrC, may be involved in resistance, but are still under investigation [37,49]. In Fusarium graminearum, copper exposure upregulates FgCrpA (ATPase exporter) and the MT FgCrdA as a means to prevent over accumulation, with the predominate method being Cu+ export activity [14]. In F. oxysporum, upregulation of oxidoreductase activity may decrease susceptibility to oxidative stress that can be induced by excessive copper exposure [210]. In Basidiomycetes, some progress has been made in identifying resistance mechanisms in F. radiculosa, where increased production of copper oxalate increases resistance [119]. However, this is the extent of the knowledge.

2.3. Iron

Iron (Fe) is a transition metal belonging to group eight of the periodic table and can exist as ferrous (Fe2+) or ferric (Fe3+) iron [211,212]. As an essential nutrient, Fe is significant for the virulence of fungi that cause disease. In A. fumigatus and F. oxysporum, survival depends on the ability to sequester iron from the host and a well-functioning homeostatic system to maintain this delicate balance [213,214]. Incapacitating the ability to do so reduces virulence and becomes a growth limiting factor, such as in the use of excessive amounts of Fe to completely overrun homeostatic systems [215,216,217]. Thus, homeostatic mechanisms are integral.

2.3.1. Iron Transport and Homeostasis

Generally, in S. cerevisiae, two iron uptake systems are described, the reductive and nonreductive systems. The reductive system recognizes Fe2+ salts and chelates for uptake through importers, while the nonreductive systems utilizes iron siderophores [218,219,220,221]. In the reductive system, high-affinity (aerobic) and low-affinity (anaerobic) transporters are responsible for ferric and ferrous iron transport, respectively [222,223]. For low-affinity uptake, iron must be reduced by ferric reductases Fre1 or Fre2, initially described by Lesuisse et al. in 1987 and later coined Fre1 and Fre2 by Georgatsou and Alexandrakin in 1994 [221,224]. Since then, both metallo-reductases have also been found to reduce both cupric and ferric ions, where FRE1 expression induces the reduction of Cu2+ when transcription factor Mac1 is bound, and Fe3+ reduction occurs via binding of transcription factor Aft1 [176,181,225]. After Fe3+ reduction, Fe2+ is then ready for uptake by a six domain, transmembrane, metal transporter, Fet4 [54,55]. Fet4 can also import other metals, but is mostly responsible for Fe2+ uptake in iron-restricted cells [223]. In anaerobic conditions, transcription factor Aft1 is required for activation, and in aerobic conditions, expression of FET4 is repressed by Rox1, which has two binding sites in the FET4 promoter region [223]. This repression is necessary to prevent the unintended uptake of toxic metals, such as Cd, where it is demonstrated that rox1Δ mutants have increased sensitivity to Cd under aerobic conditions [223,226]. A second, less utilized iron transporter in the low-affinity uptake system is Smf1, responsible for the uptake of the Fe2+/H complex [51,52]. This metal transporter is mostly known for the uptake of Cu, Mn, and Cd; however, in a study completed by Cohen et al. in 2000, it was shown that overexpression of SMF1 also results in significant iron uptake [52,65,227]. High-affinity iron uptake is also part of the reductive system. In low-iron conditions, this system dissociates and reduces ferric iron, via Fre1 and Fre2, from a wide array of Fe3+ substrates such as ferric chelates, salts, and siderophores [218,219]. Fe2+ then transitions through the Fet3/Ftr1 complex [58]. Fet3 is activated by transcription factor Aft1 in iron-deficient conditions and contains four Cu+ binding domains that must be metalated for activation [53,178,181]. Activated Fet3 goes through an aerobic reaction that oxidizes Fe2+ to Fe3+ for passage to the cytosol via iron permease Ftr1 [58,178]. The final destination and the cell’s utilization of Fe3+ is not fully elucidated.
The nonreductive system utilizes siderophores. S. cerevisiae is incapable of producing siderophores, but can sequester siderophores produced by other microorganisms via siderophore iron plasma membrane transporters Arn1—Arn4 [16,224]. Arn1 transports ferrichrome into the cell for iron acquisition; however, Arn1 is not always readily available in the plasma membrane because it is localized to endosomes or is routed to vacuoles for degradation when ferrichrome is unavailable [16,220,228]. When ferrichrome is present, Arn1 is routed through the plasma membrane, where ferrichrome adheres to either the low or high-affinity binding site and is transported to the cytosol [16,220,228]. It remains intact in the cytosol and serves as an intracellular Fe3+ storage reservoir until the cell needs iron; in this event, Fe3+ is reduced via metallo-reductases, or released via ferrichrome degradation [16,220,228,229]. Arn2 (also known as Taf1) is the second siderophore transporter in the ARN family, responsible for transporting tri-acetyl-fusarinine to the cytosol; it is unclear if Arn2 is located anywhere else aside from the plasma membrane when tri-acetyl-fusarinine is unavailable [220,230,231]. The literature is not very informative on the functions of tri-acetyl-fusarinine, but it does appear to have a similar role to ferrichrome as a store reservoir for ferric iron [230,231]. Arn3 (also known as Sit1) is a transporter for ferrioxamine B and is situated within intracellular vesicles. It appears to have a similar function to Arn1 and can progress to the plasma membrane when ferrioxamine B is available [229,232]. After ferrioxamine B is transported inside the cell, it is stored in the vacuole, likely for subsequent dissociation [232]. The first three mentioned siderophores transported by Arn1–Arn3 belong to the hydroxamate class of siderophores. However, the final transporter Arn4 (also known as Enb1) transports a siderophore of the catecholate class, ferric entero-bactin [220,233]. Unlike the other siderophore transporters, Arn4 remains at the plasma membrane regardless of the presence of its substrate [218]. Philpott and Protchenko suggested the difference in plasma membrane cycling between hydroxamate and catecholate transporters may be due to the possibility that there are toxins that can adhere to the hydroxamate transporters and not the catecholate transporters [218]. In the act of self-preservation, those transporters remove themselves as a potential source of toxicity [218]. Ferric entero-bactin is not well-studied in S. cerevisiae, but based on the function of other siderophores it may be reasonable to conclude that, upon cellular entry, ferric entero-bactin is also used as an Fe3+ storage system.
After Fe uptake, there are many intracellular destinations. Two briefly discussed here are the cytosol and the nucleus [61,62]. In the cytosol, iron–sulfur assembly (CIA) proteins Npb35 (binds two Fe–S clusters), Nar1, Cfd1 (binds one Fe–S cluster), and Cia1 form an iron–sulfur complex [61,62,234]. These complexes transfer Fe–S clusters to various apoproteins for activation [61,62,234]. In the nucleus, CIA proteins deliver Fe–S clusters to various nuclear proteins involved in DNA repair and replication [61,235].
Iron homeostasis in the fission yeast S. pombe is also well-studied and has three mechanisms of iron uptake [236]. One involves cell surface ferric reduction, and the other, in contrast to S. cerevisiae, involves the production of siderophores to capture extracellular iron and heme [236]. The first iron uptake system described here is through use of siderophore synthesis [237]. Under iron-deficient conditions, Sib2, a catalyst for ferrichrome synthesis, hydroxylates ornithine to N5-hydroxyornithine, a newly formed hydroxy-mate group molecule, and then undergoes processing by Sib1 [236,237]. This non-ribosomal peptide synthase yields the desferri-form of ferrichrome [236,237]. Schrettl, Winkelmann, and Haas suggested that the resulting ferrichrome is excreted from the cell to capture extracellular Fe3+ from the surrounding environment [237]. In an iron-dependent response, transcription factor Fep1 activates ferrichrome transporters Str1, Str2, and Str3, and the iron-loaded ferrichrome re-enters the cell (predominately by way of Str1) [59,63]. S. pombe is also able to import exogenous iron-loaded ferrioxamine B via Str2 [63]. In addition to the previously mentioned siderophore functions, it had also been suggested that, as in S. cerevisiae, imported siderophores also serve as iron storage vesicles [63,237].
The second iron uptake mechanism employed by S. pombe is the high-affinity, reductive system that depends on cell surface ferric reductase Frp1. frp1+ shares 27% homology with the S. cerevisiae Fe3+/Cu2+ reductase encoding gene, FRE1, and reduces extracellular Fe3+ to Fe2+ [238]. Transcription of frp1+ may also have some functional relation to the vacuole/cytoplasmic transporter Abc3 that transports iron from the vacuole to the cytosol in iron-deficient conditions [238,239]. Pouliot et al. found that abc3Δ mutants resulted in the activation of frp1+; however, a nucleotide-based transcription factor directly linked to frp1+ has not yet been determined and it appears to be solely activated or repressed by the absence or presence of iron, respectively [238,239]. After ferric reduction, Fe2+ enters an oxidase-permease complex, similar to that of the S. cerevisiae Fet3/Ftr1 complex, composed of proteins Fio1 and Fip1 [50]. Fio1 is a Fe2+ oxidase that shares 37% homology with the S. cerevisiae Fet3, and in an iron deprived environment, oxidizes Fe2+ in preparation for transfer across the plasma membrane via Fip1 [50]. fip1+ is a ferrous permease having 46% homology with the S. cerevisiae Ftr1 [50,236].
Heme is an iron-containing compound and its acquisition and biosynthesis are the finally discussed mechanisms of iron uptake in S. pombe. It is notable to state that while S. cerevisiae does utilize heme in other processes such as respiration and ergosterol biosynthesis, it has not been determined to be used to acquire iron [240,241]. S. pombe imports exogenous heme for iron uptake through Str3 and Shu1 [56,57]. Shu1 is a plasma membrane protein induced during iron deprivation, when heme biosynthesis is not attainable, or if Fep1 is inactivated [56,57]. The second protein involved in heme uptake is Str3, previously mentioned as a part of a ferrichrome transporter family (Str1, Str2, and Str3). Str3 shares the lowest homology (25.1%) with Str1 when compared to Str2 (29%), and its substrate specificity is undetermined [57,59,63]. Iron release and utilization from heme is not yet fully understood in S. pombe; however, studies in C. albicans (and other fungi) show that heme degradation is catalyzed upon cellular entry via heme oxygenase [56,57,242]. S. pombe also biosynthesizes heme and is encoded by hem1+, hem2+, hem3+, hem12+, hem13+, hem14+, hem15+, and ups1+ [56,57]. In iron-deficient conditions, a cascade of events between the mitochondria and the cytoplasm occurs to synthesize heme for further utilization [56,57,243].
In addition to iron acquisition in S. pombe, regulation mechanisms must be in place to prevent over-accumulation. Mercier, Pelletier, and Labbé identified the gene pcl1+ to play a role in vacuolar iron storage [244]. pcl1+ shares homology to S. cerevisiae Ccc1, an iron vacuolar transporter, and it has been shown that pcl1Δ mutants have increased sensitivity to iron; this together with the study of Mercier, Pelletier, and Labbé suggests that Pcl1 might play a similar role in iron storage in S. pombe [63,239]. As mentioned, the final destinations of heme are somewhat unclear, but based on research in other fungi, heme may be degraded, and literature suggested that there may be a group of proteins responsible for transporting those ions to the vacuole for degradation or storage [56,57,245]. Much is known about iron homeostasis in S. pombe; however, there are apparent gaps in knowledge of specific processes.
In filamentous fungi, iron homeostasis is less documented. It has been investigated in U. maydis, a pathogenic fungus that causes corn smut disease and whose virulence is associated with iron acquisition [60,121]. There are two iron uptake mechanisms, one through hydroxamate siderophores, and the other an oxidase-permease system, similar to S. pombe [60,233,246,247]. In the latter, exogenous ferric iron is reduced by a seemingly unknown reductase (possibly Fer9) and then re-oxidized by ferroxidase Fer1 for uptake through the high-affinity ferric iron permease Fer2 [60,121]. In the former, siderophore iron uptake is mediated by siderophore biosynthesis encoding genes Sdi1 and Sid2, and both are negatively regulated by transcription factor Urbs1 [60,121,246,247]. These siderophores play a role in iron acquisition; however, deletion mutants showed they are not necessary for virulence [121].

2.3.2. Iron Toxicity

Iron is involved in many biological processes, but can be toxic in excess. Studies have shown its toxicity in S. cerevisiae and fungal pathogens, but they have also demonstrated that targeting and interfering in iron acquisition mechanisms can also be detrimental. Reports indicate that iron or iron compounds are fungistatic against F. oxysporum and its mycotoxins in a dose dependent manner [216,217]. In discussing iron toxicity, it is also important to note that the interference of homeostatic systems can result in the inhibition of iron acquisition, which can also be toxic. Leal et al. demonstrated this with the utilization of lactoferrin, an iron-binding glycoprotein, as a topical agent to obstruct iron uptake mechanisms of A. fumigatus and F. oxysporum in mice [92,93]. Results indicated that, during corneal fungal infection, these fungi acquired iron through siderophores and that the iron-binding agent blocked the ability of the pathogen to acquire siderophore-bound iron, highlighting the inability of the fungi to proliferate without access to iron [93]. In S. cerevisiae, iron toxicity is related to the ability of the cell to transport cytosolic iron to the vacuole via Ccc1 [91,248]. Lin et al. showed that ccc1Δ mutants could not transfer cytosolic iron to the vacuole under anaerobic conditions, even with the overexpression of iron mitochondrial transporter Mrs3, effectively inducing toxicity [248]. The ability to alter and control homeostatic mechanisms are determinants of the fungal ability to resist excessive iron concentrations.

2.3.3. Iron Tolerance and Resistance

S. cerevisiae achieves iron resistance through the downregulation of iron import systems via Aft1, or activation of vacuolar transporter Ccc1 [64,249]. Ccc1 is regulated by the iron sensitive transcription factor Yap5; removal of YAP5 increases iron sensitivity, while its overexpression dramatically reduces cytosolic iron [120]. It may be worth the effort to investigate how the overexpression of CCC1 affects iron resistance and the vacuolar ability to store excess iron in order to prevent toxicity. Another vacuolar gene, VMA13, might also play a potentially novel role in iron tolerance [250]. Vma13 is commonly known as a vacuolar H+-ATPase subunit that plays a role in vacuolar acidification; however, a study involving vma13Δ mutants showed that they experienced increased sensitivity to iron deprivation, suggesting Vma13 plays a role in iron import [250]. The function of VMA13 in iron homeostasis combined with its role in vacuolar acidification should be studied to determine if mutants can also help increase iron resistance. Another method of iron resistance in S. cerevisiae is the expression of ferritin related genes. Ferritin is an iron storage protein found in many other eukaryotes, but is not native to fungi [122,123,124]. Its effects on increased iron resistance and storage capacity in yeast has been investigated and results indicate that the expression of human, soybean, and tadpole ferritin genes (HuFH, SFerH1/SFerH2, and TFH, respectively) resulted in the increased ability of yeast to store and carry higher concentration of iron [122,123,124]. Llanos et al. showed the ability of soybean ferritin genes, SFerH1 and SFerH2, to increase iron resistance in ccc1Δ mutants [122]. This is significant because, even without the natural vacuolar detoxification system, yeast cells with soybean ferritin were still able to store increased concentrations of iron and evade toxicity.
Fewer studies report on iron resistance in other fungi, but several inferences can be made based on knowledge of iron homeostasis. In S. pombe, ferrichome production, excretion, and subsequent uptake are used to acquire extracellular Fe3+ in iron-deficient conditions [59,63,236,237,251]. The engineering of cells to overexpress Sib2 and Sib1 could potentially serve as extra storage vesicles for any excess cytosolic iron acquired by the cell [59,63,236,237,251]. It is not clear how ferrichrome is excreted from the cell after production, therefore this exact mechanism would first need to be identified and well-studied to determine if inhibiting excretion would have any other adverse effects on cellular health. U. maydis also biosynthesizes siderophores (hydroxamate) via sid1 for iron uptake which could also be investigated for increased production for storage of excess iron [60,121].

2.4. Manganese

Manganese (Mn) is a transition metal and also an essential micronutrient in fungi. In agriculture, Mn compounds reduce mycelial growth of fungal pathogens [252,253]. In other pathogenic fungi, Mn2+ is required for virulence [254]. Some lignocellulose degrading enzymes also require Mn2+, such as manganese-dependent peroxidase, which white-rot fungi express during lignocellulose degradation, integral to nutrient uptake [255,256]. Many fungal species rely on Mn2+ and homeostatic mechanisms must exist to ensure proliferation.

2.4.1. Manganese Transport and Homeostasis

Within S. cerevisiae, Mn2+ transporters Smf1 and Smf2 (part of the Nramp metal transporter family) and phosphate transporter Pho84, have a diverging consensus on their roles in Mn2+ homeostasis. In the case of Smf1, it was initially determined to be a high-affinity plasma membrane transporter, which acquired extracellular Mn2+ in Mn2+ deficient environments [65,66]. Smf2 is localized in golgi-like vesicles and shares approximately 50% identity with Smf1 (at the amino acid level), but does not share functionality and is a low-affinity Mn2+ transporter [77,257]. Once inside the cell, the fate of Mn2+ is as a cofactor for proteins such as Sod2 [126,258]. Sod2 is a mitochondrial manganese superoxide dismutase that receives Mn2+ via Mtm1 for activation [69,126,258,259]. In smf2Δ mutants, the Sod2 primary protein structure accumulates in the mitochondria; however, they were mostly inactive due to inadequate Mn2+ transfer to the mitochondria, indicating that Smf2 is a requirement for S. cerevisiae Sod2 activity [126,258]. Smf1 and Smf2, unlike many other metal ion transporters discussed in this review, are not regulated at the transcriptional level, rather post-translationally by protein turnover and localization, which is directly related to Mn2+ availability [260]. When Mn2+ concentrations are stable or in excess (~100 nmol/(1 × 109 cells)), Smf1 and Smf2 are ubiquitinated via Rsp5 (a NEDD4 family E3 ubiquitin ligase) with the aid of Bsd2 and transferrin receptor-like proteins (Tre1 and Tre2) [260,261,262]. Smf1 and Smf2 are then trafficked to multivesicular bodies, which deliver the proteins to the vacuole for degradation [260,263,264]. This mechanism of action is supported by reports that tre1Δ, tre2Δ, and bsd2Δ mutants resulted in the accumulation of Smf1 and Smf2 [260,261,262]. Conversely, when Mn2+ starvation occurs, Bsd2 is depleted, Smf1 is localized to the cell surface, Smf2 is localized to intracellular vesicles, and Smf1 and Smf2 resume their Mn2+ uptake functions [257,260,261].
The final transport system discussed here is the phosphate transporter Pho84. It was initially characterized in S. cerevisiae as a high-affinity, six-domain, transmembrane, inorganic phosphate transporter [265]. However, Pho84 is now also known as a low-affinity Mn2+ transporter, along with other metals such as cobalt, zinc, and copper [67]. Through pho84Δ mutants, it was shown that Mn2+ uptake was the most commonly affected (in relation to the other metals) when PHO84 was removed, further proving its Mn2+ transporter role [67]. PHO84 transcription is regulated by transcription factor Pho4, which inhibits Pho84 activity when it is phosphorylated in the presence of excess phosphate; Pho4 resumes transcription when phosphate levels are low [265,266].
Once Mn2+ is inside the cell, there are an array of destinations. Pmr1 (high-affinity Ca2+/Mn2+ P-type ATPase) and Gdt1 (calcium/manganese transporter) both transport cytosolic Mn2+ to the Golgi lumen, where Mn2+ serves as a cofactor for mannosyl-transferases, such as Mnn1, Mnn2, Mnn5, and Mnn9, which glycosylate proteins in the secretory pathway [70,71,72,267,268,269,270,271]. This type of protein modification provides protein stability by preventing degradation, protecting against oxidative damage, and increasing thermodynamic equilibrium [272]. Concerning the ER, P-type ATPase Spf1 transports Mn2+ to the ER lumen; this is supported by a study showing that spf1Δ mutants had decreased luminal Mn2+; its overexpression had the opposite effect [73]. This same study also stated that Mn2+ depletion observed in spf1Δ mutants negatively impacted luminal Mn2+ dependent processes. On the contrary, it positively impacted Mn2+ associated cytosolic processes, indicating that Spf1 is integral to S. cerevisiae manganese ER and cytosolic homeostasis [73].
Mn2+ accumulation can have severe consequences on cellular health, and systems must be in place to prevent subsequent events. We will discuss two defense mechanisms in S. cerevisiae, Mn2+ trafficking to vacuoles for storage and degradation and Mn2+ export. Pmr1, previously characterized as an Mn2+ Golgi lumen transporter, also serves as a detoxifier. Presented with toxic Mn2+ levels, Mn2+ is still transported to the Golgi lumen from the cytosol, but excess ions are delivered to secretory pathway vesicles, which ultimately exit the cell, completely removing toxic Mn2+ (Figure 3) [77,273]. The HIP1 gene product also expresses export activity. Hip1 was initially characterized as a high-affinity, plasma membrane histidine permease, but has since been shown to play a role in Mn2+ resistance [78,274]. Farcasanu et al. investigated S. cerevisiae mutants having defects in Mn2+transport and found that a mutation in the HIP1 gene was responsible [78]. This mutation, originally a single base deletion, introduced a cascade of mutations that led to the protein Hip1-272 (272 amino acids long). Subsequent experiments showed that hip1-272 mutants had significantly less cytosolic Mn2+ accumulation, increased Mn2+ efflux, and increased resistance than null mutants and wild type strains [78]. Further studies into the hip1-272 mutant could elucidate the exact mechanisms of action of Mn2+ transport, determining how ions are trafficked to Hip1-272 and expelled. The second defense mechanism against Mn2+ toxicity in S. cerevisiae was Mn2+ trafficking to vacuoles through Ccc1 and Ypk9. Ccc1 (and possibly Cos16) is localized in the vacuolar membrane and is responsible for trafficking cytosolic Mn2+ to vacuoles; CCC1 overexpression results in reduced Mn2+ toxicity, lower concentrations of cytosolic Mn2+, and increased vacuolar concentrations (Figure 3) [64,75,77]. Ypk9 is also localized in the vacuolar membrane and shuttles Mn2+ to the vacuole. Gitler et al. and Schmidt et al. both demonstrated that ypk9Δ mutants expressed Mn2+ hypersensitivity when compared to wild type strains, further affirming Ypk9 involvement in Mn2+ homeostasis [74,76].
Manganese homeostasis has not been well characterized in higher fungi, but Phanerochaete chrysosporium has received some attention. P. chrysosporium is a white-rot fungus that produces lignin-degrading enzymes, which have been useful in the biodegradation of various plant biomass and an array of organo-pollutants [275,276,277]. Manganese peroxidase is a common lignin depolymerizing peroxidase utilized by white-rot Basidiomycetes [278,279]. It acts in combination with other enzymes to convert various biomass to useful bio-products of commerce and agricultural operations [255,280,281,282,283]. Homologs of the S. cerevisiae Pho84 and Smf1/2 proteins have been found in P. chrysosporium, PcPho84 and PcSmfs, respectively. PcPho84 is a plasma membrane protein involved in Mn2+ uptake, having a similar function to its S. cerevisiae homolog [68]. Smf1/2 are predicted to have similar functions in P. chrysosporium to their S. cerevisiae homologs [68]. Intracellular Mn2+ transport has also been investigated. Yeast homolog PcAtx2, localized in the Golgi membrane, was shown to function as an antioxidant through sod1Δ mutants [68]. When grown on 600 µM paraquat (inducer of oxidative stress), sod1Δ mutants experienced almost no growth; however, in mutants expressing PcATX2, growth was restored, indicating that PcAtx2 exhibits similar antioxidant functionality as Sod1 [68]. In the case of mitochondrial transport, S. cerevisiae Mtm1 traffics Mn2+ to the mitochondria for Sod1 activation; however, the function of the P. chrysosporium homolog, PcMtm1 (localized in the mitochondrial membrane), has yet to be identified, but predicted to have a similar antioxidant activity [68]. PcMnt and PcCcc1 engage in Mn2+ storage and export in P. chrysosporium, respectively. In Phanerochaete sordida, PsMnt was found to be a homolog of yeast Smf2 and plays a role in Mn2+ uptake, suggesting that it could have dual functionality, but this is still unknown [77,284]. Limited information exists on Mn2+ homeostasis in other fungi; however, due to the impact of Mn2+ on lignin-degrading enzymes in wood-rotting fungi, more studies should be conducted. Overall, Mn2+ homeostasis is critical to cellular functioning to prevent toxic Mn2+ accumulation, detoxify cells of free radicals, and provide white-rot fungi with their capacity to degrade lignin. In the absence of such mechanisms, toxicity can impede proper functioning and cause cellular damage.

2.4.2. Manganese Toxicity

In the model yeast S. cerevisiae, excessive Mn2+ can overrun homeostatic systems and create a toxic ionic imbalance that negatively impacts survival rate [104,285,286]. Expression profiles show that high levels of Mn2+ down-regulate genes associated to histidine proteins (HTB2, HTA1, HTA2, HTB1, and HHF) that are compulsory in chromatin assembly chromosome functioning, and interface in this functioning can end in cell cycle arrest [94,95,114,287]. Filamentous fungi are often studied for their lignin degradation properties which focus on how Mn2+ impacts manganese peroxidase activity, but there is a lack of knowledge on how excess Mn2+ can be toxic towards this activity [96,97,98]. Due to the reliability of much of the lignin degrading properties on manganese peroxidase in many white-rot fungi, effects of Mn2+ over accumulation should be further investigated [96,97,98]. In some cases, toxicity can be avoided by resistance mechanisms.

2.4.3. Manganese Tolerance and Resistance

As with other metals, Mn2+ resistance is usually contingent upon homeostatic systems. In S. cerevisiae, several genes involved in resistance emanate from mutations. MNR1 (also known as HUM1 and VCX1), encodes a vacuolar H+/Ca2+ antiporter, but has been implicated in Mn2+ resistance [125,288,289]. A single nucleotide alteration may affect Mnr1 function and result in increased Mn2+ sequestration to the vacuole [125,289,290]. A mutation in PHO84 is also implicated in Mn2+ resistance, where pho84Δ mutants have increased resistance, likely due to the acquired inability to import and accumulate excess Mn2+ [67]. In filamentous fungi, Diss et al. elucidated potential resistance mechanisms through P. chrysosporium, where it was demonstrated that PcPho84Δ mutants increase Mn2+ resistance, as well as expression of PcMNT, which is likely to engage in Mn2+export activity [68].
Up to this point, metals that serve as essential nutrients have been reviewed. In recent years, there has been an increase in studies on the usage of metals with no nutritional purpose, but which serve as antimicrobial agents, such as silver. This increase gives cause for further investigation into how these metals are metabolized and their intracellular functioning.

2.5. Silver

Silver (Ag) is a transition metal that shares similar properties to other transition metals in groups three through twelve, and closely resembles the properties of Cu and gold (Au) [291,292]. In fungi, silver is implicated in the eradication of pathogens. As part of agricultural research, silver nanoparticles (Ag NPs) and Ag ions (Ag+) have demonstrated their ability to control plant pathogens [293,294,295]. As a feed additive, silver has a positive effect on the intestinal microflora, aflatoxins, and mycotoxin absorption in farm animals and in the food industry is used in food packaging for its antimicrobial properties [291,296,297]. Thus, the development of silver as an antimicrobial agent should continue to be investigated, especially on the development of fungal resistance and the impacts on non-target organisms.

2.5.1. Silver Transport and Homeostasis

Silver is a non-essential metal that has no designated cellular receptors or membrane channels for ion uptake. Much of the literature has focused on silver as an antimicrobial agent, but some studies have begun to clarify homeostatic mechanisms [81,114,116,298]. Silver has properties similar to copper, which has initiated the evaluation of copper homeostatic systems to investigate how they may contribute to silver uptake and transport [21,81,114,116,298].
In S. cerevisiae, Ctr1, high-affinity Cu+ transporter, has been identified as a Ag+ importer. This is based on observed reduced Ag+ uptake in ctr1Δ mutants exposed to low silver concentrations, and transcriptional analysis that shows exposure to Ag NPs upregulates CTR1 throughout the entire transcriptome [80,81]. The involvement of copper-related genes in Ag+ homeostasis was also investigated by Hosiner et al. and Niazi et al.; both found that short-term exposure to silver resulted in increased expression of copper MTs Cup1-1 and Cup1-2, suggesting these MTs sequester Ag+ in response to silver stress [114,115]. The competitiveness of Cu+ and Ag+ for Cup1-1 and Cup1-2 should be further investigated to determine which ion the MTs have a higher affinity for. Other metal ion transporters (Pho84, Fet3, and Smf1) have been investigated for their involvement in Ag+ uptake, but results indicate they are not [81].
Once inside the cell, there are not many known Ag+ destinations. AgNO3 exposure results in Ag+ accumulation in the mitochondria, which, in return, reduces Cu+ accumulation in the mitochondrial matrix [21]. The direct result of this action is reduced copper-dependent cytochrome c oxidase activity, suggesting that cytosolic Ag+ is trafficked to the mitochondria via Cu+ mitochondrial transporter Pic2, potentially with a higher affinity, which can be toxic to cells by reducing the rate of cellular respiration [21]. No other intracellular destinations have been identified in yeast, and silver homeostasis in filamentous fungi is still unknown.

2.5.2. Silver Toxicity

Efflux systems are integral to cellular homeostasis, preventing the accumulation of toxic compounds within a cell. In S. cerevisiae, Ag+ uptake can affect these systems, resulting in toxicity. Exposure to Ag+ can increase the efflux rate of potassium ions (K+) from S. cerevisiae, resulting in almost complete K+ efflux from the cell. S. cerevisiae requires a minimum 30mM K+, suggesting those events can be toxic if the ion concentration is not restored [299,300]. Another mechanism of Ag+ toxicity is its ability to alter cellular structure [100,103]. Ionic fluids can affect cell membrane integrity of yeast Yarrowia lipolytica, reducing the amount of ergosterol, which fluidizes the membrane, and increases internal lateral pressures [100]. Ag+ exposure can also deform the cell wall, which is a likely a response to the down-regulation of genes involved in ergosterol synthesis (ERG3, ERG5, ERG6, ERG11, ERG25, and ERG28) in S. cerevisiae [80,99]. In the aquatic fungus Articulospora tetracladia, transcriptome analysis via RNAseq revealed toxicity of Ag+ and Ag NPs may result from interrupted functioning of plasma/organelle membranes and downregulation of genes associated with cellular redox [301]. Silver toxicity has also been studied in other agriculturally relevant processes and it has been determined that AgNO3 and Ag NPs can be useful in pathogen control of plant diseases [174,293,295]. It may be worthwhile to investigate silver homeostasis in addressing long-term effects of exposure.

2.5.3. Silver Tolerance and Resistance

The worldwide increase of silver usage makes studies on mechanisms of silver resistance important; presently, few studies have reported on this. CTR3 is implicated in Ag+ resistance after an observed fold increase in its expression in a silver evolved strain of S. cerevisiae [116]. Insight into the expression of the Ctr3 transcription factor MAC1 in the presence of Ag+ may clarify its role in resistance. It is possible that MTs Cup1-1 and Cup1-2 are also involved in resistance. It was previously described that exposure to AgNO3 and Ag NPs resulted in the increased expression of CUP1-1 and CUP1-2, proposing that the encoded MTs may also bind Ag+ and decrease sensitivity [81,114,115]. Similar results were observed in AgNO3 exposure, where yeast had increased expression of CUP1-1 and CUP1-2 (4.79-fold and 4.71-fold, respectively) in an extended study that resulted in an evolved yeast strain, confirming the potential role of copper MTs in silver resistance [116]. Other Ag+ transporters, Pho84, Fet3, and Smf1, were not implicated in Ag+ uptake; however, significant down regulation (68.56-fold) of PHO84 in silver evolved yeast has been observed, which may indicate that Pho84 plays a role in Ag+ uptake, and may serve as a mechanism of Ag+ resistance [81,116]. The effect of Ag+ on genes involved in ergosterol biosynthesis was also investigated in a silver evolved yeast [116]. Results indicated down-regulation of those genes, suggesting that one mechanism of action of resistance against Ag+ toxicity could be the ability to inhibit their down regulation [116]. In the filamentous fungus A. nidulans, silver induced expression of copper exporter crpA, indicating that it may play a role in silver export and resistance [90]. In A. tetracladia, resistance may be due to increased vacuolar function [301]. Overall, there has been some progress made in unveiling silver homeostasis in fungi, mostly by way of S. cerevisiae. Due to the increasing silver and Ag NP usage in many aspects of human life, silver–fungal interactions should be further investigated at the molecular level to decipher precise homeostatic and resistance mechanisms.

3. Omics and Metal Homeostasis

As the potential for commercial use of antifungal metals increases, so does the need to further investigate fungal homeostasis of essential and non-essential metals. Currently, research in this area is heavily reliant on assay based methods, which can be subjective and ambiguous. In this review, many of the discoveries of homeostatic mechanisms stemmed from the use of deletion libraries, microarrays, and PCR-based methods. This can restrict the scope of the research by only analyzing known genomic or transcriptiomic signatures.
The incorporation of an omics based approach is a resolution to this issue. The most popular omics utilizes bioinformatics to analyze fungal–metal interactions at a nucleotide and protein level, which can reveal novel genes and mutations. In genomics, the entirety of a genome is assessed and compared to others for similarities and differences that can contribute to an organism’s characteristics [302,303]. Transcriptomics relies on RNA sequencing to survey gene expression through fold-changes in transcripts and proteomics assess fold-change in subsequent proteins. In fungi, omics is already incorporated into the identification of characteristics of multi-drug resistance, analysis of genomic divergence based on species origination, some analysis of metal tolerance due to short term exposure, and the analysis of the effects of exposure to non-metal selective pressures [210,301,304,305,306].
Bioinformatics analysis is used to translate omics results via computer programming methods. In nucleotide based omics, DNA or RNA is fragmented into segments or reads prior to sequencing. After sequencing, base calling assigns a nucleotide base to an intensity signal linked to a chromatogram peak and quality control measures are taken to trim reads of adapters used in the sequencing process and trim low quality bases [307]. Next, species that have a reference genome or transcriptome are mapped or aligned to that reference (resequencing). After genomic mapping, variant calling identifies distinctions between the re-sequenced organism and the reference [307]. After transcriptomic mapping, transcripts are quantified and analyzed for differential expression. Species that do not have a reference undergo de novo assembly, which constructs a genome or transcriptome from scratch. De novo assembly utilizes the fragmented reads by overlapping or matching them based on areas of similarity until the entire -ome is constructed [308]. Genome or transcriptome annotation can then be used for further interpretation of the sequencing data. In other omics, molecules produced by an organism are also analyzed and compared to chosen reference samples.
Steps within these bioinformatics pipelines require the use of computational tools written into the command line. Multiple tools with varying parameters exist to complete the same function; however, the user must decide which tools fit their scientific needs. This can result in variation between datasets and across scientific disciplines, based on accepted standards and norms. However, this limitation does not deduct from the vast amounts of data received.
With the increasing affordability of high-throughput omics, organisms can be analyzed at multiple omics levels. This is leading to a more comprehensive understanding of characteristics, especially in fungi where there is limited knowledge of their complexity. This type of research will also illuminate unique features of fungal metal homeostasis, toxicity, and resistance, especially of non-essential metals that are becoming conventional antimicrobial agents.

4. Conclusions

Fungal–metal interactions such as the synthesis of nanoparticles and metal used as antifungal agents are on the rise. Studies on metal toxicity and resistance have uncovered preserved homeostatic mechanisms. This review discussed metal homeostasis in various fungi types and has shown that essential metals have designated uptake and transport systems that regulate metal ion balance, mostly through the model organism S. cerevisiae. However, there was a significant lack of fundamental knowledge of such mechanisms in filamentous fungi, which play critical roles in nanoparticle biosynthesis and are targets of metal antifungals, further accentuating the need to investigate molecular systems involved in metal homeostasis. Fungal homeostasis of the non-essential metal silver was also highlighted. It showed that homeostatic mechanisms were reliant on existing copper transport systems, but were largely unclear regarding overall cellular processing. There is a need to further investigate other non-essential metals’ cellular homeostasis as their commercial usage increases, due to the current lack of knowledge of future implications.

Author Contributions

J.R.R.; O.S.I.; F.N.A.; all authors contributed equally to this publication. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by a Title III HBGI Grant sponsored by The US Department of Education.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data sharing not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Ahmad, A.; Mukherjee, P.; Senapati, S.; Mandal, D.; Khan, M.I.; Kumar, R.; Sastry, M. Extracellular Biosynthesis of Silver Nanoparticles Using the Fungus Fusarium oxysporum. Colloids Surf. B Biointerfaces 2003, 28, 313–318. [Google Scholar] [CrossRef]
  2. Birla, S.S.; Gaikwad, S.C.; Gade, A.K.; Rai, M.K. Rapid Synthesis of Silver Nanoparticles from Fusarium oxysporum by Optimizing Physicocultural Conditions. Sci. World J. 2013, 2013, 796018. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  3. Naureen, B.; Miana, G.A.; Shahid, K.; Asghar, M.; Tanveer, S.; Sarwar, A. Iron (III) and Zinc (II) Monodentate Schiff Base Metal Complexes: Synthesis, Characterisation and Biological Activities. J. Mol. Struct. 2021, 1231, 129946. [Google Scholar] [CrossRef]
  4. Mani Chandrika, K.V.S.; Sharma, S. Promising Antifungal Agents: A Minireview. Bioorganic Med. Chem. 2020, 28, 115398. [Google Scholar] [CrossRef] [PubMed]
  5. Varshney, R.; Bhadauria, S.; Gaur, M.S. A Review: Biological Synthesis of Silver and Copper Nanoparticles. Nano Biomed. Eng. 2012, 4, 99–106. [Google Scholar] [CrossRef] [Green Version]
  6. Gudikandula, K.; Charya Maringanti, S. Synthesis of Silver Nanoparticles by Chemical and Biological Methods and Their Antimicrobial Properties. J. Exp. Nanosci. 2016, 11, 714–721. [Google Scholar] [CrossRef]
  7. Jamdagni, P.; Khatri, P.; Rana, J.S. Green Synthesis of Zinc Oxide Nanoparticles Using Flower Extract of Nyctanthes Arbor-tristis and Their Antifungal Activity. J. King Saud Univ. Sci. 2018, 30, 168–175. [Google Scholar] [CrossRef] [Green Version]
  8. Bundschuh, M.; Filser, J.; Lüderwald, S.; McKee, M.S.; Metreveli, G.; Schaumann, G.E.; Schulz, R.; Wagner, S. Nanoparticles in the Environment: Where Do We Come from, Where Do We Go to? Environ. Sci. Eur. 2018, 30, 1–17. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  9. Kittler, S.; Greulich, C.; Diendorf, J.; Koller, M.; Epple, M. Toxicity of Silver Nanoparticles Increases during Storage because of Slow Dissolution under Release of Silver Ions. Chem. Mater. 2010, 22, 4548–4554. [Google Scholar] [CrossRef]
  10. Mussin, J.E.; Roldán, M.V.; Rojas, F.; de los Ángeles Sosa, M.; Pellegri, N.; Giusiano, G. Antifungal Activity of Silver Nanoparticles in Combination with Ketoconazole against Malassezia Furfur. AMB Express 2019, 9, 131. [Google Scholar] [CrossRef]
  11. Khan, I.; Saeed, K.; Khan, I. Nanoparticles: Properties, Applications and Toxicities. Arab. J. Chem. 2019, 12, 908–931. [Google Scholar] [CrossRef]
  12. Cheong, Y.-K.; Arce, M.P.; Benito, A.; Chen, D.; Luengo Crisóstomo, N.; Kerai, L.V.; Rodríguez, G.; Valverde, J.L.; Vadalia, M.; Cerpa-Naranjo, A. Synergistic Antifungal Study of PEGylated Graphene Oxides and Copper Nanoparticles Against Candida Albicans. Nanomaterials 2020, 10, 819. [Google Scholar] [CrossRef] [PubMed]
  13. Mulenos, M.R.; Liu, J.; Lujan, H.; Guo, B.; Lichtfouse, E.; Sharma, V.K.; Sayes, C.M. Copper, Silver, and Titania Nanoparticles Do not Release Ions under Anoxic Conditions and Release Only Minute Ion Levels Under Oxic Conditions in Water: Evidence for the Low Toxicity of Nanoparticles. Environ. Chem. Lett. 2020, 18, 1319–1328. [Google Scholar] [CrossRef]
  14. Liu, X.; Jiang, Y.; He, D.; Fang, X.; Xu, J.; Lee, Y.-W.; Keller, N.P.; Shi, J. Copper Tolerance Mediated by FgAceA and FgCrpA in Fusarium graminearum. Front. Microbiol. 2020, 11, 1392. [Google Scholar] [CrossRef] [PubMed]
  15. Zhao, H.; Eide, D. The Yeast ZRT1 Gene Encodes the Zinc Transporter Protein of a High-affinity uptake System Induced by Zinc Limitation. Proc. Natl. Acad. Sci. USA 1996, 93, 2454–2458. [Google Scholar] [CrossRef] [Green Version]
  16. Moore, R.E.; Kim, Y.; Philpott, C.C. The Mechanism of Ferrichrome Transport through Arn1p and Its Metabolism in Saccharomyces cerevisiae. Proc. Natl. Acad. Sci. USA 2003, 100, 5664–5669. [Google Scholar] [CrossRef] [Green Version]
  17. Voß, B.; Kirschhöfer, F.; Brenner-Weiß, G.; Fischer, R. Alternaria alternata Uses Two Siderophore Systems for Iron Acquisition. Sci. Rep. 2020, 10, 3587. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  18. Butt, T.R.; Sternberg, E.; Herd, J.; Crooke, S.T. Cloning and Expression of a Yeast Copper Metallothionein Gene. Gene 1984, 27, 23–33. [Google Scholar] [CrossRef]
  19. Perinelli, M.; Tegoni, M.; Freisinger, E. Different Behavior of the Histidine Residue toward Cadmium and Zinc in a Cadmium-Specific Metallothionein from an Aquatic Fungus. Inorg. Chem. 2020, 59, 16988–16997. [Google Scholar] [CrossRef]
  20. Cho, M.; Hu, G.; Caza, M.; Horianopoulos, L.C.; Kronstad, J.W.; Jung, W.H. Vacuolar Zinc Transporter Zrc1 is Required for Detoxification of Excess Intracellular Zinc in the Human Fungal Pathogen Cryptococcus neoformans. J. Microbiol. 2018, 56, 65–71. [Google Scholar] [CrossRef]
  21. Vest, K.E.; Leary, S.C.; Winge, D.R.; Cobine, P.A. Copper Import into the Mitochondrial Matrix in Saccharomyces cerevisiae is Mediated by Pic2, a Mitochondrial Carrier Family Protein. J. Biol. Chem. 2013, 288, 23884–23892. [Google Scholar] [CrossRef] [Green Version]
  22. Zhao, H.; Eide, D. The ZRT2 Gene Encodes the Low Affinity Zinc Transporter in Saccharomyces cerevisiae. J. Biol. Chem. 1996, 271, 23203–23210. [Google Scholar] [CrossRef] [Green Version]
  23. Amich, J.; Vicentefranqueira, R.; Mellado, E.; Ruiz-Carmuega, A.; Leal, F.; Calera, J.A. The ZrfC Alkaline Zinc Transporter is Required for Aspergillus fumigatus Virulence and Its Growth in the Presence of the Zn/Mn-chelating Protein Calprotectin. Cell. Microbiol. 2014, 16, 548–564. [Google Scholar] [CrossRef]
  24. Do, E.; Hu, G.; Caza, M.; Kronstad, J.W.; Jung, W.H. The ZIP Family Zinc Transporters Support the Virulence of Cryptococcus neoformans. Med. Mycol. 2016, 54, 605–615. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  25. Dos Santos, F.M.; Piffer, A.C.; Schneider, R.O.; Ribeiro, N.S.; Garcia, A.W.A.; Schrank, A.; Kmetzsch, L.; Vainstein, M.H.; Staats, C.C. Alterations of Zinc Homeostasis in Response to Cryptococcus neoformans in a Murine Macrophage Cell Line. Future Microbiol. 2017, 12, 491–504. [Google Scholar] [CrossRef]
  26. Martha-Paz, A.M.; Eide, D.; Mendoza-Cózatl, D.; Castro-Guerrero, N.A.; Aréchiga-Carvajal, E.T. Zinc Uptake in the Basidiomycota: Characterization of Zinc Transporters in Ustilago maydis. Mol. Membr. Biol. 2019, 35, 39–50. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  27. Vicentefranqueira, R.; Amich, J.; Laskaris, P.; Ibrahim-Granet, O.; Latgé, J.P.; Toledo, H.; Leal, F.; Calera, J.A. Targeting Zinc Homeostasis to combat Aspergillus fumigatus Infections. Front. Microbiol. 2015, 6, 160. [Google Scholar] [CrossRef] [PubMed]
  28. MacDiarmid, C.W.; Milanick, M.A.; Eide, D.J. Induction of the ZRC1 Metal Tolerance Gene in Zinc-limited Yeast Confers Resistance to Zinc Shock. J. Biol. Chem. 2003, 278, 15065–15072. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  29. MacDiarmid, C.; Gaither, L.; Eide, D. Zinc Transporters That Regulate Vacuolar Zinc Storage in Saccharomyces cerevisiae. EMBO J. 2000, 19, 2845–2855. [Google Scholar] [CrossRef] [Green Version]
  30. Gross, C.; Kelleher, M.; Iyer, V.R.; Brown, P.O.; Winge, D.R. Identification of the Copper Regulon in Saccharomyces cerevisiae by DNA Microarrays. J. Biol. Chem. 2000, 275, 32310–32316. [Google Scholar] [CrossRef] [Green Version]
  31. Hassett, R.; Dix, D.R.; Eide, D.J.; Kosman, D.J. The Fe(II) Permease Fet4p Functions as a Low Affinity Copper Transporter and Supports Normal Copper Trafficking in Saccharomyces cerevisiae. Biochem. J. 2000, 351 Pt 2, 477–484. [Google Scholar] [CrossRef]
  32. Pena, M.M.; Puig, S.; Thiele, D.J. Characterization of the Saccharomyces cerevisiae High Affinity Copper Transporter Ctr3. J. Biol. Chem. 2000, 275, 33244–33251. [Google Scholar] [CrossRef] [Green Version]
  33. Peña, M.M.O.; Koch, K.A.; Thiele, D.J. Dynamic Regulation of Copper Uptake and Detoxification Genes in Saccharomyces cerevisiae. Mol. Cell. Biol. 1998, 18, 2514–2523. [Google Scholar] [CrossRef] [Green Version]
  34. Beaudoin, J.; Ekici, S.; Daldal, F.; Ait-Mohand, S.; Guérin, B.; Labbé, S. Copper Transport and Regulation in Schizosaccharomyces pombe. Biochem. Soc. Trans. 2013, 41, 1679–1686. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  35. Beaudoin, J.; Ioannoni, R.; Mailloux, S.; Plante, S.; Labbé, S. Transcriptional Regulation of the Copper Transporter Mfc1 in Meiotic Cells. Eukaryot. Cell 2013, 12, 575–590. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  36. Zhou, H.; Thiele, D.J. Identification of a Novel High Affinity Copper Transport Complex in the Fission Yeast Schizosaccharomyces pombe. J. Biol. Chem. 2001, 276, 20529–20535. [Google Scholar] [CrossRef] [Green Version]
  37. Park, Y.-S.; Lian, H.; Chang, M.; Kang, C.-M.; Yun, C.-W. Identification of High-affinity Copper Transporters in Aspergillus fumigatus. Fungal Genet. Biol. 2014, 73, 29–38. [Google Scholar] [CrossRef]
  38. Peñas, M.M.; Azparren, G.; Domínguez, Á.; Sommer, H.; Ramírez, L.; Pisabarro, A.G. Identification and Functional Characterisation of Ctr1, a Pleurotus ostreatus Gene Coding for a Copper Transporter. Mol. Genet. Genom. 2005, 274, 402–409. [Google Scholar] [CrossRef] [Green Version]
  39. Raffa, N.; Osherov, N.; Keller, N. Copper Utilization, Regulation, and Acquisition by Aspergillus fumigatus. Int. J. Mol. Sci. 2019, 20, 1980. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  40. Lin, S.-J.; Pufahl, R.A.; Dancis, A.; O’Halloran, T.V.; Culotta, V.C. A Role for the Saccharomyces cerevisiae ATX1 Gene in Copper Trafficking and Iron Transport. J. Biol. Chem. 1997, 272, 9215–9220. [Google Scholar] [CrossRef] [Green Version]
  41. Yuan, D.S.; Dancis, A.; Klausner, R.D. Restriction of Copper Export in Saccharomyces cerevisiae to a Late Golgi or Post-Golgi Compartment in the Secretory Pathway. J. Biol. Chem. 1997, 272, 25787–25793. [Google Scholar] [CrossRef] [Green Version]
  42. Peter, C.; Laliberté, J.; Beaudoin, J.; Labbé, S. Copper Distributed by Atx1 is Available to Copper Amine Oxidase 1 in Schizosaccharomyces pombe. Eukaryot. Cell 2008, 7, 1781–1794. [Google Scholar] [CrossRef] [Green Version]
  43. Beers, J.; Glerum, D.M.; Tzagoloff, A. Purification, Characterization, and Localization of Yeast Cox17p, a Mitochondrial Copper Shuttle. J. Biol. Chem. 1997, 272, 33191–33196. [Google Scholar] [CrossRef] [Green Version]
  44. Glerum, D.M.; Shtanko, A.; Tzagoloff, A. Characterization of COX17, a Yeast Gene Involved in Copper Metabolism and Assembly of Cytochrome Oxidase. J. Biol. Chem. 1996, 271, 14504–14509. [Google Scholar] [CrossRef] [Green Version]
  45. Culotta, V.C.; Klomp, L.W.J.; Strain, J.; Casareno, R.L.B.; Krems, B.; Gitlin, J.D. The Copper Chaperone for Superoxide Dismutase. J. Biol. Chem. 1997, 272, 23469–23472. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  46. Laliberté, J.; Whitson, L.J.; Beaudoin, J.; Holloway, S.P.; Hart, P.J.; Labbé, S. The Schizosaccharomyces pombe Pccs Protein Functions in Both Copper Trafficking and Metal Detoxification Pathways. J. Biol. Chem. 2004, 279, 28744–28755. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  47. Carr, H.S.; George, G.N.; Winge, D.R. Yeast Cox11, a Protein Essential for Cytochrome cOxidase Assembly, Is a Cu(I)-binding Protein. J. Biol. Chem. 2002, 277, 31237–31242. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  48. Smith, A.D.; Logeman, B.L.; Thiele, D.J. Copper Acquisition and Utilization in Fungi. Annu. Rev. Microbiol. 2017, 71, 597–623. [Google Scholar] [CrossRef] [PubMed]
  49. Wiemann, P.; Perevitsky, A.; Lim, F.Y.; Shadkchan, Y.; Knox, B.P.; Landero Figueora, J.A.; Choera, T.; Niu, M.; Steinberger, A.J.; Wüthrich, M.; et al. Aspergillus fumigatus Copper Export Machinery and Reactive Oxygen Intermediate Defense Counter Host Copper-Mediated Oxidative Antimicrobial Offense. Cell Rep. 2017, 19, 1008–1021. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  50. Askwith, C.; Kaplan, J. An Oxidase-permease-based Iron Transport System in Schizosaccharomyces pombe and Its Expression in Saccharomyces cerevisiae. J. Biol. Chem. 1997, 272, 401–405. [Google Scholar] [CrossRef] [Green Version]
  51. Chen, X.-Z.; Peng, J.-B.; Cohen, A.; Nelson, H.; Nelson, N.; Hediger, M.A. Yeast SMF1 Mediates H+-coupled Iron Uptake with Concomitant Uncoupled Cation Currents. J. Biol. Chem. 1999, 274, 35089–35094. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  52. Cohen, A.; Nelson, H.; Nelson, N. The Family of SMF Metal Ion Transporters in Yeast Cells. J. Biol. Chem. 2000, 275, 33388–33394. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  53. De Silva, D.M.; Askwith, C.C.; Eide, D.; Kaplan, J. The FET3 Gene Product Required for High Affinity Iron Transport in Yeast Is a Cell Surface Ferroxidase. J. Biol. Chem. 1995, 270, 1098–1101. [Google Scholar] [PubMed] [Green Version]
  54. Dix, D.; Bridgham, J.; Broderius, M.; Eide, D. Characterization of the FET4 Protein of Yeast Evidence for a Direct Role in the Transport of Iron. J. Biol. Chem. 1997, 272, 11770–11777. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  55. Dix, D.R.; Bridgham, J.T.; Broderius, M.A.; Byersdorfer, C.A.; Eide, D.J. The FET4 Gene Encodes the Low Affinity Fe(II) Transport Protein of Saccharomyces cerevisiae. J. Biol. Chem. 1994, 269, 26092–26099. [Google Scholar] [CrossRef]
  56. Mourer, T.; Jacques, J.-F.; Brault, A.; Bisaillon, M.; Labbé, S. Shu1 Is a Cell-surface Protein Involved in Iron Acquisition from Heme in Schizosaccharomyces pombe. J. Biol. Chem. 2015, 290, 10176–10190. [Google Scholar] [CrossRef] [Green Version]
  57. Normant, V.; Mourer, T.; Labbé, S. The Major Facilitator Transporter Str3 is Required for Low-affinity Heme Acquisition in Schizosaccharomyces pombe. J. Biol. Chem. 2018, 293, 6349–6362. [Google Scholar] [CrossRef] [Green Version]
  58. Stearman, R.; Yuan, D.S.; Yamaguchi-Iwai, Y.; Klausner, R.D.; Dancis, A. A Permease-Oxidase Complex Involved in High-Affinity Iron Uptake in Yeast. Science 1996, 271, 1552–1557. [Google Scholar] [CrossRef]
  59. Pelletier, B.; Beaudoin, J.; Philpott, C.C.; Labbé, S. Fep1 Represses Expression of the Fission Yeast Schizosaccharomyces pombe Siderophore-iron Transport System. Nucleic Acids Res. 2003, 31, 4332–4344. [Google Scholar] [CrossRef] [Green Version]
  60. Eichhorn, H.; Lessing, F.; Winterberg, B.; Schirawski, J.; Kämper, J.; Müller, P.; Kahmann, R. A Ferroxidation/permeation Iron Uptake System is Required for Virulence in Ustilago maydis. Plant Cell 2006, 18, 3332–3345. [Google Scholar] [CrossRef] [Green Version]
  61. Lindahl, P.A. A Comprehensive Mechanistic Model of Iron Metabolism in Saccharomyces cerevisiae. Met. Integr. Biometal Sci. 2019, 11, 1779–1799. [Google Scholar] [CrossRef]
  62. Netz, D.; Pierik, A.; Stümpfig, M.; Muhlenhoff, U.; Lill, R. The Cfd1-Nbp35 Complex Acts as a Scaffold for Iron-sulfur Protein Assembly in the Yeast Cytosol. Nat. Chem. Biol. 2007, 3, 278–286. [Google Scholar] [CrossRef] [PubMed]
  63. Labbé, S.; Khan, M.G.M.; Jacques, J.-F. Iron Uptake and Regulation in Schizosaccharomyces pombe. Curr. Opin. Microbiol. 2013, 16, 669–676. [Google Scholar] [CrossRef] [PubMed]
  64. Li, L.; Chen, O.S.; Ward, D.M.; Kaplan, J. CCC1 is a Transporter That Mediates Vacuolar Iron Storage in Yeast. J. Biol. Chem. 2001, 276, 29515–29519. [Google Scholar] [CrossRef] [Green Version]
  65. Supek, F.; Supekova, L.; Nelson, H.; Nelson, N. A Yeast Manganese Transporter Related to the Macrophage Protein Involved in Conferring Resistance to Mycobacteria. Proc. Natl. Acad. Sci. USA 1996, 93, 5105–5110. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  66. Supek, F.; Supekova, L.; Nelson, H.; Nelson, N. Function of Metal-ion Homeostasis in the Cell Division Cycle, Mitochondrial Protein Processing, Sensitivity to Mycobacterial Infection and Brain Function. J. Exp. Biol. 1997, 200, 321–330. [Google Scholar] [PubMed]
  67. Jensen, L.T.; Ajua-Alemanji, M.; Culotta, V.C. The Saccharomyces cerevisiae High Affinity Phosphate Transporter Encoded by PHO84 also Functions in Manganese Homeostasis. J. Biol. Chem. 2003, 278, 42036–42040. [Google Scholar] [CrossRef] [Green Version]
  68. Diss, L.; Blaudez, D.; Gelhaye, E.; Chalot, M. Genome-wide Analysis of Fungal Manganese Transporters, with an Emphasis on Phanerochaete chrysosporium. Environ. Microbiol. Rep. 2011, 3, 367–382. [Google Scholar] [CrossRef]
  69. Luk, E.; Carroll, M.; Baker, M.; Culotta, V.C. Manganese Activation of Superoxide Dismutase 2 in Saccharomyces cerevisiae Requires MTM1, a Member of the Mitochondrial Carrier Family. Proc. Natl. Acad. Sci. USA 2003, 100, 10353–10357. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  70. Dulary, E.; Yu, S.-Y.; Houdou, M.; de Bettignies, G.; Decool, V.; Potelle, S.; Duvet, S.; Krzewinski-Recchi, M.-A.; Garat, A.; Matthijs, G.; et al. Investigating the Function of Gdt1p in Yeast Golgi glycosylation. Biochim. Biophys. Acta Gen. Subj. 2018, 1862, 394–402. [Google Scholar] [CrossRef]
  71. Lapinskas, P.J.; Cunningham, K.W.; Liu, X.F.; Fink, G.R.; Culotta, V.C. Mutations in PMR1 Suppress Oxidative Damage in Yeast Cells Lacking Superoxide Dismutase. Mol. Cell. Biol. 1995, 15, 1382–1388. [Google Scholar] [CrossRef] [Green Version]
  72. Thines, L.; Deschamps, A.; Sengottaiyan, P.; Savel, O.; Stribny, J.; Morsomme, P. The Yeast Protein Gdt1p Transports Mn(2+) Ions and Thereby Regulates Manganese Homeostasis in the Golgi. J. Biol. Chem. 2018, 293, 8048–8055. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  73. Cohen, Y.; Megyeri, M.; Chen, O.C.W.; Condomitti, G.; Riezman, I.; Loizides-Mangold, U.; Abdul-Sada, A.; Rimon, N.; Riezman, H.; Platt, F.M.; et al. The Yeast P5 Type ATPase, Spf1, Regulates Manganese Transport into the Endoplasmic Reticulum. PLoS ONE 2014, 8, e85519. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  74. Gitler, A.D.; Chesi, A.; Geddie, M.L.; Strathearn, K.E.; Hamamichi, S.; Hill, K.J.; Caldwell, K.A.; Caldwell, G.A.; Cooper, A.A.; Rochet, J.-C.; et al. Alpha-synuclein is Part of a Diverse and Highly Conserved Interaction Network That Includes PARK9 and Manganese Toxicity. Nat. Genet. 2009, 41, 308–315. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  75. Lapinskas, P.J.; Lin, S.-J.; Culotta, V.C. The Role of the Saccharomyces cerevisiae CCC1 Gene in the Homeostasis of Manganese Ions. Mol. Microbiol. 1996, 21, 519–528. [Google Scholar] [CrossRef] [PubMed]
  76. Schmidt, K.; Wolfe, D.M.; Stiller, B.; Pearce, D.A. Cd2+, Mn2+, Ni2+ and Se2+ Toxicity to Saccharomyces cerevisiae Lacking YPK9p the Orthologue of Human ATP13A2. Biochem. Biophys. Res. Commun. 2009, 383, 198–202. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  77. Culotta, V.C.; Yang, M.; Hall, M.D. Manganese Transport and Trafficking: Lessons Learned from Saccharomyces cerevisiae. Eukaryot. Cell 2005, 4, 1159–1165. [Google Scholar] [CrossRef] [Green Version]
  78. Farcasanu, I.; Mizunuma, M.; Hirata, D.; Miyakawa, T. Involvement of Histidine Permease (Hip1p) in Manganese Transport in Saccharomyces cerevisiae. Mol. Gen. Genetics MGG 1998, 259, 541–548. [Google Scholar] [CrossRef]
  79. Ton, V.-K.; Mandal, D.; Vahadji, C.; Rao, R. Functional Expression in Yeast of the Human Secretory Pathway Ca2+, Mn2+-ATPase Defective in Hailey-Hailey Disease. J. Biol. Chem. 2002, 277, 6422–6427. [Google Scholar] [CrossRef] [Green Version]
  80. Horstmann, C.; Campbell, C.; Kim, D.S.; Kim, K. Transcriptome Profile with 20 nm Silver Nanoparticles in Yeast. FEMS Yeast Res. 2019, 19. [Google Scholar] [CrossRef] [Green Version]
  81. Ruta, L.L.; Banu, M.A.; Neagoe, A.D.; Kissen, R.; Bones, A.M.; Farcasanu, I.C. Accumulation of Ag(I) by Saccharomyces cerevisiae Cells Expressing Plant Metallothioneins. Cells 2018, 7, 266. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  82. B, A.W. Oligodynamic Phenomena of Living Cells. Nature 1893, 48, 331. [Google Scholar] [CrossRef] [Green Version]
  83. Galván Márquez, I.; Ghiyasvand, M.; Massarsky, A.; Babu, M.; Samanfar, B.; Omidi, K.; Moon, T.W.; Smith, M.L.; Golshani, A. Zinc Oxide and Silver Nanoparticles Toxicity in the Baker’s Yeast, Saccharomyces cerevisiae. PLoS ONE 2018, 13, e0193111. [Google Scholar] [CrossRef] [PubMed]
  84. Sinisi, V.; Pelagatti, P.; Carcelli, M.; Migliori, A.; Mantovani, L.; Righi, L.; Leonardi, G.; Pietarinen, S.; Hubsch, C.; Rogolino, D. A Green Approach to Copper-Containing Pesticides: Antimicrobial and Antifungal Activity of Brochantite Supported on Lignin for the Development of Biobased Plant Protection Products. ACS Sustain. Chem. Eng. 2019, 7, 3213–3221. [Google Scholar] [CrossRef]
  85. Pasquet, J.; Chevalier, Y.; Pelletier, J.; Couval, E.; Bouvier, D.; Bolzinger, M.A. The Contribution of Zinc Ions to the Antimicrobial Activity of Zinc Oxide. Colloids Surf. A Physicochem. Eng. Asp. 2014, 457, 263–274. [Google Scholar] [CrossRef]
  86. Xue, J.; Moyer, A.; Peng, B.; Wu, J.; Hannafon, B.N.; Ding, W.-Q. Chloroquine is a Zinc Ionophore. PLoS ONE 2014, 9, e109180. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  87. 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]
  88. Lanfranco, L.; Balsamo, R.; Martino, E.; Perotto, S.; Bonfante, P. Zinc Ions Alter Morphology and Chitin Deposition in an Ericoid Fungus. Eur. J. Histochem. 2002, 46, 341–350. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  89. Noor, S.; Shah, Z.; Javed, A.; Ali, A.; Hussain, S.B.; Zafar, S.; Ali, H.; Muhammad, S.A. A Fungal Based Synthesis Method for Copper Nanoparticles with the Determination of Anticancer, Antidiabetic and Antibacterial Activities. J. Microbiol. Methods 2020, 174, 105966. [Google Scholar] [CrossRef] [PubMed]
  90. Antsotegi-Uskola, M.; Markina-Iñarrairaegui, A.; Ugalde, U. Copper Resistance in Aspergillus nidulans Relies on the PI-Type ATPase CrpA, Regulated by the Transcription Factor AceA. Front. Microbiol. 2017, 8, 912. [Google Scholar] [CrossRef]
  91. Li, L.; Ward, D.M. Iron Toxicity in Yeast: Transcriptional Regulation of the Vacuolar Iron Importer Ccc1. Curr. Genet. 2018, 64, 413–416. [Google Scholar] [CrossRef]
  92. Ward, P.P.; Conneely, O.M. Lactoferrin: Role in Iron Homeostasis and Host Defense against Microbial Infection. Biometals 2004, 17, 203–208. [Google Scholar] [CrossRef]
  93. Leal, S.M., Jr.; Roy, S.; Vareechon, C.; Carrion, S.D.; Clark, H.; Lopez-Berges, M.S.; diPietro, A.; Schrettl, M.; Beckmann, N.; Redl, B.; et al. Targeting Iron Acquisition Blocks Infection with the Fungal Pathogens Aspergillus fumigatus and Fusarium oxysporum. PLoS Pathog. 2013, 9, e1003436. [Google Scholar] [CrossRef]
  94. Dollard, C.; Ricupero-Hovasse, S.L.; Natsoulis, G.; Boeke, J.D.; Winston, F. SPT10 and SPT21 are Required for Transcription of Particular Histone Genes in Saccharomyces cerevisiae. Mol. Cell. Biol. 1994, 14, 5223–5228. [Google Scholar] [CrossRef] [Green Version]
  95. Norris, D.; Osley, M.A. The Two Gene Pairs Encoding H2A and H2B Play Different Roles in the Saccharomyces cerevisiae Life Cycle. Mol. Cell. Biol. 1987, 7, 3473–3481. [Google Scholar] [CrossRef]
  96. Janusz, G.; Kucharzyk, K.H.; Pawlik, A.; Staszczak, M.; Paszczynski, A.J. Fungal Laccase, Manganese Peroxidase and Lignin Peroxidase: Gene Expression and Regulation. Enzym. Microb. Technol. 2013, 52, 1–12. [Google Scholar] [CrossRef]
  97. Manavalan, T.; Manavalan, A.; Heese, K. Characterization of Lignocellulolytic Enzymes from White-rot Fungi. Curr. Microbiol. 2015, 70, 485–498. [Google Scholar] [CrossRef]
  98. Xu, H.; Guo, M.-Y.; Gao, Y.-H.; Bai, X.-H.; Zhou, X.-W. Expression and Characteristics of Manganese Peroxidase from Ganoderma lucidum in Pichia pastoris and Its Application in the Degradation of Four Dyes and Phenol. BMC Biotechnol. 2017, 17, 19. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  99. Das, D.; Ahmed, G. Silver Nanoparticles Damage Yeast Cell Wall. J. Biotechnol. 2012, 3, 36–39. [Google Scholar]
  100. Walker, C.; Ryu, S.; Trinh, C. Exceptional Solvent Tolerance in Yarrowia lipolytica Is Enhanced by Sterols. bioRxiv 2018, 324681. [Google Scholar] [CrossRef] [PubMed]
  101. Ghannoum, M.A.; Rice, L.B. Antifungal Agents: Mode of Action, Mechanisms of Resistance, and Correlation of These Mechanisms with Bacterial Resistance. Clin. Microbiol. Rev. 1999, 12, 501–517. [Google Scholar] [CrossRef] [Green Version]
  102. Avery, S.V.; Howlett, N.G.; Radice, S. Copper Toxicity towards Saccharomyces cerevisiae: Dependence on Plasma Membrane Fatty Acid Composition. Appl. Environ. Microbiol. 1996, 62, 3960. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  103. Babele, P.K.; Singh, A.K.; Srivastava, A. Bio-Inspired Silver Nanoparticles Impose Metabolic and Epigenetic Toxicity to Saccharomyces cerevisiae. Front. Pharmacol. 2019, 10, 1016. [Google Scholar] [CrossRef] [PubMed]
  104. Blackwell, K.J.; Tobin, J.M.; Avery, S.V. Manganese Toxicity towards Saccharomyces cerevisiae: Dependence on Intracellular and Extracellular Magnesium Concentrations. Appl. Microbiol. Biotechnol. 1998, 49, 751–757. [Google Scholar] [CrossRef] [PubMed]
  105. Gerwien, F.; Skrahina, V.; Kasper, L.; Hube, B.; Brunke, S. Metals in Fungal Virulence. FEMS Microbiol. Rev. 2017, 42. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  106. Simm, C.; Lahner, B.; Salt, D.; LeFurgey, A.; Ingram, P.; Yandell, B.; Eide, D.J. Saccharomyces cerevisiae Vacuole in Zinc Storage and Intracellular Zinc Distribution. Eukaryot. Cell 2007, 6, 1166. [Google Scholar] [CrossRef] [Green Version]
  107. Staats, C.; Kmetzsch, L.; Schrank, A.; Vainstein, M. Fungal Zinc Metabolism and Its Connections to Virulence. Front. Cell. Infect. Microbiol. 2013, 3, 65. [Google Scholar] [CrossRef] [Green Version]
  108. Wilson, D.; Citiulo, F.; Hube, B. Zinc Exploitation by Pathogenic Fungi. PLoS Pathog. 2012, 8, e1003034. [Google Scholar] [CrossRef] [Green Version]
  109. González Matute, R.; Serra, A.; Figlas, D.; Curvetto, N. Copper and Zinc Bioaccumulation and Bioavailability of Ganoderma lucidum. J. Med. Food 2011, 14, 1273–1279. [Google Scholar] [CrossRef]
  110. Ruytinx, J.; Nguyen, H.; Van Hees, M.; Op De Beeck, M.; Vangronsveld, J.; Carleer, R.; Colpaert, J.V.; Adriaensen, K. Zinc Export Results in Adaptive Zinc Tolerance in the Ectomycorrhizal Basidiomycete Suillus bovinus. Metallomics 2013, 5, 1225–1233. [Google Scholar] [CrossRef] [PubMed]
  111. Kalsotra, T.; Khullar, S.; Agnihotri, R.; Reddy, M.S. Metal Induction of Two Metallothionein Genes in the Ectomycorrhizal Fungus Suillus himalayensis and Their Role in Metal Tolerance. Microbiology 2018, 164, 868–876. [Google Scholar] [CrossRef]
  112. Tucker, S.L.; Thornton, C.R.; Tasker, K.; Jacob, C.; Giles, G.; Egan, M.; Talbot, N.J. A Fungal Metallothionein is Required for Pathogenicity of Magnaporthe grisea. Plant Cell 2004, 16, 1575–1588. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  113. Yasokawa, D.; Murata, S.; Kitagawa, E.; Iwahashi, Y.; Nakagawa, R.; Hashido, T.; Iwahashi, H. Mechanisms of Copper Toxicity in Saccharomyces cerevisiae Determined by Microarray Analysis. Environ. Toxicol. 2008, 23, 599–606. [Google Scholar] [CrossRef]
  114. Hosiner, D.; Gerber, S.; Lichtenberg-Frate, H.; Glaser, W.; Schüller, C.; Klipp, E. Impact of Acute Metal Stress in Saccharomyces cerevisiae. PLoS ONE 2014, 9, e83330. [Google Scholar] [CrossRef]
  115. 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]
  116. Terzioğlu, E.; Alkım, C.; Arslan, M.; Balaban, B.G.; Holyavkin, C.; Kısakesen, H.İ.; Topaloğlu, A.; Yılmaz Şahin, Ü.; Gündüz Işık, S.; Akman, S.; et al. Genomic, Transcriptomic and Physiological Analyses of Silver-resistant Saccharomyces cerevisiae Obtained by Evolutionary Engineering. Yeast 2020, 37, 413–426. [Google Scholar] [CrossRef]
  117. Akgul, A.; Akgul, A. Mycoremediation of Copper: Exploring the Metal Tolerance of Brown Rot Fungi. Bioresources 2018, 13, 7155–7171. [Google Scholar]
  118. Ohno, K.M.; Clausen, C.A.; Green, F.; Diehl, S.V. Insights into the Mechanism of Copper-Tolerance in Fibroporia radiculosa: The Biosynthesis of Oxalate. Int. Biodeterior. Biodegrad. 2015, 105, 90–96. [Google Scholar] [CrossRef]
  119. Tang, J.D.; Parker, L.A.; Perkins, A.D.; Sonstegard, T.S.; Schroeder, S.G.; Nicholas, D.D.; Diehl, S.V. Gene Expression Analysis of Copper Tolerance and Wood Decay in the Brown Rot Fungus Fibroporia radiculosa. Appl. Environ. Microbiol. 2013, 79, 1523–1533. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  120. Li, L.; Bagley, D.; Ward, D.M.; Kaplan, J. Yap5 Is an Iron-Responsive Transcriptional Activator That Regulates Vacuolar Iron Storage in Yeast. Mol. Cell. Biol. 2008, 28, 1326–1337. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  121. Mei, B.; Budde, A.D.; Leong, S.A. sid1, a Gene Initiating Siderophore Biosynthesis in Ustilago maydis: Molecular Characterization, Regulation by Iron, and Role in Phytopathogenicity. Proc. Natl. Acad. Sci. USA 1993, 90, 903–907. [Google Scholar] [CrossRef] [Green Version]
  122. de Llanos, R.; Martínez-Garay, C.A.; Fita-Torró, J.; Romero, A.M.; Martínez-Pastor, M.T.; Puig, S. Soybean Ferritin Expression in Saccharomyces cerevisiae Modulates Iron Accumulation and Resistance to Elevated Iron Concentrations. Appl. Environ. Microbiol. 2016, 82, 3052–3060. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  123. Kim, H.-J.; Kim, H.-M.; Kim, J.-H.; Ryu, K.-S.; Park, S.-M.; Jahng, K.-Y.; Yang, M.-S.; Kim, D.-H. Expression of Heteropolymeric Ferritin Improves Iron Storage in Saccharomyces cerevisiae. Appl. Environ. Microbiol. 2003, 69, 1999–2005. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  124. Shin, Y.-M.; Kwon, T.-H.; Kim, K.-S.; Chae, K.-S.; Kim, D.-H.; Kim, J.-H.; Yang, M.-S. Enhanced Iron Uptake of Saccharomyces cerevisiae by Heterologous Expression of a Tadpole Ferritin Gene. Appl. Environ. Microbiol. 2001, 67, 1280–1283. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  125. del Pozo, L.; Osaba, L.; Corchero, J.; Jimenez, A. A Single Nucleotide Change in the MNR1 (VCX1/HUM1) Gene Determines Resistance to Manganese in Saccharomyces cerevisiae. Yeast 1999, 15, 371–375. [Google Scholar] [CrossRef]
  126. Luk, E.E.-C.; Culotta, V.C. Manganese Superoxide Dismutase in Saccharomyces cerevisiae Acquires Its Metal Co-factor through a Pathway Involving the Nramp Metal Transporter, Smf2p. J. Biol. Chem. 2001, 276, 47556–47562. [Google Scholar] [CrossRef] [Green Version]
  127. Reza, M.H.; Shah, H.; Manjrekar, J.; Chattoo, B.B. Magnesium Uptake by CorA Transporters Is Essential for Growth, Development and Infection in the Rice Blast Fungus Magnaporthe oryzae. PLoS ONE 2016, 11, e0159244. [Google Scholar] [CrossRef] [Green Version]
  128. Mendel, R.R. Molybdenum: Biological Activity and Metabolism. Dalton Trans. 2005, 21, 3404–3409. [Google Scholar] [CrossRef]
  129. Novotny, J.A.; Peterson, C.A. Molybdenum. Adv. Nutr. 2018, 9, 272–273. [Google Scholar] [CrossRef]
  130. Clarance, P.; Luvankar, B.; Sales, J.; Khusro, A.; Agastian, P.; Tack, J.C.; Al Khulaifi, M.M.; Al-Shwaiman, H.A.; Elgorban, A.M.; Syed, A.; et al. Green Synthesis and Characterization of Gold Nanoparticles Using Endophytic Fungi Fusarium solani and Its in vitro Anticancer and Biomedical Applications. Saudi J. Biol. Sci. 2020, 27, 706–712. [Google Scholar] [CrossRef]
  131. Sharma, K.; Giri, R.; Sharma, R. Lead, Cadmium and Nickel Removal Efficiency of White-rot Fungus Phlebia brevispora. Lett. Appl. Microbiol. 2020, 71, 637–644. [Google Scholar] [CrossRef] [PubMed]
  132. Brown, K.H.; Wuehler, S.E.; Peerson, J.M. The Importance of Zinc in Human Nutrition and Estimation of the Global Prevalence of Zinc Deficiency. Food Nutr. Bull. 2001, 22, 113–125. [Google Scholar] [CrossRef] [Green Version]
  133. Feldmann, H.; Branduardi, P. Yeast: Molecular and Cell Biology; Wiley-Blackwell: Weinheim, Germany, 2012; p. 464. [Google Scholar]
  134. Zhang, C.; Huang, H.; Deng, W.; Li, T. Genome-Wide Analysis of the Zn(II)₂Cys₆ Zinc Cluster-Encoding Gene Family in Tolypocladium guangdongense and Its Light-Induced Expression. Genes 2019, 10, 179. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  135. Todd, R.B.; Andrianopoulos, A. Evolution of a Fungal Regulatory Gene Family: The Zn(II)2Cys6 Binuclear Cluster DNA Binding Motif. Fungal Genet. Biol. 1997, 21, 388–405. [Google Scholar] [CrossRef] [PubMed]
  136. Grotz, N.; Fox, T.; Connolly, E.; Park, W.; Guerinot, M.L.; Eide, D. Identification of a Family of Zinc Transporter Genes from Arabidopsis That Respond to Zinc Deficiency. Proc. Natl. Acad. Sci. USA 1998, 95, 7220–7224. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  137. Fisher, M.C.; Henk, D.A.; Briggs, C.J.; Brownstein, J.S.; Madoff, L.C.; McCraw, S.L.; Gurr, S.J. Emerging Fungal Threats to Animal, Plant and Ecosystem Health. Nature 2012, 484, 186–194. [Google Scholar] [CrossRef]
  138. Savary, S.; Ficke, A.; Aubertot, J.-N.; Hollier, C. Crop Losses Due to Diseases and Their Implications for Global Food Production Losses and Food Security. Food Secur. 2012, 4, 519–537. [Google Scholar] [CrossRef]
  139. 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]
  140. Fravel, D.; Olivain, C.; Alabouvette, C. Fusarium oxysporum and Its Biocontrol. New Phytol. 2003, 157, 493–502. [Google Scholar] [CrossRef]
  141. Sadasivam, N.; Sivaraj, R. Biogenic ZnO Nanoparticles Synthesized Using L. Aculeata Leaf Extract and Their Antifungal Activity against Plant Fungal Pathogens. Bull. Mater. Sci. 2016, 39, 1–5. [Google Scholar]
  142. Sharma, R. Pathogenecity of Aspergillus niger in Plants. Cibtech J. Microbiol. 2012, 1, 47–51. [Google Scholar]
  143. WIlliamson, B.; Tudzynski, B.; Tudzynski, P.; Van Kan, J.A.L. Botrytis cinerea: The Cause of Grey Mould Disease. Mol. Plant Pathol. 2007, 8, 561–580. [Google Scholar] [CrossRef]
  144. Yehia, R.; Ahmed, O. In vitro Study of the Antifungal Efficacy of Zinc Oxide Nanoparticles against Fusarium oxysporum and Penicillium expansum. Afr. J. Microbiol. Res. 2013, 7, 1917–1923. [Google Scholar]
  145. Bryden, W.L. Mycotoxin Contamination of the Feed Supply Chain: Implications for Animal Productivity and Feed Security. Anim. Feed Sci. Technol. 2012, 173, 134–158. [Google Scholar] [CrossRef]
  146. Flores-Flores, M.E.; Lizarraga, E.; López de Cerain, A.; González-Peñas, E. Presence of Mycotoxins in Animal Milk: A Review. Food Control 2015, 53, 163–176. [Google Scholar] [CrossRef]
  147. Appell, M.; Jackson, M.A.; Wang, L.C.; Ho, C.-H.; Mueller, A. Determination of Fusaric Acid in Maize Using Molecularly Imprinted SPE Clean-up. J. Sep. Sci. 2014, 37, 281–286. [Google Scholar] [CrossRef]
  148. Daubner, S.C.; Le, T.; Wang, S. Tyrosine Hydroxylase and Regulation of Dopamine Synthesis. Arch. Biochem. Biophys. 2011, 508, 1–12. [Google Scholar] [CrossRef] [Green Version]
  149. Rahman, M.K.; Rahman, F.; Rahman, T.; Kato, T. Dopamine-β-Hydroxylase (DBH), Its Cofactors and Other Biochemical Parameters in the Serum of Neurological Patients in Bangladesh. Int. J. Biomed. Sci. IJBS 2009, 5, 395–401. [Google Scholar] [CrossRef] [PubMed]
  150. Cuero, R.; Ouellet, T. Metal Ions Modulate Gene Expression and Accumulation of the Mycotoxins Aflatoxin and Zearalenone. J. Appl. Microbiol. 2005, 98, 598–605. [Google Scholar] [CrossRef] [PubMed]
  151. Savi, G.D.; Bortoluzzi, A.J.; Scussel, V.M. Antifungal Properties of Zinc-compounds against Toxigenic Fungi and Mycotoxin. Int. J. Food Sci. Technol. 2013, 48, 1834–1840. [Google Scholar] [CrossRef]
  152. Eide, D.J. Zinc Transporters and the Cellular Trafficking of Zinc. Biochim. Biophys. Acta Mol. Cell Res. 2006, 1763, 711–722. [Google Scholar] [CrossRef] [Green Version]
  153. Gaither, L.A.; Eide, D.J. Eukaryotic Zinc Transporters and Their Regulation. Biometals 2001, 14, 251–270. [Google Scholar] [CrossRef] [PubMed]
  154. Waters, B.M.; Eide, D.J. Combinatorial Control of Yeast FET4 Gene Expression by Iron, Zinc, and Oxygen. J. Biol. Chem. 2002, 277, 33749–33757. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  155. Eide, D.J. Multiple Regulatory Mechanisms Maintain Zinc Homeostasis in Saccharomyces cerevisiae. J. Nutr. 2003, 133, 1532S–1535S. [Google Scholar] [CrossRef] [PubMed]
  156. Zhao, H.; Butler, E.; Rodgers, J.; Spizzo, T.; Duesterhoeft, S.; Eide, D. Regulation of Zinc Homeostasis in Yeast by Binding of the ZAP1 Transcriptional Activator to Zinc-responsive Promoter Elements. J. Biol. Chem. 1998, 273, 28713–28720. [Google Scholar] [CrossRef] [Green Version]
  157. Wang, Y.; Weisenhorn, E.; MacDiarmid, C.W.; Andreini, C.; Bucci, M.; Taggart, J.; Banci, L.; Russell, J.; Coon, J.J.; Eide, D.J. The Cellular Economy of the Saccharomyces cerevisiae Zinc Proteome. Met. Integr. Biometal Sci. 2018, 10, 1755–1776. [Google Scholar] [CrossRef]
  158. Frey, A.G.; Eide, D.J. Zinc-responsive Coactivator Recruitment by the Yeast Zap1 Transcription Factor. Microbiologyopen 2012, 1, 105–114. [Google Scholar] [CrossRef]
  159. Frey, A.G.; Bird, A.J.; Evans-Galea, M.V.; Blankman, E.; Winge, D.R.; Eide, D.J. Zinc-Regulated DNA Binding of the Yeast Zap1 Zinc-Responsive Activator. PLoS ONE 2011, 6, e22535. [Google Scholar] [CrossRef] [Green Version]
  160. Guerinot, M.L. The ZIP Family of Metal Transporters. Biochim. Biophys. Acta Biomembr. 2000, 1465, 190–198. [Google Scholar] [CrossRef] [Green Version]
  161. Vicentefranqueira, R.; Moreno, M.Á.; Leal, F.; Calera, J.A. The zrfA and zrfB Genes of Aspergillus fumigatus Encode the Zinc Transporter Proteins of a Zinc Uptake System Induced in an Acid, Zinc-Depleted Environment. Eukaryot. Cell 2005, 4, 837–848. [Google Scholar] [CrossRef] [Green Version]
  162. Moreno, M.Á.; Ibrahim-Granet, O.; Vicentefranqueira, R.; Amich, J.; Ave, P.; Leal, F.; Latgé, J.-P.; Calera, J.A. The Regulation of Zinc Homeostasis by the ZafA Transcriptional Activator is Essential for Aspergillus fumigatus Virulence. Mol. Microbiol. 2007, 64, 1182–1197. [Google Scholar] [CrossRef] [PubMed]
  163. López-Berges, M.S. ZafA-Mediated Regulation of Zinc Homeostasis is Required for Virulence in the Plant Pathogen Fusarium oxysporum. Mol. Plant Pathol. 2020, 21, 244–249. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  164. Reeder, N.L.; Xu, J.; Youngquist, R.S.; Schwartz, J.R.; Rust, R.C.; Saunders, C.W. The Antifungal Mechanism of Action of Zinc Pyrithione. Br. J. Dermatol. 2011, 165, 9–12. [Google Scholar] [CrossRef] [PubMed]
  165. Park, M.; Cho, Y.-J.; Lee, Y.W.; Jung, W.H. Understanding the Mechanism of Action of the Anti-Dandruff Agent Zinc Pyrithione against Malassezia restricta. Sci. Rep. 2018, 8, 12086. [Google Scholar] [CrossRef]
  166. Stehling, O.; Lill, R. The Role of Mitochondria in Cellular Iron-sulfur Protein Biogenesis: Mechanisms, Connected Processes, and Diseases. Cold Spring Harbor Perspect. Biol. 2013, 5, a011312. [Google Scholar] [CrossRef] [Green Version]
  167. Buchman, C.; Skroch, P.; Welch, J.; Fogel, S.; Karin, M. The CUP2 Gene Product, Regulator of Yeast Metallothionein Expression, is a Copper-activated DNA-binding Protein. Mol. Cell. Biol. 1989, 9, 4091–4095. [Google Scholar] [CrossRef] [Green Version]
  168. Dancis, A.; Haile, D.; Yuan, D.S.; Klausner, R.D. The Saccharomyces cerevisiae Copper Transport Protein (Ctr1p). Biochemical Characterization, Regulation by Copper, and Physiologic Role in Copper Uptake. J. Biol. Chem. 1994, 269, 25660–25667. [Google Scholar] [CrossRef]
  169. Crawford, A.C.; Lehtovirta-Morley, L.E.; Alamir, O.; Niemiec, M.J.; Alawfi, B.; Alsarraf, M.; Skrahina, V.; Costa, A.C.B.P.; Anderson, A.; Yellagunda, S.; et al. Biphasic Zinc Compartmentalisation in a Human Fungal Pathogen. PLoS Pathog. 2018, 14, e1007013. [Google Scholar] [CrossRef] [Green Version]
  170. Khouja, H.R.; Abbà, S.; Lacercat-Didier, L.; Daghino, S.; Doillon, D.; Richaud, P.; Martino, E.; Vallino, M.; Perotto, S.; Chalot, M.; et al. OmZnT1 and OmFET, Two Metal Transporters from the Metal-tolerant Strain Zn of the Ericoid Mycorrhizal Fungus Oidiodendron maius, Confer Zinc Tolerance in Yeast. Fungal Genet. Biol. 2013, 52, 53–64. [Google Scholar] [CrossRef] [PubMed]
  171. González-Guerrero, M.; Melville, L.H.; Ferrol, N.; Lott, J.N.A.; Azcón-Aguilar, C.; Peterson, R.L. Ultrastructural Localization of Heavy Metals in the Extraradical Mycelium and Spores of the Arbuscular Mycorrhizal Fungus Glomus intraradices. Can. J. Microbiol. 2008, 54, 103–110. [Google Scholar] [CrossRef] [PubMed]
  172. Leonhardt, T.; Sácký, J.; Šimek, P.; Šantrůček, J.; Kotrba, P. Metallothionein-like Peptides Involved in Sequestration of Zn in the Zn-accumulating Ectomycorrhizal Fungus Russula atropurpurea. Metallomics 2014, 6, 1693–1701. [Google Scholar] [CrossRef] [Green Version]
  173. Sácký, J.; Leonhardt, T.; Borovička, J.; Gryndler, M.; Briksí, A.; Kotrba, P. Intracellular Sequestration of Zinc, Cadmium and Silver in Hebeloma mesophaeum and Characterization of Its Metallothionein Genes. Fungal Genet. Biol. 2014, 67, 3–14. [Google Scholar] [CrossRef]
  174. Ouda, S. 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]
  175. Zhang, P.; Zhang, D.; Zhao, X.; Wei, D.; Wang, Y.; Zhu, X. Effects of CTR4 Deletion on Virulence and Stress Response in Cryptococcus neoformans. Antonie van Leeuwenhoek 2016, 109, 1081–1090. [Google Scholar] [CrossRef]
  176. Hassett, R.; Kosman, D.J. Evidence for Cu(II) Reduction as a Component of Copper Uptake by Saccharomyces cerevisiae. J. Biol. Chem. 1995, 270, 128–134. [Google Scholar] [CrossRef] [Green Version]
  177. Zhu, Z.; Labbé, S.; Peña, M.M.O.; Thiele, D.J. Copper Differentially Regulates the Activity and Degradation of Yeast Mac1 Transcription Factor. J. Biol. Chem. 1998, 273, 1277–1280. [Google Scholar] [CrossRef] [Green Version]
  178. de Silva, D.; Davis-Kaplan, S.; Fergestad, J.; Kaplan, J. Purification and Characterization of Fet3 Protein, a Yeast Homologue of Ceruloplasmin. J. Biol. Chem. 1997, 272, 14208–14213. [Google Scholar] [CrossRef] [Green Version]
  179. Askwith, C.; Eide, D.; Van Ho, A.; Bernard, P.S.; Li, L.; Davis-Kaplan, S.; Sipe, D.M.; Kaplan, J. The FET3 Gene of S. Cerevisiae Encodes a Multicopper Oxidase Required for Ferrous Iron Uptake. Cell 1994, 76, 403–410. [Google Scholar] [CrossRef]
  180. Shakoury-Elizeh, M.; Protchenko, O.; Berger, A.; Cox, J.; Gable, K.; Dunn, T.M.; Prinz, W.A.; Bard, M.; Philpott, C.C. Metabolic Response to Iron Deficiency in Saccharomyces cerevisiae. J. Biol. Chem. 2010, 285, 14823–14833. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  181. Yamaguchi-Iwai, Y.; Dancis, A.; Klausner, R.D. AFT1: A Mediator of Iron Regulated Transcriptional Control in Saccharomyces cerevisiae. EMBO J. 1995, 14, 1231–1239. [Google Scholar] [CrossRef]
  182. Horng, Y.-C.; Cobine, P.A.; Maxfield, A.B.; Carr, H.S.; Winge, D.R. Specific Copper Transfer from the Cox17 Metallochaperone to Both Sco1 and Cox11 in the Assembly of Yeast Cytochrome c Oxidase. J. Biol. Chem. 2004, 279, 35334–35340. [Google Scholar] [CrossRef] [Green Version]
  183. Garay-Arroyo, A.; Lledías, F.; Hansberg, W.; Covarrubias, A.A. Cu,Zn-superoxide Dismutase of Saccharomyces cerevisiae is Required for Resistance to Hyperosmosis. FEBS Lett. 2003, 539, 68–72. [Google Scholar] [CrossRef] [Green Version]
  184. Zyrina, A.N.; Smirnova, E.A.; Markova, O.V.; Severin, F.F.; Knorre, D.A. Mitochondrial Superoxide Dismutase and Yap1p Act as a Signaling Module Contributing to Ethanol Tolerance of the Yeast Saccharomyces cerevisiae. Appl. Environ. Microbiol. 2017, 83, e02759–e16. [Google Scholar] [CrossRef] [Green Version]
  185. Fridovich, I. Superoxide Anion Radical (O·2), Superoxide Dismutases, and Related Matters. J. Biol. Chem. 1997, 272, 18515–18517. [Google Scholar] [CrossRef] [Green Version]
  186. Imlay, J.A.; Fridovich, I. Suppression of Oxidative Envelope Damage by Pseudoreversion of a Superoxide Dismutase-deficient Mutant of Escherichia coli. J. Bacteriol. 1992, 174, 953. [Google Scholar] [CrossRef] [Green Version]
  187. Lushchak, V.; Semchyshyn, H.; Mandryk, S.; Lushchak, O. Possible Role of Superoxide Dismutases in the Yeast Saccharomyces cerevisiae under Respiratory Conditions. Arch. Biochem. Biophys. 2005, 441, 35–40. [Google Scholar] [CrossRef]
  188. Seah, T.C.M.; Kaplan, J.G. Purification and Properties of the Catalase of Bakers’ Yeast. J. Biol. Chem. 1973, 248, 2889–2893. [Google Scholar] [CrossRef]
  189. Culotta, V.C.; Howard, W.R.; Liu, X.F. CRS5 Encodes a Metallothionein-like Protein in Saccharomyces cerevisiae. J. Biol. Chem. 1994, 269, 25295–25302. [Google Scholar] [CrossRef]
  190. George, G.N.; Byrd, J.; Winge, D.R. X-ray Absorption Studies of Yeast Copper Metallothionein. J. Biol. Chem. 1988, 263, 8199–8203. [Google Scholar] [CrossRef]
  191. Jensen, L.T.; Howard, W.R.; Strain, J.J.; Winge, D.R.; Culotta, V.C. Enhanced Effectiveness of Copper Ion Buffering by CUP1 Metallothionein Compared with CRS5 Metallothionein in Saccharomyces cerevisiae. J. Biol. Chem. 1996, 271, 18514–18519. [Google Scholar] [CrossRef] [Green Version]
  192. Nagakubo, T.; Kumano, T.; Ohta, T.; Hashimoto, Y.; Kobayashi, M. Copper Amine Oxidases Catalyze the Oxidative Deamination and Hydrolysis of Cyclic Imines. Nat. Commun. 2019, 10, 413. [Google Scholar] [CrossRef] [PubMed]
  193. Antsotegi-Uskola, M.; Markina-Iñarrairaegui, A.; Ugalde, U. New Insights into Copper Homeostasis in Filamentous Fungi. Int. Microbiol. 2020, 23, 65–73. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  194. Kusuya, Y.; Hagiwara, D.; Sakai, K.; Yaguchi, T.; Gonoi, T.; Takahashi, H. Transcription Factor Afmac1 Controls Copper Import Machinery in Aspergillus fumigatus. Curr. Genet. 2017, 63, 777–789. [Google Scholar] [CrossRef]
  195. Park, Y.-S.; Kang, S.; Seo, H.; Yun, C.-W. A copper Transcription Factor, AfMac1, Regulates Both Iron and Copper Homeostasis in the Opportunistic Fungal Pathogen Aspergillus fumigatus. Biochem. J. 2018, 475, 2831–2845. [Google Scholar] [CrossRef] [PubMed]
  196. Park, Y.-S.; Kim, T.-H.; Yun, C.-W. Functional Characterization of the Copper Transcription Factor AfMac1 from Aspergillus fumigatus. Biochem. J. 2017, 474, 2365–2378. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  197. Tang, J.D.; Perkins, A.D.; Sonstegard, T.S.; Schroeder, S.G.; Burgess, S.C.; Diehl, S.V. Short-Read Sequencing for Genomic Analysis of the Brown Rot Fungus Fibroporia radiculosa. Appl. Environ. Microbiol. 2012, 78, 2272. [Google Scholar] [CrossRef] [Green Version]
  198. Bayat, N.; Rajapakse, K.; Marinsek-Logar, R.; Drobne, D.; Cristobal, S. The Effects of Engineered Nanoparticles on the Cellular Structure and Growth of Saccharomyces cerevisiae. Nanotoxicology 2014, 8, 363–373. [Google Scholar] [CrossRef]
  199. Kasemets, K.; Käosaar, S.; Vija, H.; Fascio, U.; Mantecca, P. Toxicity of Differently Sized and Charged Silver Nanoparticles to Yeast Saccharomyces cerevisiae BY4741: A Nano-biointeraction Perspective. Nanotoxicology 2019, 13, 1041–1059. [Google Scholar] [CrossRef]
  200. Jo, W.J.; Loguinov, A.; Chang, M.; Wintz, H.; Nislow, C.; Arkin, A.P.; Giaever, G.; Vulpe, C.D. Identification of Genes Involved in the Toxic Response of Saccharomyces cerevisiae against Iron and Copper Overload by Parallel Analysis of Deletion Mutants. Toxicol. Sci. 2007, 101, 140–151. [Google Scholar] [CrossRef]
  201. Kasemets, K.; Ivask, A.; Dubourguier, H.-C.; Kahru, A. Toxicity of Nanoparticles of ZnO, CuO and TiO2 to Yeast Saccharomyces cerevisiae. Toxicol. In Vitro 2009, 23, 1116–1122. [Google Scholar] [CrossRef]
  202. Giannousi, K.; Sarafidis, G.; Mourdikoudis, S.; Pantazaki, A.; Dendrinou-Samara, C. Selective Synthesis of Cu2O and Cu/Cu2O NPs: Antifungal Activity to Yeast Saccharomyces cerevisiae and DNA Interaction. Inorg. Chem. 2014, 53, 9657–9666. [Google Scholar] [CrossRef]
  203. Hitchcock, C.A.; Pye, G.W.; Troke, P.F.; Johnson, E.M.; Warnock, D.W. Fluconazole Resistance in Candida glabrata. Antimicrob. Agents Chemother. 1993, 37, 1962–1965. [Google Scholar] [CrossRef] [Green Version]
  204. Orozco, A.S.; Higginbotham, L.M.; Hitchcock, C.A.; Parkinson, T.; Falconer, D.; Ibrahim, A.S.; Ghannoum, M.A.; Filler, S.G. Mechanism of Fluconazole Resistance in Candida krusei. Antimicrob. Agents Chemother. 1998, 42, 2645–2649. [Google Scholar] [CrossRef] [Green Version]
  205. Pfaller, M.A.; Diekema, D.J.; Gibbs, D.L.; Newell, V.A.; Nagy, E.; Dobiasova, S.; Rinaldi, M.; Barton, R.; Veselov, A. Global Antifungal Surveillance, G. Candida krusei, a Multidrug-resistant Opportunistic Fungal Pathogen: Geographic and Temporal Trends from the ARTEMIS DISK Antifungal Surveillance Program, 2001 to 2005. J. Clin. Microbiol. 2008, 46, 515–521. [Google Scholar] [CrossRef] [Green Version]
  206. Gomes da Silva Dantas, F.; Araújo de Almeida-Apolonio, A.; Pires de Araújo, R.; Regiane Vizolli Favarin, L.; Fukuda de Castilho, P.; de Oliveira Galvão, F.; Inez Estivalet Svidzinski, T.; Antônio Casagrande, G.; Mari Pires de Oliveira, K. A Promising Copper(II) Complex as Antifungal and Antibiofilm Drug against Yeast Infection. Molecules 2018, 23, 1856. [Google Scholar] [CrossRef] [Green Version]
  207. Coyle, B.; Kavanagh, K.; McCann, M.; Devereux, M.; Geraghty, M. Mode of Anti-fungal Activity of 1,10-phenanthroline and Its Cu(II), Mn(II) and Ag(I) Complexes. Biometals Int. J. Role Met. Ions Biol. Biochem. Med. 2003, 16, 321–329. [Google Scholar] [CrossRef]
  208. Borgatta, J.; Ma, C.; Hudson-Smith, N.; Elmer, W.; Plaza Pérez, C.D.; De La Torre-Roche, R.; Zuverza-Mena, N.; Haynes, C.L.; White, J.C.; Hamers, R.J. Copper Based Nanomaterials Suppress Root Fungal Disease in Watermelon (Citrullus lanatus): Role of Particle Morphology, Composition and Dissolution Behavior. ACS Sustain. Chem. Eng. 2018, 6, 14847–14856. [Google Scholar] [CrossRef]
  209. Cai, Z.; Du, W.; Zhang, Z.; Guan, L.; Zeng, Q.; Chai, Y.; Dai, C.; Lu, L. The Aspergillus fumigatus Transcription Factor AceA is Involved not only in Cu but also in Zn Detoxification through Regulating Transporters CrpA and ZrcA. Cell. Microbiol. 2018, 20, e12864. [Google Scholar] [CrossRef]
  210. Ragasa, L.R.P.; Joson, S.E.A.; Bagay, W.L.R.; Perez, T.R.; Velarde, M.C. Transcriptome Analysis Reveals Involvement of Oxidative Stress Response in a Copper-tolerant Fusarium oxysporum Strain. Fungal Biol. 2021. [Google Scholar] [CrossRef]
  211. Bolm, C. A New Iron Age. Nat. Chem. 2009, 1, 420. [Google Scholar] [CrossRef] [Green Version]
  212. Weber, K.A.; Achenbach, L.A.; Coates, J.D. Microorganisms Pumping Iron: Anaerobic Microbial Iron Oxidation and Reduction. Nat. Rev. Microbiol. 2006, 4, 752–764. [Google Scholar] [CrossRef] [Green Version]
  213. Hissen, A.H.T.; Wan, A.N.C.; Warwas, M.L.; Pinto, L.J.; Moore, M.M. The Aspergillus fumigatus Siderophore Biosynthetic Gene sidA, Encoding l-Ornithine N5-Oxygenase, Is Required for Virulence. Infect. Immun. 2005, 73, 5493–5503. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  214. Schrettl, M.; Bignell, E.; Kragl, C.; Joechl, C.; Rogers, T.; Arst, H.N., Jr.; Haynes, K.; Haas, H. Siderophore Biosynthesis but Not Reductive Iron Assimilation Is Essential for Aspergillus fumigatus Virulence. J. Exp. Med. 2004, 200, 1213–1219. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  215. López-Berges, M.S.; Capilla, J.; Turrà, D.; Schafferer, L.; Matthijs, S.; Jöchl, C.; Cornelis, P.; Guarro, J.; Haas, H.; Di Pietro, A. HapX-mediated Iron Homeostasis is Essential for Rhizosphere Competence and Virulence of the Soilborne Pathogen Fusarium oxysporum. Plant Cell 2012, 24, 3805–3822. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  216. Dong, X.; Wang, M.; Ling, N.; Shen, Q.; Guo, S. Effects of Iron and Boron Combinations on the Suppression of Fusarium Wilt in Banana. Sci. Rep. 2016, 6, 38944. [Google Scholar] [CrossRef]
  217. Hashem, A.R. Influence of Iron on the Growth of the Tomato Wilt Pathogen, Fusarium oxysporum, Isolated in Saudi Arabia. J. Plant Dis. Prot. 1995, 102, 326–330. [Google Scholar]
  218. Philpott, C.C. Iron Uptake in Fungi: A System for Every Source. Biochim. Biophys. Acta Mol. Cell Res. 2006, 1763, 636–645. [Google Scholar] [CrossRef] [Green Version]
  219. Yun, C.-W.; Ferea, T.; Rashford, J.; Ardon, O.; Brown, P.O.; Botstein, D.; Kaplan, J.; Philpott, C.C. Desferrioxamine-mediated Iron Uptake in Saccharomyces cerevisiae: Evidence for Two Pathways of Iron Uptake. J. Biol. Chem. 2000, 275, 10709–10715. [Google Scholar] [CrossRef] [Green Version]
  220. Yun, C.-W.; Tiedeman, J.S.; Moore, R.E.; Philpott, C.C. Siderophore-Iron Uptake in Saccharomyces cerevisiae: Identification of Ferrichrome and Fusarinine Transporters J. Biol. Chem. 2000, 275, 16354–16359. [Google Scholar] [CrossRef] [Green Version]
  221. Georgatsou, E.; Alexandraki, D. Two distinctly Regulated Genes are Required for Ferric Reduction, the First Step of Iron Uptake in Saccharomyces cerevisiae. Mol. Cell. Biol. 1994, 14, 3065–3073. [Google Scholar] [CrossRef] [Green Version]
  222. Martínez-Pastor, M.T.; Perea-García, A.; Puig, S. Mechanisms of Iron Sensing and Regulation in the Yeast Saccharomyces cerevisiae. World J. Microbiol. Biotechnol. 2017, 33, 75. [Google Scholar] [CrossRef] [Green Version]
  223. Jensen, L.T.; Culotta, V.C. Regulation of Saccharomyces cerevisiae FET4 by Oxygen and Iron. J. Mol. Biol. 2002, 318, 251–260. [Google Scholar] [CrossRef]
  224. Lesuisse, E.; Raguzzi, F.; Crichton, R. Iron Uptake by the Yeast Saccharomyces cerevisiae: Involvement of a Reduction Step. J. Gen. Microbiol. 1987, 133, 3229–3236. [Google Scholar] [CrossRef] [Green Version]
  225. Georgatsou, E.; Mavrogiannis, L.A.; Fragiadakis, G.S.; Alexandraki, D. The Yeast Fre1p/Fre2p Cupric Reductases Facilitate Copper Uptake and Are Regulated by the Copper-modulated Mac1p Activator. J. Biol. Chem. 1997, 272, 13786–13792. [Google Scholar] [CrossRef] [Green Version]
  226. Caetano, S.M.; Menezes, R.; Amaral, C.; Rodrigues-Pousada, C.; Pimentel, C. Repression of the Low Affinity Iron Transporter Gene FET4: A Novel Mechanism against Cadmium Toxicity Orchestrated by Yap1 via Rox1. J. Biol. Chem. 2015, 290, 18584–18595. [Google Scholar] [CrossRef] [Green Version]
  227. Liu, X.F.; Supek, F.; Nelson, N.; Culotta, V.C. Negative Control of Heavy Metal Uptake by the Saccharomyces cerevisiae BSD2 Gene. J. Biol. Chem. 1997, 272, 11763–11769. [Google Scholar] [CrossRef] [Green Version]
  228. Kim, Y.; Lampert, S.M.; Philpott, C.C. A Receptor Domain Controls the Intracellular Sorting of the Ferrichrome Transporter, ARN1. EMBO J. 2005, 24, 952–962. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  229. Kim, Y.; Yun, C.-W.; Philpott, C.C. Ferrichrome Induces Endosome to Plasma Membrane Cycling of the Ferrichrome Transporter, Arn1p, in Saccharomyces cerevisiae. EMBO J. 2002, 21, 3632–3642. [Google Scholar] [CrossRef] [Green Version]
  230. Heymann, P.; Ernst, J.F.; Winkelmann, G. Identification of a Fungal Triacetylfusarinine C Siderophore Transport Gene (TAF1) in Saccharomyces cerevisiae as a Member of the Major Facilitator Superfamily. Biometals 1999, 12, 301–306. [Google Scholar] [CrossRef]
  231. Lesuisse, E.; Blaiseau, P.-L.; Dancis, A.; Camadro, J.-M. Siderophore Uptake and Use by the Yeast Saccharomyces cerevisiae. Microbiology 2001, 147, 289–298. [Google Scholar] [CrossRef] [Green Version]
  232. Froissard, M.; Belgareh-Touzé, N.; Dias, M.; Buisson, N.; Camadro, J.M.; Haguenauer-Tsapis, R.; Lesuisse, E. Trafficking of Siderophore Transporters in Saccharomyces cerevisiae and Intracellular Fate of Ferrioxamine B Conjugates. Traffic 2007, 8, 1601–1616. [Google Scholar] [CrossRef]
  233. Kosman, D.J. Molecular Mechanisms of Iron Uptake in Fungi. Mol. Microbiol. 2003, 47, 1185–1197. [Google Scholar] [CrossRef]
  234. Kohlhaw, G.B. Leucine Biosynthesis in Fungi: Entering Metabolism through the Back Door. Microbiol. Mol. Biol. Rev. 2003, 67, 1–15. [Google Scholar] [CrossRef] [Green Version]
  235. Pokharel, S.; Campbell, J.L. Cross Talk between the Nuclease and Helicase Activities of Dna2: Role of an Essential Iron-sulfur Cluster Domain. Nucleic Acids Res. 2012, 40, 7821–7830. [Google Scholar] [CrossRef] [Green Version]
  236. Labbé, S.; Pelletier, B.; Mercier, A. Iron Homeostasis in the Fission Yeast Schizosaccharomyces pombe. Biometals 2007, 20, 523–537. [Google Scholar] [CrossRef]
  237. Schrettl, M.; Winkelmann, G.; Haas, H. Ferrichrome in Schizosaccharomyces pombe–an Iron Transport and Iron Storage Compound. Biometals 2004, 17, 647–654. [Google Scholar] [CrossRef]
  238. Roman, D.G.; Dancis, A.; Anderson, G.J.; Klausner, R.D. The Fission Yeast Ferric Reductase Gene frp1+ is Required for Ferric Iron Uptake and Encodes a Protein That is Homologous to the gp91-phox Subunit of the Human NADPH Phagocyte Oxidoreductase. Mol. Cell. Biol. 1993, 13, 4342–4350. [Google Scholar] [CrossRef] [Green Version]
  239. Pouliot, B.; Jbel, M.; Mercier, A.; Labbé, S. abc3 + Encodes an Iron-Regulated Vacuolar ABC-Type Transporter in Schizosaccharomyces pombe. Eukaryot. Cell 2010, 9, 59–73. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  240. Lowry, C.V.; Zitomer, R.S. ROX1 Encodes a Heme-induced Repression Factor Regulating ANB1 and CYC7 of Saccharomyces cerevisiae. Mol. Cell. Biol. 1988, 8, 4651–4658. [Google Scholar] [CrossRef] [Green Version]
  241. Hickman, M.J.; Winston, F. Heme Levels Switch the Function of Hap1 of Saccharomyces cerevisiae between Transcriptional Activator and Transcriptional Repressor. Mol. Cell. Biol. 2007, 27, 7414–7424. [Google Scholar] [CrossRef] [Green Version]
  242. Santos, R.; Buisson, N.; Knight, S.; Dancis, A.; Camadro, J.-M.; Lesuisse, E. Haemin Uptake and Use as an Iron Source by Candida albicans: Role of CaHMX1-encoded Haem Oxygenase. Microbiology 2003, 149, 579–588. [Google Scholar] [CrossRef]
  243. Tsiftsoglou, A.S.; Tsamadou, A.I.; Papadopoulou, L.C. Heme as Key Regulator of Major Mammalian Cellular Functions: Molecular, Cellular, and Pharmacological Aspects. Pharmacol. Ther. 2006, 111, 327–345. [Google Scholar] [CrossRef]
  244. Mercier, A.; Pelletier, B.; Labbé, S. A Transcription Factor Cascade Involving Fep1 and the CCAAT-Binding Factor Php4 Regulates Gene Expression in Response to Iron Deficiency in the Fission Yeast Schizosaccharomyces pombe. Eukaryot. Cell 2006, 5, 1866–1881. [Google Scholar] [CrossRef] [Green Version]
  245. Weissman, Z.; Shemer, R.; Conibear, E.; Kornitzer, D. An Endocytic Mechanism for Haemoglobin-iron Acquisition in Candida albicans. Mol. Microbiol. 2008, 69, 201–217. [Google Scholar] [CrossRef]
  246. Ardon, O.; Nudelman, R.; Caris, C.; Libman, J.; Shanzer, A.; Chen, Y.; Hadar, Y. Iron Uptake in Ustilago maydis: Tracking the Iron Path. J. Bacteriol. 1998, 180, 2021–2026. [Google Scholar] [CrossRef] [Green Version]
  247. Wang, J.; Budde, A.D.; Leong, S.A. Analysis of Ferrichrome Biosynthesis in the Phytopathogenic Fungus Ustilago maydis: Cloning of an Ornithine-N5-oxygenase Gene. J. Bacteriol. 1989, 171, 2811–2818. [Google Scholar] [CrossRef] [Green Version]
  248. Lin, H.; Li, L.; Jia, X.; Ward, D.M.; Kaplan, J. Genetic and Biochemical Analysis of High Iron Toxicity in Yeast: Iron Toxicity is Due to the Accumulation of Cytosolic Iron and Occurs under Both Aerobic and Anaerobic Conditions. J. Biol. Chem. 2011, 286, 3851–3862. [Google Scholar] [CrossRef] [Green Version]
  249. Berthelet, S.; Usher, J.; Shulist, K.; Hamza, A.; Maltez, N.; Johnston, A.; Fong, Y.; Harris, L.J.; Baetz, K. Functional Genomics Analysis of the Saccharomyces cerevisiae Iron Responsive Transcription Factor Aft1 Reveals Iron-independent Functions. Genetics 2010, 185, 1111–1128. [Google Scholar] [CrossRef] [Green Version]
  250. van Bakel, H.; Strengman, E.; Wijmenga, C.; Holstege, F.C.P. Gene Expression Profiling and Phenotype Analyses of S. Cerevisiae in Response to Changing Copper Reveals Six Genes with New Roles in Copper and Iron Metabolism. Physiol. Genom. 2005, 22, 356–367. [Google Scholar] [CrossRef]
  251. Adamczyk, Z.; Oćwieja, M.; Mrowiec, H.; Walas, S.; Lupa, D. Oxidative Dissolution of Silver Nanoparticles: A New Theoretical Approach. J. Colloid Interface Sci. 2016, 469, 355–364. [Google Scholar] [CrossRef] [PubMed]
  252. Hassan, N.; Elsharkawy, M.M.; Shimizu, M.; Hyakumachi, M. Control of Root Rot and Wilt Diseases of Roselle under Field Conditions. Mycobiology 2014, 42, 376–384. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  253. Adhilakshmi, M.; Karthikeyan, M.; Devadason, A. Effect of Combination of Bio-agents and Mineral Nutrients for the Management of Alfalfa Wilt Pathogen Fusarium oxysporum f. sp. medicaginis. Arch. Phytopathol. Plant Prot. 2008, 47, 514–525. [Google Scholar] [CrossRef]
  254. Clark, H.L.; Jhingran, A.; Sun, Y.; Vareechon, C.; de Jesus Carrion, S.; Skaar, E.P.; Chazin, W.J.; Calera, J.A.; Hohl, T.M.; Pearlman, E. Zinc and Manganese Chelation by Neutrophil S100A8/A9 (Calprotectin) Limits Extracellular Aspergillus fumigatus Hyphal Growth and Corneal Infection. J. Immunol. 2016, 196, 336–344. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  255. Isikhuemhen, O.S.; Mikiashvilli, N.A. Lignocellulolytic Enzyme Activity, Substrate Utilization, and Mushroom Yield by Pleurotus ostreatus Cultivated on Substrate Containing Anaerobic Digester Solids. J. Ind. Microbiol. Biotechnol. 2009, 36, 1353–1362. [Google Scholar] [CrossRef] [PubMed]
  256. Isikhuemhen, O.S.; Nerud, F. Preliminary Studies on the Ligninolytic Enzymes Produced by the Tropical Fungus Pleurotus tuber-regium (Fr.) Sing. Antonie van Leeuwenhoek 1999, 75, 257–260. [Google Scholar] [CrossRef]
  257. Portnoy, M.E.; Liu, X.F.; Culotta, V.C. Saccharomyces cerevisiae Expresses Three Functionally Distinct Homologues of the Nramp Family of Metal Transporters. Mol. Cell. Biol. 2000, 20, 7893–7902. [Google Scholar] [CrossRef]
  258. Luk, E.; Yang, M.; Jensen, L.T.; Bourbonnais, Y.; Culotta, V.C. Manganese Activation of Superoxide Dismutase 2 in the Mitochondria of Saccharomyces cerevisiae. J. Biol. Chem. 2005, 280, 22715–22720. [Google Scholar] [CrossRef] [Green Version]
  259. Van Loon, A.; Pesold-Hurt, B.; Schatz, G. A Yeast Mutant Lacking Mitochondrial Manganese-superoxide Dismutase is Hypersensitive to Oxygen. Proc. Natl. Acad. Sci. USA 1986, 83, 3820–3824. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  260. Reddi, A.R.; Jensen, L.T.; Naranuntarat, A.; Rosenfeld, L.; Leung, E.; Shah, R.; Culotta, V.C. The Overlapping Roles of Manganese and Cu/Zn SOD in Oxidative Stress Protection. Free Radic. Biol. Med. 2009, 46, 154–162. [Google Scholar] [CrossRef] [Green Version]
  261. Liu, X.F.; Culotta, V.C. Post-translation Control of Nramp Metal Transport in Yeast: Role of Metal Ions and the BSD2 Gene. J. Biol. Chem. 1999, 274, 4863–4868. [Google Scholar] [CrossRef] [Green Version]
  262. Stimpson, H.E.; Lewis, M.J.; Pelham, H.R. Transferrin Receptor-like Proteins Control the Degradation of a Yeast Metal Transporter. EMBO J. 2006, 25, 662–672. [Google Scholar] [CrossRef]
  263. Jensen, L.T.; Carroll, M.C.; Hall, M.D.; Harvey, C.J.; Beese, S.E.; Culotta, V.C. Down-regulation of a Manganese Transporter in the Face of Metal Toxicity. Mol. Biol. Cell 2009, 20, 2810–2819. [Google Scholar] [CrossRef] [Green Version]
  264. Yang, B.; Kumar, S. Nedd4 and Nedd4-2: Closely Related Ubiquitin-protein Ligases with Distinct Physiological Functions. Cell Death Differ. 2010, 17, 68–77. [Google Scholar] [CrossRef]
  265. Bun-Ya, M.; Nishimura, M.; Harashima, S.; Oshima, Y. The PHO84 Gene of Saccharomyces cerevisiae Encodes an Inorganic Phosphate Transporter. Mol. Cell. Biol. 1991, 11, 3229–3238. [Google Scholar] [CrossRef]
  266. Wykoff, D.D.; Rizvi, A.H.; Raser, J.M.; Margolin, B.; O’Shea, E.K. Positive Feedback Regulates Switching of Phosphate Transporters in S. cerevisiae. Mol. Cell 2007, 27, 1005–1013. [Google Scholar] [CrossRef] [Green Version]
  267. Reddi, A.R.; Jensen, L.T.; Culotta, V.C. Manganese Homeostasis in Saccharomyces cerevisiae. Chem. Rev. 2009, 109, 4722–4732. [Google Scholar] [CrossRef] [Green Version]
  268. Rayner, J.C.; Munro, S. Identification of the MNN2 and MNN5Mannosyltransferases Required for Forming and Extending the Mannose Branches of the Outer Chain Mannans of Saccharomyces cerevisiae. J. Biol. Chem. 1998, 273, 26836–26843. [Google Scholar] [CrossRef] [Green Version]
  269. Romero, P.A.; Herscovics, A. Glycoprotein Biosynthesis in Saccharomyces cerevisiae. Characterization of Alpha-1,6-mannosyltransferase Which Initiates outer Chain Formation. J. Biol. Chem. 1989, 264, 1946–1950. [Google Scholar] [CrossRef]
  270. Wiggins, C.A.; Munro, S. Activity of the Yeast MNN1 Alpha-1,3-mannosyltransferase Requires a Motif Conserved in Many Other Families of Glycosyltransferases. Proc. Natl. Acad. Sci. USA 1998, 95, 7945–7950. [Google Scholar] [CrossRef] [Green Version]
  271. Striebeck, A.; Robinson, D.A.; Schüttelkopf, A.W.; van Aalten, D.M.F. Yeast Mnn9 is Both a Priming Glycosyltransferase and an Allosteric Activator of Mannan Biosynthesis. Open Biol. 2013, 3, 130022. [Google Scholar] [CrossRef]
  272. Solá, R.J.; Griebenow, K. Effects of Glycosylation on the Stability of Protein Pharmaceuticals. J. Pharm. Sci. 2009, 98, 1223–1245. [Google Scholar] [CrossRef] [Green Version]
  273. Xiang, M.; Mohamalawari, D.; Rao, R. A Novel Isoform of the Secretory Pathway Ca2+,Mn(2+)-ATPase, hSPCA2, Has Unusual Properties and is Expressed in the Brain. J. Biol. Chem. 2005, 280, 11608–11614. [Google Scholar] [CrossRef] [Green Version]
  274. Tanaka, J.; Fink, G.R. The Histidine Permease Gene (HIP1) of Saccharomyces cerevisiae. Gene 1985, 38, 205–214. [Google Scholar] [CrossRef]
  275. Bakshi, D.K.; Saha, S.; Sindhu, I.; Sharma, P. Use of Phanerochaete Chrysosporium Biomass for the Removal of Textile Dyes from a Synthetic Effluent. World J. Microbiol. Biotechnol. 2006, 22, 835–839. [Google Scholar] [CrossRef]
  276. Bending, G.D.; Friloux, M.; Walker, A. Degradation of Contrasting Pesticides by White Rot Fungi and Its Relationship with Ligninolytic Potential. FEMS Microbiol. Lett. 2002, 212, 59–63. [Google Scholar] [CrossRef]
  277. Kirk, T.K.; Connors, W.; Zeikus, J.G. Requirement for a Growth Substrate during Lignin Decomposition by Two Wood-rotting Fungi. Appl. Environ. Microbiol. 1976, 32, 192–194. [Google Scholar] [CrossRef] [Green Version]
  278. Kuwahara, M.; Glenn, J.K.; Morgan, M.A.; Gold, M.H. Separation and Characterization of Two Extracelluar H2O2-dependent Oxidases from Ligninolytic Cultures of Phanerochaete chrysosporium. FEBS Lett. 1984, 169, 247–250. [Google Scholar] [CrossRef] [Green Version]
  279. Bonnarme, P.; Jeffries, T.W. Mn(II) Regulation of Lignin Peroxidases and Manganese-dependent Peroxidases from Lignin-degrading White Rot Fungi. Appl. Environ. Microbiol. 1990, 56, 210–217. [Google Scholar] [CrossRef] [Green Version]
  280. Isikhuemhen, O.S.; Mikiashvili, N.A.; Kelkar, V. Application of Solid Waste from Anaerobic Digestion of Poultry Litter in Agrocybe aegerita Cultivation: Mushroom Production, Lignocellulolytic Enzymes Activity and Substrate Utilization. Biodegradation 2009, 20, 351–361. [Google Scholar] [CrossRef]
  281. Isikhuemhen, O.S.; Mikiashvili, N.A.; Adenipekun, C.O.; Ohimain, E.I.; Shahbazi, G. The Tropical White Rot Fungus, Lentinus squarrosulus Mont.: Lignocellulolytic Enzymes Activities and Sugar Release from Cornstalks under Solid State Fermentation. World J. Microbiol. Biotechnol. 2012, 28, 1961–1966. [Google Scholar] [CrossRef]
  282. Isikhuemhen, O.S.; Mikiashvili, N.A.; Senwo, Z.N.; Ohimain, E.I. Biodegradation and Sugar Release from Canola Plant Biomass by Selected White Rot Fungi. Adv. Biol. Chem. 2014, 4, 395. [Google Scholar] [CrossRef] [Green Version]
  283. Zebulun, H.O.; Isikhuemhen, O.S.; Inyang, H. Decontamination of Anthracene-polluted Soil through White Rot Fungus-induced Biodegradation. Environmentalist 2011, 31, 11–19. [Google Scholar] [CrossRef]
  284. Mori, T.; Nagai, Y.; Kawagishi, H.; Hirai, H. Functional Characterization of the Manganese Transporter smf2 Homologue Gene, PsMnt, of Phanerochaete sordida YK-624 via Homologous Overexpression. FEMS Microbiol. Lett. 2018, 365. [Google Scholar] [CrossRef] [PubMed]
  285. Liang, Q.; Zhou, B. Copper and Manganese Induce Yeast Apoptosis via Different Pathways. Mol. Biol. Cell 2007, 18, 4741–4749. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  286. Yang, M.; Jensen, L.T.; Gardner, A.J.; Culotta, V.C. Manganese Toxicity and Saccharomyces cerevisiae Mam3p, a Member of the ACDP (Ancient Conserved Domain Protein) Family. Biochem. J. 2005, 386, 479–487. [Google Scholar] [CrossRef] [Green Version]
  287. Hereford, L.; Fahrner, K.; Woolford, J., Jr.; Rosbash, M.; Kaback, D.B. Isolation of Yeast Histone Genes H2A and H2B. Cell 1979, 18, 1261–1271. [Google Scholar] [CrossRef]
  288. Cunningham, K.W.; Fink, G.R. Calcineurin Inhibits VCX1-dependent H+/Ca2+ Exchange and Induces Ca2+ ATPases in Saccharomyces cerevisiae. Mol. Cell. Biol. 1996, 16, 2226–2237. [Google Scholar] [CrossRef] [Green Version]
  289. Pozos, T.C.; Sekler, I.; Cyert, M.S. The Product of HUM1, a Novel Yeast Gene, is Required for Vacuolar Ca2+/H+ Exchange and is Related to Mammalian Na+/Ca2+ Exchangers. Mol. Cell. Biol. 1996, 16, 3730–3741. [Google Scholar] [CrossRef] [Green Version]
  290. Bianchi, M.; Carbone, M.; Lucchini, G.; Magni, G. Mutants Resistant to Manganese in Saccharomyces cerevisiae. Curr. Genet. 1981, 4, 215–220. [Google Scholar] [CrossRef]
  291. Peters, R.J.B.; Bouwmeester, H.; Gottardo, S.; Amenta, V.; Arena, M.; Brandhoff, P.; Marvin, H.J.P.; Mech, A.; Moniz, F.B.; Pesudo, L.Q.; et al. Nanomaterials for Products and Application in Agriculture, Feed and Food. Trends Food Sci. Technol. 2016, 54, 155–164. [Google Scholar] [CrossRef]
  292. Wesley, A. History of the Medical Use of Silver. Surg. Infect. 2009, 10, 289–292. [Google Scholar]
  293. 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] [Green Version]
  294. Kaur, P.; Thakur, R.; Duhan, J.S.; Chaudhury, A. Management of Wilt Disease of Chickpea in vivo by Silver Nanoparticles Biosynthesized by Rhizospheric Microflora of Chickpea (Cicer arietinum). J. Chem. Technol. Biotechnol. 2018, 93, 3233–3243. [Google Scholar] [CrossRef]
  295. Farooq, M.; Ilyas, N.; Khan, I.; Saboor, A.; Khan, K.; Khan, M.N.; Qayum, A.; Ilyas, N.; Bakhtia, M. Antifungal Activity of Plant Extracts and Silver Nano Particles Against Citrus Brown Spot Pathogen (Alternaria citri). Int. J. Environ. Agric. Res. 2018, 4, 118–125. [Google Scholar]
  296. Gholami-Ahangaran, M.; Zia-Jahromi, N. Nanosilver Effects on Growth Parameters in Experimental Aflatoxicosis in Broiler Chickens. Toxicol. Ind. Health 2013, 29, 121–125. [Google Scholar] [CrossRef]
  297. Carbone, M.; Donia, D.T.; Sabbatella, G.; Antiochia, R. Silver Nanoparticles in Polymeric Matrices for Fresh Food Packaging. J. King Saud Univ. Sci. 2016, 28, 273–279. [Google Scholar] [CrossRef] [Green Version]
  298. Vest, K.E.; Wang, J.; Gammon, M.G.; Maynard, M.K.; White, O.L.; Cobine, J.A.; Mahone, W.K.; Cobine, P.A. Overlap of Copper and Iron Uptake Systems in Mitochondria in Saccharomyces cerevisiae. Open Biol. 2016, 6, 150223. [Google Scholar] [CrossRef] [Green Version]
  299. Vagabov, V.; Yu Ivanov, A.; Kulakovskaya, T.; Kulakovskaya, E.; Petrov, V.; Kulaev, I. Efflux of Potassium Ions from Cells and Spheroplasts of Saccharomyces cerevisiae Yeast Treated with Silver and Copper Ions. Biochemistry 2008, 73, 1224–1227. [Google Scholar] [CrossRef]
  300. Cyert, M.S.; Philpott, C.C. Regulation of Cation Balance in Saccharomyces cerevisiae. Genetics 2013, 193, 677–713. [Google Scholar] [CrossRef] [Green Version]
  301. Barros, D.; Pradhan, A.; Pascoal, C.; Cássio, F. Transcriptomics Reveals the Action Mechanisms and Cellular Targets of Citrate-coated Silver Nanoparticles in a Ubiquitous Aquatic Fungus. Environ. Pollut. 2021, 268, 115913. [Google Scholar] [CrossRef]
  302. Ball, B.; Langille, M.; Geddes-McAlister, J. Fun(gi)omics: Advanced and Diverse Technologies to Explore Emerging Fungal Pathogens and Define Mechanisms of Antifungal Resistance. Mbio 2020, 11, e01020–e20. [Google Scholar] [CrossRef]
  303. Chitarrini, G.; Riccadonna, S.; Zulini, L.; Vecchione, A.; Stefanini, M.; Larger, S.; Pindo, M.; Cestaro, A.; Franceschi, P.; Magris, G.; et al. Two-omics Data Revealed Commonalities and Differences between Rpv12- and Rpv3-mediated Resistance in Grapevine. Sci. Rep. 2020, 10, 12193. [Google Scholar] [CrossRef] [PubMed]
  304. Bazzicalupo, A.L.; Ruytinx, J.; Ke, Y.-H.; Coninx, L.; Colpaert, J.V.; Nguyen, N.H.; Vilgalys, R.; Branco, S. Fungal Heavy Metal Adaptation through Single Nucleotide Polymorphisms and Copy-number Variation. Mol. Ecol. 2020, 29, 4157–4169. [Google Scholar] [CrossRef] [PubMed]
  305. Libkind, D.; Peris, D.; Cubillos, F.A.; Steenwyk, J.L.; Opulente, D.A.; Langdon, Q.K.; Rokas, A.; Hittinger, C.T. Into the Wild: New Yeast Genomes from Natural Environments and New Tools for Their Analysis. FEMS Yeast Res. 2020, 20. [Google Scholar] [CrossRef]
  306. Samaras, A.; Ntasiou, P.; Myresiotis, C.; Karaoglanidis, G. Multidrug Resistance of Penicillium expansum to Fungicides: Whole Transcriptome Analysis of MDR Strains Reveals Overexpression of Efflux Transporter Genes. Int. J. Food Microbiol. 2020, 335, 108896. [Google Scholar] [CrossRef] [PubMed]
  307. Pereira, R.; Oliveira, J.; Sousa, M. Bioinformatics and Computational Tools for Next-Generation Sequencing Analysis in Clinical Genetics. J. Clin. Med. 2020, 9, 132. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  308. Liu, P.-H.; Huang, Z.-X.; Luo, X.-H.; Chen, H.; Weng, B.-Q.; Wang, Y.-X.; Chen, L.-S. Comparative Transcriptome Analysis Reveals Candidate Genes Related to Cadmium Accumulation and Tolerance in Two Almond Mushroom (Agaricus brasiliensis) Strains with Contrasting Cadmium Tolerance. PLoS ONE 2020, 15, e0239617. [Google Scholar]
Figure 1. S. cerevisiae zinc homeostatic systems.
Figure 1. S. cerevisiae zinc homeostatic systems.
Jof 07 00225 g001
Figure 2. Yeast copper transport systems. In S. cerevisiae, cupric reductase, Fre1 reduces extracellular cupric oxide for transport across high and low-affinity copper membrane transports Ctr1 and Fet4. From the cytoplasm, Ccc2 shuttles Cu+ to Golgi bodies, and Pic2 shuttles Cu+ to the mitochondrial matrix. During meiosis in S. pombe, Mfc1 transports Cu+ across the forespore membrane.
Figure 2. Yeast copper transport systems. In S. cerevisiae, cupric reductase, Fre1 reduces extracellular cupric oxide for transport across high and low-affinity copper membrane transports Ctr1 and Fet4. From the cytoplasm, Ccc2 shuttles Cu+ to Golgi bodies, and Pic2 shuttles Cu+ to the mitochondrial matrix. During meiosis in S. pombe, Mfc1 transports Cu+ across the forespore membrane.
Jof 07 00225 g002
Figure 3. Mn2+ uptake and detoxification systems in S. cerevisiae.
Figure 3. Mn2+ uptake and detoxification systems in S. cerevisiae.
Jof 07 00225 g003
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Robinson, J.R.; Isikhuemhen, O.S.; Anike, F.N. Fungal–Metal Interactions: A Review of Toxicity and Homeostasis. J. Fungi 2021, 7, 225. https://doi.org/10.3390/jof7030225

AMA Style

Robinson JR, Isikhuemhen OS, Anike FN. Fungal–Metal Interactions: A Review of Toxicity and Homeostasis. Journal of Fungi. 2021; 7(3):225. https://doi.org/10.3390/jof7030225

Chicago/Turabian Style

Robinson, Janelle R., Omoanghe S. Isikhuemhen, and Felicia N. Anike. 2021. "Fungal–Metal Interactions: A Review of Toxicity and Homeostasis" Journal of Fungi 7, no. 3: 225. https://doi.org/10.3390/jof7030225

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