Next Article in Journal
Marine Bioactive Compounds against Aspergillus fumigatus: Challenges and Future Prospects
Previous Article in Journal
Antimicrobial Activity of Bee-Collected Pollen and Beebread: State of the Art and Future Perspectives
Previous Article in Special Issue
Repurposing Disulfiram (Tetraethylthiuram Disulfide) as a Potential Drug Candidate against Borrelia burgdorferi In Vitro and In Vivo
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Perspective

Antifungal Drug Repurposing

1
Foodborne Toxin Detection and Prevention Research Unit, Western Regional Research Center, Agricultural Research Service, United States Department of Agriculture, Albany, CA 94710, USA
2
Healthy Processed Foods Research Unit, Western Regional Research Center, Agricultural Research Service, United States Department of Agriculture, Albany, CA 94710, USA
3
Henry E. Riggs School of Applied Life Sciences, Keck Graduate Institute, Claremont, CA 91711, USA
4
Department of Biological Sciences, University of the Pacific, Stockton, CA 95211, USA
*
Author to whom correspondence should be addressed.
Antibiotics 2020, 9(11), 812; https://doi.org/10.3390/antibiotics9110812
Submission received: 17 September 2020 / Revised: 30 October 2020 / Accepted: 13 November 2020 / Published: 15 November 2020
(This article belongs to the Special Issue Drugs Repurposing for the Treatment of Bacterial Infections)

Abstract

:
Control of fungal pathogens is increasingly problematic due to the limited number of effective drugs available for antifungal therapy. Conventional antifungal drugs could also trigger human cytotoxicity associated with the kidneys and liver, including the generation of reactive oxygen species. Moreover, increased incidences of fungal resistance to the classes of azoles, such as fluconazole, itraconazole, voriconazole, or posaconazole, or echinocandins, including caspofungin, anidulafungin, or micafungin, have been documented. Of note, certain azole fungicides such as propiconazole or tebuconazole that are applied to agricultural fields have the same mechanism of antifungal action as clinical azole drugs. Such long-term application of azole fungicides to crop fields provides environmental selection pressure for the emergence of pan-azole-resistant fungal strains such as Aspergillus fumigatus having TR34/L98H mutations, specifically, a 34 bp insertion into the cytochrome P450 51A (CYP51A) gene promoter region and a leucine-to-histidine substitution at codon 98 of CYP51A. Altogether, the emerging resistance of pathogens to currently available antifungal drugs and insufficiency in the discovery of new therapeutics engender the urgent need for the development of new antifungals and/or alternative therapies for effective control of fungal pathogens. We discuss the current needs for the discovery of new clinical antifungal drugs and the recent drug repurposing endeavors as alternative methods for fungal pathogen control.

1. Introduction

Fungal infections, such as human invasive aspergillosis, candidiasis, or cryptococcosis caused by Aspergillus, Candida, or Cryptococcus species, respectively, are persistent problems since effective antifungal drugs are often limited [1]. In addition to 5-flucytosine (5FC), only three classes of antifungal drugs are currently applied in clinical settings, namely, azoles, polyenes, and echinocandins; azoles and polyenes were introduced before 1980, while the echinocandin drug caspofungin (CAS) was approved for use in clinics after the year 2000 [2]. These are the three drug classes used in clinical routines to treat invasive/systemic fungal infections and, therefore, emerging resistance incidences of fungal pathogens to these drugs make fungal diseases a global human health concern [3]. Immunocompromised people are at risk of developing fungal diseases.
For instance, the yeast pathogen Candida species are the most common causative agents of bloodstream infections, linked to high morbidity and mortality. While Candida albicans is the most prevalent species infecting humans, cases of non-albicans infections are also continuously growing. Of note, the non-albicans C. auris is an emerging yeast pathogen showing severe human infections and multidrug resistance, with up to 323 United States (US) clinical cases reported in 2018 [4]. C. auris spreads easily between hospitalized patients or nursing home residents, and it has been classified as an “urgent threat” pathogen according to the United States Centers for Disease Control and Prevention’s (CDC’s) 2019 Antibiotic Resistance Threats Report (ARTR) [4]. Numerous other Candida species have also been grouped as “serious threat” pathogens because they cause different types of fungal infections including oral and vaginal infections or severe invasive diseases. Many species of this group are resistant to conventional antifungal drugs, with estimated US hospitalization cases in 2017 of 34,800, resulting in 1700 deaths [4].
In addition, azole-resistant strains of other fungal species, including Aspergillus fumigatus, also cause life-threatening infections, especially in high-risk, immunocompromised people. Azoles are broadly used for treating human fungal pathogens, where the mechanism of action is to inhibit cellular lanosterol 14 alpha-demethylases involved in fungal ergosterol biosynthesis. Azoles are also increasingly applied in agricultural fields to prevent or treat phytopathogenic fungal diseases in crops. Currently, more than 25% of total fungicide sales are azoles [5]. This provides environmental selection pressure for the emergence of pan-azole-resistant strains such as A. fumigatus having the TR34/L98H mutation [5]. The development of azole resistance due to their increased use in human medicine and agriculture is a public health concern, leading to the placement of azole-resistant A. fumigatus on the microorganism watchlist, according to CDC’s 2019 ARTR [4].
The spectrum of activity for the current systemic antifungal drugs, including amphotericin B (AMB), 5FC, fluconazole (FLU), itraconazole (ITR), voriconazole (VOR), posaconazole (POS), isavuconazole (ISA), CAS, micafungin (MICA), or anidulafungin (ANI), have been documented [6,7,8]. In brief, the antifungal spectrum has been determined as follows: C. albicans, Candida glabrata, Candida parapsilosis, and Candida tropicalis (AMB, 5FC, FLU, ITR, VOR, POS, ISA, CAS, MICA, and ANI); Candida krusei (AMB, 5FC, ITR, VOR, POS, ISA, CAS, MICA, and ANI); Candida lusitaniae (5FC, ITR, VOR, POS, ISA, CAS, MICA, and ANI); A. fumigatus (AMB, ITR, VOR, POS, ISA, CAS, MICA, and ANI); Cryptococcus neoformans (AMB, 5FC, FLU, ITR, VOR, POS, and ISA); Fusarium species (AMB, ITR, VOR, POS, and ISA); Scedosporium species (AMB, ITR, VOR, POS, and ISA); Blastomyces dermatitidis, Coccidioides immitis, and Histoplasma capsulatum (AMB, FLU, ITR, VOR, POS, and ISA); Mucorales (AMB, POS, and ISA). Studies indicated that differential susceptibilities of fungal pathogens to the drugs exist depending on the types of fungi or drugs applied. For example, the yeast pathogens C. albicans, C. glabrata, C. parapsilosis, and C. tropicalis were susceptible to all antifungal drugs described (AMB, 5FC, FLU, ITR, VOR, POS, ISA, CAS, MICA, and ANI), while the other two Candida species (C. krusei and C. lusitaniae) did not show sensitivity to FLU or AMB, respectively. Of note, except for the Candida species and A. fumigatus, the other fungi mentioned (namely, C. neoformans, Fusarium species, Scedosporium species, Blastomyces dermatitidis, Coccidioides immitis, and Histoplasma capsulatum) did not exhibit susceptibility to the echinocandin drugs (CAS, MICA, and ANI), whereas the azoles POS and ISA were effective against all fungal pathogens described above.
There have also been persistent efforts to improve the efficacy or to reduce the toxicity of conventional antifungal drugs/intervention strategies. For instance, AMB was the first antifungal drug introduced to clinics over five decades ago, for which several types of formulations have been developed with varying toxicity such as infusion-related reactions and nephrotoxicity [9,10]. Three lipid-associated AMB formulations have been developed, which include the AMB lipid complex (AMB-LC), liposomal AMB (L-AMB), and colloidal dispersion of AMB (AMB-CD) with the recommended therapeutic doses of 5, 3–6, and 3–4 mg/kg/day, respectively [9,10].
However, the development of entirely new antifungal drugs is a very expensive and time-consuming process. It is estimated that the overall timelines and costs from new antifungal lead discovery to regulatory approval, especially for those overcoming drug-resistant fungal pathogens, are 10 years and USD >300 million, respectively. In addition, marketing is estimated to cost USD 400 million over the lifespan of a product [11]. Recently, there have been alternative approaches termed antifungal ‘drug repurposing” via which the new utility of various types of marketed, non-antifungal drugs are repositioned as novel antifungal agents. Here, we discuss the current clinical needs for the development of new antifungal therapy, and we comment on the recent antifungal drug repurposing efforts as alternative approaches for the control of fungal pathogens.

2. Drug Repurposing Approaches

2.1. Repurposing Approaches for the Human Therapeutic Drugs (Non-Antifungals)

Drug repurposing for “medical treatments (other than fungal diseases)” is the repositioning platform of already marketed drugs for treating human diseases to cure new, other types of disorders/health problems such as viral infection, lupus nephritis, and neurodegenerative disease. One of the merits of drug repurposing is that the mechanisms of action, cellular targets, toxicity profile, or safety of the commercial drugs have already been identified, which enables expedited regulatory approval [12,13]. The methodical drug repurposing pipeline largely involves two types of approaches, “experimental testing” approaches such as microtiter plate-based high-throughput screenings and “in silico/computational” approaches that utilize currently available data (omics, drug target, and real-world data, such as the data pertaining to individual’s health status or to the healthcare routinely provided) for the identification of potential new drugs to cure diseases. Systematic drug repurposing needs the accession to and interpretation of molecular, protein, and real-world data, as well as experimental analysis, where data validation in the multicellular or higher organism is the key for industry implementation [12]. In principle, the “antifungal” drug repurposing processes also apply similar approaches for successful drug/compound repositioning (Table S1, Supplementary Materials).

2.2. Repurposing Approaches for the New Antifungal Drugs

We performed a PubMed database search in the National Center for Biotechnology Information (NCBI) [14] (https://www.ncbi.nlm.nih.gov/) by applying the keywords “antifungal drug repurposing”, “repurposing [and] in silico [and] fungi (or antifungal, Candida, Cryptococcus, Aspergillus)” plus “antifungal resistance [and] Food and Drug Administration (FDA) (or FDA-approved drug)”, which retrieved a total of 747 articles (accessed on 14 August and 20 October 2020). Each repurposed drug identified was then searched further in PubMed with the search terms “fungi [and] antifungal [and] repurposed drug name (individual)” to provide a comprehensive antifungal spectrum. Articles relevant to the new antifungal drug development are summarized in Table S1 (Supplementary Materials) [15,16,17,18,19,20,21,22,23,24,25,26,27,28,29,30,31,32,33,34,35,36,37,38,39,40,41,42,43,44,45,46,47,48,49,50,51,52,53,54,55,56,57,58,59,60,61,62,63,64,65,66,67,68,69,70,71,72,73,74,75,76,77,78,79,80,81,82,83,84,85,86,87,88,89,90,91,92,93,94,95,96,97,98,99,100,101,102,103,104,105,106,107,108,109,110,111,112,113,114,115,116,117,118,119,120,121,122,123,124,125,126,127,128,129,130,131,132,133,134,135,136,137,138,139,140,141,142,143,144,145,146,147,148,149,150,151,152,153,154,155,156,157,158,159,160,161,162,163,164,165,166,167,168,169,170]. The remaining articles not selected here mainly described (1) antibacterial, antiviral, or antiprotozoal drug development, (2) anticancer drug development, or (3) drug development for other human diseases/conditions including metabolic diseases such as arachidonic acid metabolism, Parkinson’s/neurodegenerative diseases, immune-mediated disease, altered gene expression, and ATP synthase disorder. Six studies adopted “in silico/computational” approaches including experimental validation (Table 1), while the remaining investigations used “experimental testing” approaches, including the utilization of standard antifungal testing protocols such as CLSI M27-A, CLSI M38-A, and EUCAST-AFST E.DEF 7.3 developed by the Clinical and Laboratory Standard Institute (CLSI) [171] or the European Committee on Antimicrobial Susceptibility Testing (EUCAST) [172], respectively, for breakout determination.

3. Antifungal Drug Repurposing: Current Measures

3.1. In Silico/Computational Repurposing Approaches

In silico/computational repurposing approaches typically use four steps: (1) mining and compilation of pathogen genome data, (2) homology modeling, (3) ligand preparation and molecular docking, and (4) experimental validation in the target pathogens [173]. As described in Table 1, public or nonprofit research sectors such as NCBI, Broad Institute (USA), and the European Molecular Biology Laboratory’s European Bioinformatics Institute (EMBL-EBI) provide updated fungal genome or protein data, which makes the comprehensive mining and compilation of fungal genome/protein data feasible. Protein or chemical databases such as the SWISS-MODEL server, Protein Data Bank (PDB), and PubChem server allow the execution of protein structure homology modeling of drug targets or the computation of 3D structures of candidate compounds. Other online tools, such as the Visual Molecular Dynamics (VMD) program or LigPlot program that automatically plot the protein–ligand interactions, are also currently available (Table 1). The execution of ligand preparation and the molecular docking step rely mainly on the application software, as documented in Table 1.
Noteworthy is the repurposing study performed by de Oliveira et al. [15] targeting the saprobic/dimorphic Paracoccidioides species, a causative agent of the systemic mycosis paracoccidioidomycosis, which adopted three additional steps: (a) identification of “orthologs” in different isolates of the target pathogen, (b) identification of “homologs” in the drug–target databases, and (c) selection of essential targets in the model fungus Saccharomyces cerevisiae system. The study compiled proteins of three Paracoccidioides species (P. lutzii, P. americana, and P. brasiliensis) via the Broad Institute Fungal Genomics Database, followed by the identification of orthologs in Paracoccidioides species. The study by de Oliveira et al. resulted in the selection of two anticancer drug candidates as new, repurposed antifungals, where the mode of action was to inhibit the fungal phosphatidylinositol 3-kinase TOR2 (Target Of Rapamycin 2) [15].
The validation of antifungal activity of new, repurposed drugs to achieve more than 99.9% fungal death (breakpoints) requires standard testing methods developed by CLSI [171] or EUCAST [172] (Section 3.2). Of the six in silico/computational studies described in Table 1, two investigations adopted the CLSI protocols while four other studies applied various agar- or liquid-based antifungal assays. It is expected that the number of in silico/computational investigations will increase in the coming years, especially with the increasing numbers of omics, drug target, and chemical structural data being generated. The adoption of standard methods, such as CLSI or EUCAST, is highly desired for the unbiased/repeatable determination of the breakpoints (and, thus, the efficacy) of repurposed antifungal drugs.

3.2. Experimental Repurposing Approaches

3.2.1. Standard Dilution Methods: CLSI and EUCAST

The majority of articles (Table S1, Supplementary Materials) performed drug repurposing via experimental testing approaches. These include CLSI, EUCAST, or other microdilution/agar assays such as biofilm bioassay, high-throughput ATP content assay, microdilution, and fluorescent microscopic analysis, metabolism and hyphal inhibitory assays, drug diffusion susceptibility testing, human neutrophils, epithelial cell adhesion and invasion assays, murine model, and macrophage assay, among others.
The standard dilution methods developed by the CLSI and EUCAST quantitatively determine (1) minimum inhibitory concentrations (MICs) and minimum fungicidal concentrations (MFC) of drugs/compounds via the microdilution assay settings, as well as whether the antifungal efficacy of drugs/compounds is fungicidal or fungistatic, where fungicidal indicates a ratio of MFC/MIC ≤4 [211], and (2) the levels and types of drug/compound interactions when two drugs/compounds are co-applied, thus calculating the fractional inhibitory concentration indices (FICI) using MIC values or the fractional fungicidal concentration indices (FFCI) using MFC values; synergism indicates FICI or FFCI values ≤0.5, while indifference indicates FICI or FFCI values >0.5–4 [212].
The other method termed the “disc diffusion” test is a relatively inexpensive assay compared to the standard dilution method, for which a few standard assay protocols have been documented in CLSI. CLSI M44 was validated only for azoles and echinocandins for the isolates of Candida species, while CLSI M51-A and supplement M51-S1 qualitatively analyze the efficacy of CAS, triazoles, AMB, etc. [213]. There is a lower agreement between the disc diffusion test results and that of the standard dilution assays, especially in the values from Aspergillus flavus (AMB and VOR) and A. fumigatus (AMB), suggesting that the standard dilution assay seems useful to determine the interpretative breakpoints for both Candida and Aspergillus species.

3.2.2. Biofilm Analysis

Studies have shown that many fungi can alternate planktonic (freely floating, homogeneous cells) and sessile (surface-aggregated, heterogeneous cells) growth, which significantly affects fungal pathogenesis and human infection [214,215,216]. The sessile, multicellular communities of fungi, also known as biofilms, are highly structured fungal communities, which are either adherent to biological or physical surfaces, such as oral mucosa, denture acrylic substrates, and catheters, or form aggregates within the protective extracellular matrix (ECM). The majority of clinically important fungi can produce biofilms, which include filamentous fungal pathogens (Aspergillus, Fusarium, and zygomycetes), yeast pathogens (Blastoschizomyces, Saccharomyces, Malassezia, Trichosporon, Cryptococcus, and Candida species), and Pneumocysitis [214,215,216]. There are also variations in biofilm morphology depending on the types of fungi, namely, (1) C. albicans forms complex morphology with blastospores, hyphae, and ECM, (2) C. neoformans forms an organized structure having yeast cells with a matrix, (3) A. fumigatus forms hyphal cells with ECM, (4) Trichosporon asahii forms yeast and hyphal cells with ECM, etc. [216]. Therefore, in addition to the standard in vitro testing for the planktonic cells, such as CLSI or EUCAST assays, a highly reproducible microtiter plate-based colorimetric measurement determining metabolic activities of the pathogen’s biofilm are also used [217].
In general, the development cycle of a fungal biofilm consists of initial adhesion, colonization, proliferation with ECM production, biofilm maturation, and dispersion [214,215,216]. Various environmental factors induce surface attachment and biofilm formation of fungi, including the flow of body fluids (urine, blood, saliva, mucus), pH, temperature, and host immune factors, whereby biofilms protect fungi from the harsh environments including antifungal drugs (e.g., ECM shields fungal cells from drugs and reduces drug penetration), chemical and physical stress, etc., or enable a community-coordinated gene regulation or metabolism [214,215,216].
Biofilm-forming fungal infections are very difficult to treat, which often involve increased drug-resistance phenotypes [218]. In particular, the triazoles and traditional formulations of polyene drugs are considered inactive against fungal biofilms [219]. Differential antifungal activity of drugs has also been documented against biofilms from different fungal pathogens. For instance, echinocandins and AMB lipid formulations exhibited in vitro and in vivo antifungal activities against C. albicans biofilms, while other fungal biofilms, such as A. fumigatus or C. auris biofilms, were resistant to echinocandin drugs including CAS [214,219].

3.2.3. Phenotypic Variability of Infecting Fungi: Conidia, Hyphal, Yeast, and Filamentous Growth

In addition to biofilm formation, the phenotypic variability of fungi could also play an important role in clinical outcomes of therapeutic interventions, including the repurposed drugs/compounds (Table 2). In C. albicans, the morphological switch from yeast cells to hyphae (filamentous forms) serves as a crucial virulence factor, which promotes infection and invasion in hosts. There was also a positive correlation between the level of azole resistance and the capability to form a hyphal structure; under hypha-inducing conditions, only the C. albicans resistant to azoles could form hyphae while the susceptible isolates could not [220]. In the antifungal drug screenings, the manganese nitrosyl [Mn(PaPy3)(NO)](ClO4) ({Mn-NO}), a biocompatible NO-donating reagent that delivers NO under visible light, has been determined more effective against the hyphal form of C. albicans, when compared to the yeast cells [221]. Triclosan has also been used in oral hygiene products with a broad-spectrum antimicrobial activity. Of note, in C. albicans, triclosan at subinhibitory concentrations antagonized the antifungal activity of the azole drug FLU, which was specific under hypha-inducing conditions [222]. This antagonism could be due to the membranotropic characteristic of triclosan and also the unique composition of hyphal membranes [222]. Meanwhile, in A. fumigatus, the conidia and hyphal forms of fungal fragments were equally susceptible to the AMB and azole drugs, while hyphal clumps were only susceptible to the relatively high concentrations of AMB [223].

3.2.4. Animal Model Systems

Animal model systems (mammalian and nonmammalian models) are important components for antifungal drug development/discovery including the validation of repurposed drugs [224] (Table 3).
Mammalian models are represented by murine, rat, guinea pigs, and rabbits, including both naïve and compromised mice, whereby the antifungal activity of the repurposed drugs can be examined for pharmacokinetics (PK) (e.g., tissue distribution, excretion), pharmacodynamics (PD), immune responses elicited by fungal pathogens, and vaccination attempts [224,225] (see [138] in Table S1, Supplementary Materials). There are various animal models and infection routes, together with various immune suppression regimens. Examples include, but are not limited to, drug screening murine model, murine neutropenic thigh model for determining PK/PD of antifungal drugs, murine model for testing mucocutaneous candidiasis, diabetic murine model of disseminated mucormycosis, murine pulmonary mucormycosis, cyclophosphamide/cortisone immunocompromised murine model of pulmonary mucormycosis, persistently neutropenic rabbit model for investigating acute, invasive pulmonary aspergillosis (IPA), and persistently granulocytopenic rabbit model for characterizing the efficacy of L-AMB against IPA [224,225]. It is important to note that animal testing should be compliant with animal welfare regulation, including Institutional Animal Care and Use Committee (IACUC) review, for all proposed animal experiments [226].
Nonmammalian models currently consist of Drosophila melanogaster (fruit fly), Caenorhabditis elegans (free-living nematode) and Galleria mellonella (wax moth). Examples include biofilm formation in D. melanogaster, melanization and toxicity testing in G. mellonela, and slow and fast killing testing in C. elegans, among others. [227]. When compared to the mammalian models, the nonmammalian models are considered affordable and easy to handle (see [154] in Table S1, Supplementary Materials). While studies have shown that many data from nonmammalian models are in parallel with those obtained from mammalian models, some results did not correlate well. The major drawback of nonmammalian models lies in their unsuitability for microbial vaccination, colonization assessment, challenge research, and immune response. Accordingly, mammalian models are considered to better represent the human condition during the efficacy assessment of the repurposed drugs [224], as also described in Section 4.1.
There have been several types of candidate drugs used for antifungal repurposing (Table S1, Supplementary Materials). These include antipsychiatric, estrogen modulator, antidepressant, antiplatelet aggregation, enzyme (serine palmitoyl-transferase) inhibitor, anticardiovascular, antiarthritis, antistroke, antiatherosclerosis, anticancer, and anthelmintic drugs. However, the use of various drug libraries in the antifungal drug repurposing process is also increasing recently; drug libraries have been prepared/preserved by either public or commercial institutes/vendors, as shown in Table 4.

3.3. Synergism between Repurposed Agents and Conventional Antifungals

Combination therapy in controlling fungal pathogens is defined as a co-application of two or more antifungal drugs to treat fungal infections [230]. Combination therapy has been developed on the basis that co-administration of antifungal drugs having different cellular/molecular targets could effectively eliminate fungal pathogens, especially those resistant to conventional drugs. However, the efficacy of combination therapies often varies depending upon the types of drugs co-applied. Although many studies determined better results for fungal pathogen control with drug combinations, other data exhibited no added merit of drug co-treatment over the individual application of each drug alone, which may be associated with drug antagonism [231]. Risk factors also exist during combination therapy, such as multidrug interactions and cytotoxicity.
Studies have shown the synergistic interaction between repurposed agents and conventional antifungal drugs, such as FLU, AMB, or CAS (see Table 5 for a summary). For example, Spitzer et al. determined that the antifungal capacity of chemicals can be systematically enhanced via the combined application of known commercial drugs, such as FLU, with other types of bioactive compounds from drug repurposing [23]. They found that the repurposed compounds did not have to possess potent antifungal activity on their own, but that the compounds potentiated the FLU antifungal activity with considerable species specificity. These synergistic drug combinations were different from the traditional combination therapies mentioned above [23].
Similarly, the anticholesterol drug lovastatin has been repurposed as a synergistic antifungal modulator to the azole drug ITR against the planktonic cells and biofilms of the yeast pathogen C. albicans; the lovastatin regulation of the ergosterol biosynthetic pathway has been the proposed mechanism of antifungal action [45]. The antifungal mechanism of lovastatin in A. fumigatus also involved ergosterol biosynthesis, which was controlled further by cellular iron homeostasis [65]. In A. fumigatus, iron starvation induces the production of the siderophore triacetylfusarinine C (TAFC), for which mevalonate is the key intermediate for the synthesis of both ergosterol and TAFC, the critical virulence factors [65]. Of note, the expression of the enzyme 3-hydroxy-3-methyl-glutaryl (HMG)-CoA reductase (Hmg1), responsible for the production of mevalonate, was increased under iron starvation, while the synthesis of TAFC was reduced following the lovastatin-mediated inhibition of Hmg1 [65].
In our prior study, considerable augmentation of the control of Cryptococcus species was achieved by the co-application of repurposed compounds, such as octyl gallate or benzaldehydes, with conventional antifungal agents [232]. The Cryptococcus species exhibited higher susceptibility to the inhibition of mitochondrial respiration compared to other yeast pathogens Candida species [232]. This “species-specific” enhancement of sensitivity to the co-treatments resulted from the inability of Cryptococcus species to produce cellular energy (ATPs) via the fermentation process. Studies by Spitzer et al. mentioned above indicated that the differential susceptibility of pathogens to the newly developed drugs or interventions is triggered by the differences in physiological/genetic settings of test strains, which could result in species-specific antifungal responses.
Related antifungal “chemosensitization” has been developed recently as a new intervention strategy, where co-application of a repurposed compound (chemosensitizer), such as food additives, with conventional drugs enhanced the antifungal efficacy of the co-applied drugs [77,233]. A chemosensitizer causes the target pathogen to be more susceptible to the co-applied conventional drug via the modulation of the pathogen’s defense system, such as the oxidative stress signaling system or cell-wall integrity pathway. Considering that the chemosensitizers could also function as probe-like chemicals by negatively affecting specific cellular targets such as antioxidant systems, types of “drug–compound (repurposed) combinations” enable target-specific control of fungal pathogens, including augmentation of the activity of the echinocandin drug CAS by the cell-wall-targeting octyl gallate [234]. The chemosensitization strategy has been applied further to the development of a high-efficiency drug repurposing protocol that could enhance the sensitivity of target pathogens to the drug candidates, thus reducing time/costs for screening new antifungal drugs, as well as overcoming drug/fungicide resistance of fungal pathogens [104].

4. Challenges

4.1. Pioglitazone: Needs for In Vivo Drug Validation

Pioglitazone (PIO) has been used as an adjuvant of AMB for the treatment of cryptococcosis. AMB causes excessive generation of reactive oxygen species linked to compromised renal function. PIO is an agonist of peroxisome proliferator-activated receptor γ, which is used to treat type 2 diabetes and is also used as an adjuvant of many drugs triggering side-effects due to its redox-active and anti-inflammatory characteristics [235]. In a murine model, co-application of PIO and AMB exhibited higher efficacy than AMB alone for the inhibition of yeast pathogens, whereby the combination (PIO + AMB) disrupted yeast transmission from the lungs to the brain, which also eliminated yeasts that reached the central nervous system [236]. PIO did not exhibit in vitro antifungal activity, nor did it affect the AMB-mediated fungicidal activity of macrophages; however, PIO as a therapeutic adjuvant counteracted the oxidative bursts after the reduction of the fungal burden, thus relieving the oxidative stress damages to the host (in vivo) and preventing the establishment of meningoencephalitis [236]. The fact that the in vitro antifungal activity of PIO and AMB co-application was determined as “indifferent” while that of in vivo administration increased the survivability of the animals, compared to AMB alone, strongly suggests the importance of the in vivo validation of drug repurposing. The in vitro CLSI or EUCAST testing alone could miss identifying highly effective antifungal adjuvants such as PIO (also observed in other drug developments, personal communication [237], American Chemical Society National Meeting, 2020).

4.2. Resistance to Repurposed Drugs/Compounds: Cinnamic and Benzoic Derivatives

The model yeast Saccharomyces cerevisiae has been used as a useful screening system for identifying antifungal agents in view that (1) the S. cerevisiae genome has been sequenced and well-characterized ([238], accessed 3 September 2020), (2) S. cerevisiae gene deletion mutants have been very useful for investigating the mechanisms or target genes of screened leads [239], and (3) many genes in S. cerevisiae are homologous to those of fungal pathogens [240]. The antifungal compounds screened via S. cerevisiae (wild type or mutants) also exhibited broad-spectrum antifungal activities against pathogenic yeasts (Candida and Cryptococcus species) and filamentous fungi (Aspergillus, Fusarium, and Scedosporium species) [2].
Caution should be exercised during the high-throughput repurposing process so as not to overlook the tolerant response of certain mutants. For instance, cinnamic acids are generally recognized as safe (GRAS) compounds, which have been used as food additives [241]. In recent repurposing studies, cinnamic acid derivatives have been investigated as antifungal alternatives which target fungal cell-wall biosynthesis and integrity [242,243,244]. While the wild type and cell-wall integrity mutants of S. cerevisiae showed a sensitive response to the selected cinnamic acids, such as 3- or 4-methoxycinnamic acids, the glutathione reductase mutant (glr1Δ) was hyper-tolerant to 4-methoxycinnamic acid when compared to other test strains [243] (Figure 1). This type of hyper-tolerance was eliminated by 4-methylcinnamic acid, which is the structural derivative of 4-methoxycinnamic acid having a deoxygenated para methyl moiety (Figure 1). Glutathione reductase is necessary for the reduction of the oxidized glutathione (GSSG) to reduced glutathione (GSH) to maintain cellular redox homeostasis [245]. While the study showed the structure–activity relationship of cinnamic derivatives in targeting fungal cell-wall components where the para methyl moiety is critical to overcoming the glr1Δ hyper-tolerance to 4-methoxycinnamic acid, the investigation also highlighted the importance of comprehensive determination of the “gene–compound” interaction/response, thus avoiding unfavorable outcomes including fungal tolerance to the repurposed agents (Figure 1; Figure 2b for the scheme) during drug repurposing.
Edible plants including herbs are rich sources of bioactive metabolites that possess various hepatoprotective, antihypertensive, antitumor, or immunomodulatory effects. However, natural ingredients in plant extracts, such as benzoic derivatives, could negatively affect the fungal signaling mutants (for example, A. fumigatus antioxidant mitogen-activated protein kinase (MAPK) mutants sakAΔ and mpkCΔ [246,247]) where fungal MAPK mutants showed tolerance to the benzoic ingredients, while the wild-type strains remain susceptible to the molecules (our unpublished observation; see Figure 2a for fungal bioassay and Figure 2b for the scheme). Collectively, studies proved the significance of “gene–compound” interaction analysis during the preclinical stage of drug repurposing (regardless of the sources of the drugs or compounds whether synthetic, natural, crude extracts, or purified), thus circumventing the unfavorable downside of repurposed drugs.

5. Summary

Current antifungal intervention strategies often encounter limited efficiency in controlling fungal pathogens. Infections of the bloodstream or lungs by Candida species or the airborne Cryptococcus/Aspergillus species, respectively, that are resistant to conventional drugs cause serious health issues. Resistance to drugs develops via the repeated usage of antifungal agents over time (acquired resistance), while certain fungal species are intrinsically resistant to the conventional drugs (intrinsic resistance); examples of intrinsic resistance include azole resistance (C. glabrata, C. krusei, and C. auris), echinocandin resistance (Cryptococcus and Fusarium species), and polyene resistance (C. auris and A. terreus) [248,249].
Drug repurposing for fungal treatments is an alternative strategy for developing new antifungals. In this paper, two types of drug repurposing approaches were discussed, in silico/computational approaches and experimental approaches. While the majority of drug repurposing studies adopted experimental repurposing platforms, it is expected that the numbers of in silico/computational investigations will be increased in the future considering the increasing numbers of in silico data including omics, drug target, and chemical structural data. The preclinical validation of the efficacy of the repurposed drugs will require the testing of molecules in the model systems with the unbiased determination of breakpoints via the standard microdilution protocols developed by CLSI or EUCAST.
While repurposed drugs could be applied independently as novel antifungal agents for treating fungal pathogens, they can also function as effective synergists/adjuvants in formulations to conventional antifungal drugs such as FLU [23]. In contrast to traditional combination therapy, co-application with a repurposed drug could avoid drug antagonism, multidrug interactions, and cytotoxicity. The One Health approach acknowledges that human, animal, and environmental health is closely linked [250], for which drug repurposing could provide solutions to eliminate resistant fungi such as pan-azole-resistant Aspergillus species [251].
In summary, drug repurposing could provide promising alternatives to current antifungal practices. Future inclusion of additional resources, in addition to the one described in this paper, such as the DrugCentral database [252] and Aggregate Analysis of ClinicalTrials.gov (AACT) database [253], would improve the antifungal drug repurposing processes that have the potential to benefit agriculture, food security, and animal and human health.

Supplementary Materials

The following are available online at https://www.mdpi.com/2079-6382/9/11/812/s1, Table S1: Characteristics of repurposed drugs/compounds for control of fungal pathogens.

Author Contributions

Conceptualization, J.H.K. and L.W.C.; methodology, J.H.K. and K.M.L.; formal analysis, J.H.K., C.C.T., M.M.S., and K.M.L.; investigation, J.H.K., K.M.L., and C.C.T.; resources, M.F. and N.M.; data curation, J.H.K. and K.L.C.; writing—original draft preparation, J.H.K.; writing—review and editing, J.H.K., C.C.T., K.L.C., M.M.S., M.F., N.M., K.M.L., and L.W.C.; visualization, K.L.C. and N.M.; supervision, L.W.C.; project administration, L.W.C. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Acknowledgments

This research was conducted under USDA-ARS CRIS Project 5325-42000-039-00D. We thank Siov Sarreal and De Angela Ford, Foodborne Toxin Detection and Prevention Research Unit, Western Regional Research Center, USDA-ARS, for technical assistance.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Perfect, J.R. The antifungal pipeline: A reality check. Nat. Rev. Drug Discov. 2017, 16, 603–616. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  2. Roemer, T.; Krysan, D.J. Antifungal drug development: Challenges, unmet clinical needs, and new approaches. Cold Spring Harb. Perspect. Med. 2014, 4. [Google Scholar] [CrossRef] [PubMed]
  3. McCarthy, M.W.; Kontoyiannis, D.P.; Cornely, O.A.; Perfect, J.R.; Walsh, T.J. Novel agents and drug targets to meet the challenges of resistant Fungi. J. Infect. Dis. 2017, 216, S474–S483. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  4. CDC. Antibiotic Resistance Threats in the United States; U.S. Department of Health and Human Services, CDC: Atlanta, GA, USA, 2019. [CrossRef] [Green Version]
  5. Bowyer, P.; Denning, D.W. Environmental fungicides and triazole resistance in Aspergillus. Pest. Manag. Sci. 2014, 70, 173–178. [Google Scholar] [CrossRef] [PubMed]
  6. Nett, J.E.; Andes, D.R. Antifungal agents: Spectrum of activity, pharmacology, and clinical indications. Infect. Dis. Clin. N. Am. 2016, 30, 51–83. [Google Scholar] [CrossRef] [PubMed]
  7. Nami, S.; Aghebati-Maleki, A.; Morovati, H.; Aghebati-Maleki, L. Current antifungal drugs and immunotherapeutic approaches as promising strategies to treatment of fungal diseases. Biomed. Pharmacother. 2019, 110, 857–868. [Google Scholar] [CrossRef] [PubMed]
  8. Houšť, J.; Spížek, J.; Havlíček, V. Antifungal drugs. Metabolites 2020, 10, 106. [Google Scholar] [CrossRef] [Green Version]
  9. Marena, G.D.; Ramos, M.A.D.S.; Bauab, T.M.; Chorilli, M. A critical review of analytical methods for quantification of amphotericin B in biological samples and pharmaceutical formulations. Crit. Rev. Anal. Chem. 2020. [Google Scholar] [CrossRef]
  10. Hamill, R.J. Amphotericin B formulations: A comparative review of efficacy and toxicity. Drugs 2013, 73, 919–934. [Google Scholar] [CrossRef]
  11. Tillotson, J.; Tillotson, G.S. The regulatory pathway for antifungal drugs: A US perspective. Clin. Infect. Dis. 2015, 61, S678–S683. [Google Scholar] [CrossRef] [Green Version]
  12. Cha, Y.; Erez, T.; Reynolds, I.J.; Kumar, D.; Ross, J.; Koytiger, G.; Kusko, R.; Zeskind, B.; Risso, S.; Kagan, E.; et al. Drug repurposing from the perspective of pharmaceutical companies. Br. J. Pharmacol. 2018, 175, 168–180. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  13. Hernandez, J.J.; Pryszlak, M.; Smith, L.; Yanchus, C.; Kurji, N.; Shahani, V.M.; Molinski, S.V. Giving drugs a second chance: Overcoming regulatory and financial hurdles in repurposing approved drugs as cancer therapeutics. Front. Oncol. 2017, 7. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  14. National Center for Biotechnology Information. PubMed. Available online: https://www.ncbi.nlm.nih.gov/pubmed (accessed on 14 August 2020).
  15. De Oliveira, A.A.; Neves, B.J.; Silva, L.D.C.; Soares, C.M.D.A.; Andrade, C.H.; Pereira, M. Drug repurposing for paracoccidioidomycosis through a computational chemogenomics framework. Front. Microbiol 2019, 10. [Google Scholar] [CrossRef] [PubMed]
  16. Khedr, M.A.; Massarotti, A.; Mohamed, M.E. Rational discovery of (+) (S) abscisic acid as a potential antifungal agent: A repurposing approach. Sci. Rep. 2018, 8, 8565. [Google Scholar] [CrossRef]
  17. Krajaejun, T.; Lohnoo, T.; Yingyong, W.; Rujirawat, T.; Kumsang, Y.; Jongkhajornpong, P.; Theerawatanasirikul, S.; Kittichotirat, W.; Reamtong, O.; Yolanda, H. The repurposed drug disulfiram inhibits urease and aldehyde dehydrogenase and prevents In Vitro growth of the oomycete Pythium insidiosum. Antimicrob. Agents Chemother. 2019, 63, e00609–e00619. [Google Scholar] [CrossRef] [Green Version]
  18. Rana, R.; Sharma, R.; Kumar, A. Repurposing of fluvastatin against Candida albicans CYP450 lanosterol 14 α-demethylase, a target enzyme for antifungal therapy: An In silico and In vitro study. Curr. Mol. Med. 2019, 19, 506–524. [Google Scholar] [CrossRef]
  19. Tavakkoli, A.; Johnston, T.P.; Sahebkar, A. Antifungal effects of statins. Pharmacol. Ther. 2020, 208, 107483. [Google Scholar] [CrossRef]
  20. Capoci, I.R.G.; Faria, D.R.; Sakita, K.M.; Rodrigues-Vendramini, F.A.V.; Bonfim-Mendonça, P.D.S.; Becker, T.C.A.; Kioshima, É.S.; Svidzinski, T.I.E.; Maigret, B. Repurposing approach identifies new treatment options for invasive fungal disease. Bioorg. Chem. 2019, 84, 87–97. [Google Scholar] [CrossRef]
  21. Gowri, M.; Jayashree, B.; Jeyakanthan, J.; Girija, E.K. Sertraline as a promising antifungal agent: Inhibition of growth and biofilm of Candida auris with special focus on the mechanism of action in vitro. J. Appl. Microbiol. 2020, 128, 426–437. [Google Scholar] [CrossRef]
  22. Khan, S.; Singhal, S.; Mathur, T.; Upadhyay, D.J.; Rattan, A. Antifungal potential of disulfiram. Nippon Ishinkin Gakkai Zasshi 2007, 48, 109–113. [Google Scholar] [CrossRef] [Green Version]
  23. Spitzer, M.; Griffiths, E.; Blakely, K.M.; Wildenhain, J.; Ejim, L.; Rossi, L.; De Pascale, G.; Curak, J.; Brown, E.; Tyers, M.; et al. Cross-species discovery of syncretic drug combinations that potentiate the antifungal fluconazole. Mol. Syst. Biol. 2011, 7, 499. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  24. Villanueva-Lozano, H.; González, G.M.; Espinosa-Mora, J.E.; Bodden-Mendoza, B.A.; Andrade, A.; Martínez-Reséndez, M.F.; Treviño-Rangel, R.d.J. Evaluation of the expanding spectrum of sertraline against uncommon fungal pathogens. J. Infect. Chemother. 2020, 26, 309–311. [Google Scholar] [CrossRef] [PubMed]
  25. Wall, G.; Chaturvedi, A.K.; Wormley, F.L.; Wiederhold, N.P.; Patterson, H.P.; Patterson, T.F.; Lopez-Ribot, J.L. Screening a repurposing library for inhibitors of multidrug-resistant Candida auris identifies ebselen as a repositionable candidate for antifungal drug development. Antimicrob. Agents Chemother. 2018, 62, e01084-18. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  26. Venturini, T.P.; Chassot, F.; Loreto, É.S.; Keller, J.T.; Azevedo, M.I.; Zeni, G.; Santurio, J.M.; Alves, S.H. Antifungal activities of diphenyl diselenide and ebselen alone and in combination with antifungal agents against Fusarium spp. Med. Mycol. 2016, 54, 550–555. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  27. Giurg, M.; Gołąb, A.; Suchodolski, J.; Kaleta, R.; Krasowska, A.; Piasecki, E.; Piętka-Ottlik, M. Reaction of bis[(2-chlorocarbonyl)phenyl] diselenide with phenols, aminophenols, and other amines towards diphenyl diselenides with antimicrobial and antiviral properties. Molecules 2017, 22, 974. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  28. Felli Kubiça, T.; Bedin Denardi, L.; de Loreto, É.S.; Zeni, G.; Weiblen, C.; Oliveira, V.; Morais Santurio, J.; Hartz Alves, S. In vitro activity of diphenyl diselenide and ebselen alone and in combination with antifungal agents against Trichosporon. Asahii. Mycoses 2019, 62, 428–433. [Google Scholar] [CrossRef] [PubMed]
  29. Cong, L.; Liao, Y.; Yang, S.; Yang, R. In vitro antifungal activity of sertraline and synergistic effects in combination with antifungal drugs against planktonic forms and biofilms of clinical Trichosporon asahii isolates. PLoS ONE 2016, 11, e0167903. [Google Scholar] [CrossRef]
  30. De Cremer, K.; Lanckacker, E.; Cools, T.L.; Bax, M.; De Brucker, K.; Cos, P.; Cammue, B.P.A.; Thevissen, K. Artemisinins, new miconazole potentiators resulting in increased activity against Candida albicans biofilms. Antimicrob. Agents Chemother. 2015, 59, 421–426. [Google Scholar] [CrossRef] [Green Version]
  31. Gao, L.; Sun, Y.; He, C.; Zeng, T.; Li, M. Synergy between pyrvinium pamoate and azoles against Exophiala dermatitidis. Antimicrob. Agents Chemother. 2018, 62, e02361-17. [Google Scholar] [CrossRef] [Green Version]
  32. Holbrook, S.Y.L.; Garzan, A.; Dennis, E.K.; Shrestha, S.K.; Garneau-Tsodikova, S. Repurposing antipsychotic drugs into antifungal agents: Synergistic combinations of azoles and bromperidol derivatives in the treatment of various fungal infections. Eur. J. Med. Chem. 2017, 139, 12–21. [Google Scholar] [CrossRef]
  33. Simm, C.; May, R.C. Zinc and iron homeostasis: Target-based drug screening as new route for antifungal drug development. Front. Cell Infect. Microbiol. 2019, 9. [Google Scholar] [CrossRef] [PubMed]
  34. Zeng, B.; Li, J.; Wang, Y.; Chen, P.; Wang, X.; Cui, J.; Liu, L.; Hu, X.; Cao, Q.; Xiao, Y.; et al. In vitro and in vivo effects of suloctidil on growth and biofilm formation of the opportunistic fungus Candida albicans. Oncotarget 2017, 8, 69972. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  35. Yousfi, H.; Ranque, S.; Rolain, J.-M.; Bittar, F. In vitro polymyxin activity against clinical multidrug-resistant fungi. Antimicrob. Resist. Infect. Control 2019, 8, 66. [Google Scholar] [CrossRef] [PubMed]
  36. Voget, M.; Lorenz, D.; Lieber-Tenorio, E.; Hauck, R.; Meyer, M.; Cieslicki, M. Is transmission electron microscopy (TEM) a promising approach for qualitative and quantitative investigations of polymyxin B and miconazole interactions with cellular and subcellular structures of Staphylococcus pseudintermedius, Escherichia coli, Pseudomonas aeruginosa and Malassezia pachydermatis? Vet. Microbiol. 2015, 181, 261–270. [Google Scholar] [CrossRef] [PubMed]
  37. Schemuth, H.; Dittmer, S.; Lackner, M.; Sedlacek, L.; Hamprecht, A.; Steinmann, E.; Buer, J.; Rath, P.-M.; Steinmann, J. In vitro activity of colistin as single agent and in combination with antifungals against filamentous fungi occurring in patients with cystic fibrosis. Mycoses 2013, 56, 297–303. [Google Scholar] [CrossRef] [PubMed]
  38. Mashaly, G.; Shrief, R. Candida glabrata complex from patients with healthcare-associated infections in Mansoura University Hospitals, Egypt: Distribution, antifungal susceptibility and effect of fluconazole and polymyxin B combination. Germs 2019, 9, 125–132. [Google Scholar] [CrossRef] [PubMed]
  39. Cui, J.; Ren, B.; Tong, Y.; Dai, H.; Zhang, L. Synergistic combinations of antifungals and anti-virulence agents to fight against Candida albicans. Virulence 2015, 6, 362–371. [Google Scholar] [CrossRef] [Green Version]
  40. Kirchhoff , L.; Olsowski, M.; Zilmans, K.; Dittmer, S.; Haase, G.; Sedlacek, L.; Steinmann, E.; Buer, J.; Rath, P.-M.; Steinmann, J. Biofilm formation of the black yeast-like fungus Exophiala dermatitidis and its susceptibility to antiinfective agents. Sci. Rep. 2017, 7, 42886. [Google Scholar] [CrossRef] [Green Version]
  41. Hsu, L.-H.; Wang, H.-F.; Sun, P.-L.; Hu, F.-R.; Chen, Y.-L. The antibiotic polymyxin B exhibits novel antifungal activity against Fusarium species. Int. J. Antimicrob. Agents 2017, 49, 740–748. [Google Scholar] [CrossRef]
  42. Borjihan, B.; Ogita, A.; Fujita, K.-i.; Doe, M.; Tanaka, T. The cyclic organosulfur compound zwiebelane a from onion (Allium cepa) functions as an enhancer of polymyxin b in fungal vacuole disruption. Planta Med. 2010, 76, 1864–1866. [Google Scholar] [CrossRef] [Green Version]
  43. Blooi, M.; Pasmans, F.; Rouffaer, L.; Haesebrouck, F.; Vercammen, F.; Martel, A. Successful treatment of Batrachochytrium salamandrivorans infections in salamanders requires synergy between voriconazole, polymyxin E and temperature. Sci. Rep. 2015, 5, 11788. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  44. Bidaud, A.L.; Djenontin, E.; Botterel, F.; Chowdhary, A.; Dannaoui, E. Colistin interacts synergistically with echinocandins against Candida auris. Int. J. Antimicrob. Agents 2020, 55, 105901. [Google Scholar] [CrossRef] [PubMed]
  45. Zhou, Y.; Yang, H.; Zhou, X.; Luo, H.; Tang, F.; Yang, J.; Alterovitz, G.; Cheng, L.; Ren, B. Lovastatin synergizes with itraconazole against planktonic cells and biofilms of Candida albicans through the regulation on ergosterol biosynthesis pathway. Appl. Microbiol. Biotechnol. 2018, 102, 5255–5264. [Google Scholar] [CrossRef] [PubMed]
  46. Zhang, X.; Fang, Y.; Jaiseng, W.; Hu, L.; Lu, Y.; Ma, Y.; Furuyashiki, T. Characterization of tamoxifen as an antifungal agent using the yeast schizosaccharomyces pombe model organism. Kobe J. Med. Sci. 2015, 61, E54–E63. [Google Scholar] [PubMed]
  47. Muthular, M.; Bálsamo, F.; Passero, P.; Jewtuchowicz, V.; Miozza, V.; Villalba, M.B.; Brusca, M.I.; Pérez, C. Effects of tamoxifen on periodontal disease and Candida albicans of patients with breast cancer and other pathologies. Future Microbiol. 2019, 14, 129–137. [Google Scholar] [CrossRef] [PubMed]
  48. Hai, T.P.; Van, A.D.; Ngan, N.T.T.; Nhat, L.T.H.; Lan, N.P.H.; Chau, N.V.V.; Thwaites, G.E.; Krysan, D.; Day, J.N. The combination of tamoxifen with amphotericin B, but not with fluconazole, has synergistic activity against the majority of clinical isolates of Cryptococcus neoformans. Mycoses 2019, 62, 818–825. [Google Scholar] [CrossRef] [Green Version]
  49. Butts, A.; Koselny, K.; Chabrier-Roselló, Y.; Semighini, C.P.; Brown, J.C.S.; Wang, X.; Annadurai, S.; DiDone, L.; Tabroff, J.; Childers, W.E.; et al. Estrogen receptor antagonists are anti-cryptococcal agents that directly bind ef hand proteins and synergize with fluconazole In Vivo. mBio 2014, 5, e00765-13. [Google Scholar] [CrossRef] [Green Version]
  50. Naeimi Eshkaleti, M.; Kordbacheh, P.; Hashemi, S.J.; Falahati, M.; Zaini, F.; Mirhendi, H.; Safara, M.; Hosseinpoor, L. In vitro activity of amphotericin B in combination with statins against clinical and environmental Rhizopus oryzae strains. Iran. J. Public Health 2019, 48, 943–948. [Google Scholar] [CrossRef] [Green Version]
  51. Nyilasi, I.; Kocsubé, S.; Krizsán, K.; Galgóczy, L.; Papp, T.; Pesti, M.; Nagy, K.; Vágvölgyi, C. Susceptibility of clinically important dermatophytes against statins and different statin-antifungal combinations. Med. Mycol. 2013, 52, 140–148. [Google Scholar] [CrossRef] [Green Version]
  52. Nyilasi, I.; Kocsubé, S.; Krizsán, K.; Galgóczy, L.; Pesti, M.; Papp, T.; Vágvölgyi, C. In vitro synergistic interactions of the effects of various statins and azoles against some clinically important fungi. FEMS Microbiol. Lett. 2010, 307, 175–184. [Google Scholar] [CrossRef] [Green Version]
  53. Qiao, J.; Kontoyiannis, D.P.; Wan, Z.; Li, R.; Liu, W. Antifungal activity of statins against Aspergillus species. Med. Mycol. 2007, 45, 589–593. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  54. Roze, L.V.; Linz, J.E. Lovastatin triggers an apoptosis-like cell death process in the fungus Mucor. Racemosus. Fungal. Genet. Biol. 1998, 25, 119–133. [Google Scholar] [CrossRef] [PubMed]
  55. Cabral, M.E.; Figueroa, L.I.C.; Fariña, J.I. Synergistic antifungal activity of statin–azole associations as witnessed by Saccharomyces cerevisiae- and Candida utilis-bioassays and ergosterol quantification. Rev. Iberoam. Micol. 2013, 30, 31–38. [Google Scholar] [CrossRef] [PubMed]
  56. Chamilos, G.; Lewis, R.E.; Kontoyiannis, D.P. Lovastatin has significant activity against zygomycetes and interacts synergistically with voriconazole. Antimicrob. Agents Chemother. 2006, 50, 96–103. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  57. Galgóczy, L.; Papp, T.; Kovács, L.; Ördögh, L.; Vágvölgyi, C. In vitro activity of phenothiazines and their combinations with amphotericin B against Zygomycetes causing rhinocerebral zygomycosis. Med. Mycol. 2009, 47, 331–335. [Google Scholar] [CrossRef] [Green Version]
  58. Lorenz, R.T.; Parks, L.W. Effects of lovastatin (mevinolin) on sterol levels and on activity of azoles in Saccharomyces cerevisiae. Antimicrob. Agents Chemother. 1990, 34, 1660–1665. [Google Scholar] [CrossRef] [Green Version]
  59. Macreadie, I.G.; Johnson, G.; Schlosser, T.; Macreadie, P.I. Growth inhibition of Candida species and Aspergillus fumigatus by statins. FEMS Microbiol. Lett. 2006, 262, 9–13. [Google Scholar] [CrossRef] [Green Version]
  60. Alem, M.A.S.; Douglas, L.J. Effects of aspirin and other nonsteroidal anti-inflammatory drugs on biofilms and planktonic cells of Candida albicans. Antimicrob. Agents Chemother. 2004, 48, 41–47. [Google Scholar] [CrossRef] [Green Version]
  61. Leeuw, N.J.; Swartt, C.W.; Ncango, D.M.; Kriell, W.M.; Pohl, C.H.; van Wyk, P.W.J.; Kock, J.L.F. Anti-inflammatory drugs selectively target sporangium development in Mucor. Can. J. Microbiol. 2009, 55, 1392–1396. [Google Scholar] [CrossRef]
  62. Yang, S.; Liao, Y.; Cong, L.; Lu, X.; Yang, R. In Vitro interactions between non-steroidal anti-inflammatory drugs and antifungal agents against planktonic and biofilm forms of Trichosporon asahii. PLoS ONE 2016, 11, e0157047. [Google Scholar] [CrossRef]
  63. Ogundeji, A.O.; Pohl, C.H.; Sebolai, O.M. Repurposing of aspirin and ibuprofen as candidate anti-Cryptococcus drugs. Antimicrob. Agents Chemother. 2016, 60, 4799–4808. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  64. Rosato, A.; Catalano, A.; Carocci, A.; Carrieri, A.; Carone, A.; Caggiano, G.; Franchini, C.; Corbo, F.; Montagna, M.T. In vitro interactions between anidulafungin and nonsteroidal anti-inflammatory drugs on biofilms of Candida spp. Bioorg. Med. Chem. 2016, 24, 1002–1005. [Google Scholar] [CrossRef] [PubMed]
  65. Yasmin, S.; Alcazar-Fuoli, L.; Gründlinger, M.; Puempel, T.; Cairns, T.; Blatzer, M.; Lopez, J.F.; Grimalt, J.O.; Bignell, E.; Haas, H. Mevalonate governs interdependency of ergosterol and siderophore biosyntheses in the fungal pathogen Aspergillus fumigatus. Proc. Natl. Acad. Sci. USA 2012, 109, E497–E504. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  66. Rossi, S.A.; de Oliveira, H.C.; Agreda-Mellon, D.; Lucio, J.; Mendes-Giannini, M.J.S.; García-Cambero, J.P.; Zaragoza, O. Identification of off-patent drugs that show synergism with amphotericin B or that present antifungal action against Cryptococcus neoformans and Candida spp. Antimicrob. Agents Chemother. 2020, 64, e01921-19. [Google Scholar] [CrossRef]
  67. Król, J.; Nawrot, U.; Bartoszewicz, M. Anti-candidal activity of selected analgesic drugs used alone and in combination with fluconazole, itraconazole, voriconazole, posaconazole and isavuconazole. J. Mycol. Med. 2018, 28, 327–331. [Google Scholar] [CrossRef]
  68. Delattin, N.; De Brucker, K.; Vandamme, K.; Meert, E.; Marchand, A.; Chaltin, P.; Cammue, B.P.A.; Thevissen, K. Repurposing as a means to increase the activity of amphotericin B and caspofungin against Candida albicans biofilms. J. Antimicrob. Chemother. 2013, 69, 1035–1044. [Google Scholar] [CrossRef] [Green Version]
  69. Al-Janabi, A.A.H.S. Determination of antidermatophytic effects of non-steroidal anti-inflammatory drugs on Trichophyton mentagrophytes and Epidermophyton floccosum. Mycoses 2011, 54, e443–e448. [Google Scholar] [CrossRef]
  70. Al-Bakri, A.G.; Othman, G.; Bustanji, Y. The assessment of the antibacterial and antifungal activities of aspirin, EDTA and aspirin-EDTA combination and their effectiveness as antibiofilm agents. J. Appl. Microbiol. 2009, 107, 280–286. [Google Scholar] [CrossRef]
  71. Roohparvar, R.; Huser, A.; Zwiers, L.-H.; De Waard, M.A. Control of Mycosphaerella graminicola on wheat seedlings by medical drugs known to modulate the activity of ATP-binding cassette transporters. Appl. Environ. Microbiol. 2007, 73, 5011–5019. [Google Scholar] [CrossRef] [Green Version]
  72. Loreto, E.S.; Tondolo, J.S.M.; Santurio, J.M.; Alves, S.H. Screening of antibacterial drugs for antimicrobial activity against Pythium insidiosum. Med. Mycol. 2018, 57, 523–525. [Google Scholar] [CrossRef]
  73. Ko, H.-T.; Hsu, L.-H.; Yang, S.-Y.; Chen, Y.-L. Repurposing the thrombopoietin receptor agonist eltrombopag as an anticryptococcal agent. Med. Mycol. 2019, 58, 493–504. [Google Scholar] [CrossRef] [PubMed]
  74. Caldara, M.; Marmiroli, N. Tricyclic antidepressants inhibit Candida albicans growth and biofilm formation. Int. J. Antimicrob. Agents 2018, 52, 500–505. [Google Scholar] [CrossRef] [PubMed]
  75. Butts, A.; DiDone, L.; Koselny, K.; Baxter, B.K.; Chabrier-Rosello, Y.; Wellington, M.; Krysan, D.J. A repurposing approach identifies off-patent drugs with fungicidal cryptococcal activity, a common structural chemotype, and pharmacological properties relevant to the treatment of cryptococcosis. Eukaryot. Cell 2013, 12, 278–287. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  76. Bagar, T.; Benčina, M. Antiarrhythmic drug amiodarone displays antifungal activity, induces irregular calcium response and intracellular acidification of Aspergillus niger—Amiodarone targets calcium and pH homeostasis of A. niger. Fungal. Genet. Biol. 2012, 49, 779–791. [Google Scholar] [CrossRef] [PubMed]
  77. Eldesouky, H.E.; Salama, E.A.; Li, X.; Hazbun, T.R.; Mayhoub, A.S.; Seleem, M.N. Repurposing approach identifies pitavastatin as a potent azole chemosensitizing agent effective against azole-resistant Candida species. Sci. Rep. 2020, 10, 7525. [Google Scholar] [CrossRef]
  78. Ji, C.; Liu, N.; Tu, J.; Li, Z.; Han, G.; Li, J.; Sheng, C. Drug repurposing of haloperidol: Discovery of new benzocyclane derivatives as potent antifungal agents against cryptococcosis and candidiasis. ACS Infect. Dis. 2020, 6, 768–786. [Google Scholar] [CrossRef]
  79. Stylianou, M.; Kulesskiy, E.; Lopes, J.P.; Granlund, M.; Wennerberg, K.; Urban, C.F. Antifungal application of nonantifungal drugs. Antimicrob. Agents Chemother. 2014, 58, 1055–1062. [Google Scholar] [CrossRef] [Green Version]
  80. Venturini, T.P.; Al-Hatmi, A.M.S.; Rossato, L.; Azevedo, M.I.; Keller, J.T.; Weiblen, C.; Santurio, J.M.; Alves, S.H. Do antibacterial and antifungal combinations have better activity against clinically relevant fusarium species? in vitro synergism. Int. J. Antimicrob. Agents 2018, 51, 784–788. [Google Scholar] [CrossRef]
  81. Aneke, C.I.; Rhimi, W.; Otranto, D.; Cafarchia, C. Synergistic effects of efflux pump modulators on the azole antifungal susceptibility of Microsporum canis. Mycopathologia 2020, 185, 279–288. [Google Scholar] [CrossRef]
  82. Iatta, R.; Puttilli, M.R.; Immediato, D.; Otranto, D.; Cafarchia, C. The role of drug efflux pumps in Malassezia pachydermatis and Malassezia furfur defence against azoles. Mycoses 2017, 60, 178–182. [Google Scholar] [CrossRef]
  83. Mamouei, Z.; Alqarihi, A.; Singh, S.; Xu, S.; Mansour, M.K.; Ibrahim, A.S.; Uppuluri, P. Alexidine Dihydrochloride Has Broad-Spectrum Activities against Diverse Fungal Pathogens. mSphere 2018, 3. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  84. Shekhar-Guturja, T.; Gunaherath, G.M.K.B.; Wijeratne, E.M.K.; Lambert, J.-P.; Averette, A.F.; Lee, S.C.; Kim, T.; Bahn, Y.-S.; Tripodi, F.; Ammar, R.; et al. Dual action antifungal small molecule modulates multidrug efflux and TOR signaling. Nat. Chem. Biol. 2016, 12, 867–875. [Google Scholar] [CrossRef] [PubMed]
  85. Tong, Y.; Liu, M.; Zhang, Y.; Liu, X.; Huang, R.; Song, F.; Dai, H.; Ren, B.; Sun, N.; Pei, G.; et al. Beauvericin counteracted multi-drug resistant Candida albicans by blocking ABC transporters. Synth. Syst. Biotechnol. 2016, 1, 158–168. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  86. Zhang, L.; Yan, K.; Zhang, Y.; Huang, R.; Bian, J.; Zheng, C.; Sun, H.; Chen, Z.; Sun, N.; An, R.; et al. High-throughput synergy screening identifies microbial metabolites as combination agents for the treatment of fungal infections. Proc. Natl. Acad. Sci. USA 2007, 104, 4606–4611. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  87. Shekhar-Guturja, T.; Tebung, W.A.; Mount, H.; Liu, N.; Köhler, J.R.; Whiteway, M.; Cowen, L.E. Beauvericin potentiates azole activity via inhibition of multidrug efflux, blocks Candida albicans morphogenesis, and is effluxed via Yor1 and circuitry controlled by Zcf29. Antimicrob. Agents Chemother. 2016, 60, 7468–7480. [Google Scholar] [CrossRef] [Green Version]
  88. Mishra, N.N.; Ali, S.; Shukla, P.K. Arachidonic acid affects biofilm formation and PGE2 level in Candida albicans and non-albicans species in presence of subinhibitory concentration of fluconazole and terbinafine. Braz. J. Infect. Dis. 2014, 18, 287–293. [Google Scholar] [CrossRef] [Green Version]
  89. Fukuda, T.; Arai, M.; Yamaguchi, Y.; Masuma, R.; Tomoda, H.; Omura, S. New beauvericins, potentiators of antifungal miconazole activity, Produced by Beauveria sp. FKI-1366. I. Taxonomy, fermentation, isolation and biological properties. J. Antibiot. 2004, 57, 110–116. [Google Scholar] [CrossRef] [Green Version]
  90. Ells, R.; Kock, J.L.F.; Van Wyk, P.W.J.; Botes, P.J.; Pohl, C.H. Arachidonic acid increases antifungal susceptibility of Candida albicans and Candida dubliniensis. J. Antimicrob. Chemother. 2008, 63, 124–128. [Google Scholar] [CrossRef] [Green Version]
  91. Coelho, R.A.; Joffe, L.S.; Alves, G.M.; Figueiredo-Carvalho, M.H.G.; Brito-Santos, F.; Amaral, A.C.F.; Rodrigues, M.L.; Almeida-Paes, R. A screening of the MMV Pathogen Box® reveals new potential antifungal drugs against the etiologic agents of chromoblastomycosis. PLoS ONE 2020, 15, e0229630. [Google Scholar] [CrossRef]
  92. Thangamani, S.; Maland, M.; Mohammad, H.; Pascuzzi, P.E.; Avramova, L.; Koehler, C.M.; Hazbun, T.R.; Seleem, M.N. Repurposing approach identifies auranofin with broad spectrum antifungal activity that targets Mia40-Erv1 pathway. Front. Cell Infect. Microbiol. 2017, 7. [Google Scholar] [CrossRef] [Green Version]
  93. Wiederhold, N.P.; Patterson, T.F.; Srinivasan, A.; Chaturvedi, A.K.; Fothergill, A.W.; Wormley, F.L.; Ramasubramanian, A.K.; Lopez-Ribot, J.L. Repurposing auranofin as an antifungal: In vitro activity against a variety of medically important fungi. Virulence 2017, 8, 138–142. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  94. Yousfi, H.; Cassagne, C.; Ranque, S.; Rolain, J.-M.; Bittar, F. Repurposing of ribavirin as an adjunct therapy against invasive Candida strains in an In Vitro study. Antimicrob. Agents Chemother. 2019, 63. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  95. Zhang, M.; Yan, H.; Lu, M.; Wang, D.; Sun, S. Antifungal activity of ribavirin used alone or in combination with fluconazole against Candida albicans is mediated by reduced virulence. Int. J. Antimicrob. Agents 2020, 55, 105804. [Google Scholar] [CrossRef] [PubMed]
  96. Yousfi, H.; Ranque, S.; Cassagne, C.; Rolain, J.-M.; Bittar, F. Identification of repositionable drugs with novel antimycotic activity by screening the Prestwick Chemical Library against emerging invasive moulds. J. Glob. Antimicrob. Resist. 2020, 21, 314–317. [Google Scholar] [CrossRef]
  97. Xu, Y.; He, Y.; Zhou, L.; Gao, C.; Sun, S.; Wang, X.; Pang, G. Effects of contact lens solution disinfectants against filamentous fungi. Optom. Vis. Sci. 2014, 91, 1440–1445. [Google Scholar] [CrossRef]
  98. Venturini, T.P.; Rossato, L.; Chassot, F.; Keller, J.T.; Piasentin, F.B.; Santurio, J.M.; Alves, S.H. In vitro synergistic combinations of pentamidine, polymyxin B, tigecycline and tobramycin with antifungal agents against Fusarium spp. J. Med. Microbiol. 2016, 65, 770–774. [Google Scholar] [CrossRef]
  99. Santos, V.A.; Viera, P.V.; Oliveira, A.M.; Zanin, M.H.; Borsatti, M.A. Antifungal effect of electrospun nanofibers containing cetylpyridinium chloride against Candida albicans. Braz. Oral. Res. 2014, 28. [Google Scholar] [CrossRef] [Green Version]
  100. Ortiz, S.C.; Huang, M.; Hull, C.M. Spore germination as a target for antifungal therapeutics. Antimicrob. Agents Chemother. 2019, 63. [Google Scholar] [CrossRef]
  101. Sun, W.; Park, Y.-D.; Sugui, J.A.; Fothergill, A.; Southall, N.; Shinn, P.; McKew, J.C.; Kwon-Chung, K.J.; Zheng, W.; Williamson, P.R. Rapid identification of antifungal compounds against Exserohilum rostratum using high throughput drug repurposing screens. PLoS ONE 2013, 8, e70506, Erratum in 2013, 8, doi:10.1371/annotation/df5a5a46-cf30-4842-bf11-b6cee36c1f9b. [Google Scholar] [CrossRef]
  102. Schwarz, P.; Schwarz, P.V.; Felske-Zech, H.; Dannaoui, E. In vitro interactions between isavuconazole and tacrolimus, cyclosporin A or sirolimus against Mucorales. J. Antimicrob. Chemother. 2019, 74, 1921–1927. [Google Scholar] [CrossRef]
  103. Schwarz, P.; Dannaoui, E. In vitro interaction between isavuconazole and tacrolimus, cyclosporin A, or sirolimus against Aspergillus species. J. Fungi 2020, 6, 103. [Google Scholar] [CrossRef] [PubMed]
  104. Kim, J.H.; Chan, K.L.; Cheng, L.W.; Tell, L.A.; Byrne, B.A.; Clothier, K.; Land, K.M. High efficiency drug repurposing design for new antifungal agents. Meth. Protoc. 2019, 2, 31. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  105. Clark, A.; Hemmelgarn, T.; Danziger-Isakov, L.; Teusink, A. Intravenous pentamidine for Pneumocystis carinii/jiroveci pneumonia prophylaxis in pediatric transplant patients. Pediatr. Transplant. 2015, 19, 326–331. [Google Scholar] [CrossRef] [PubMed]
  106. Li, W.; Zhang, Z.-W.; Luo, Y.; Liang, N.; Pi, X.-X.; Fan, Y.-M. Molecular epidemiology, in vitro susceptibility and exoenzyme screening of Malassezia clinical isolates. J. Med. Microbiol. 2020, 69, 436–442. [Google Scholar] [CrossRef]
  107. Nixon, G.L.; McEntee, L.; Johnson, A.; Farrington, N.; Whalley, S.; Livermore, J.; Natal, C.; Washbourn, G.; Bibby, J.; Berry, N.; et al. Repurposing and reformulation of the antiparasitic agent flubendazole for treatment of cryptococcal meningoencephalitis, a neglected fungal disease. Antimicrob. Agents Chemother. 2018, 62, e01909–e01917. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  108. Pandey, N.; Tripathi, M.; Gupta, M.K.; Tilak, R. Overexpression of efflux pump transporter genes and mutations in ERG11 pave the way to fluconazole resistance in Candida tropicalis: A study from a North India region. J. Glob. Antimicrob. Resist. 2020, 22, 374–378. [Google Scholar] [CrossRef] [PubMed]
  109. Park, Y.-D.; Sun, W.; Salas, A.; Antia, A.; Carvajal, C.; Wang, A.; Xu, X.; Meng, Z.; Zhou, M.; Tawa, G.J.; et al. Identification of multiple cryptococcal fungicidal drug targets by combined gene dosing and drug affinity responsive target stability screening. mBio 2016, 7. [Google Scholar] [CrossRef] [Green Version]
  110. Truong, M.; Monahan, L.G.; Carter, D.A.; Charles, I.G. Repurposing drugs to fast-track therapeutic agents for the treatment of cryptococcosis. PeerJ 2018, 6, e4761. [Google Scholar] [CrossRef] [Green Version]
  111. Afeltra, J.; Vitale, R.G.; Mouton, J.W.; Verweij, P.E. Potent synergistic in vitro interaction between nonantimicrobial membrane-active compounds and itraconazole against clinical isolates of Aspergillus fumigatus resistant to itraconazole. Antimicrob. Agents Chemother. 2004, 48, 1335–1343. [Google Scholar] [CrossRef] [Green Version]
  112. Alnajjar, L.M.; Bulatova, N.R.; Darwish, R.M. Evaluation of four calcium channel blockers as fluconazole resistance inhibitors in Candida glabrata. J. Glob. Antimicrob. Resist. 2018, 14, 185–189. [Google Scholar] [CrossRef]
  113. Eldesouky, H.E.; Mayhoub, A.; Hazbun, T.R.; Seleem, M.N. Reversal of azole resistance in Candida albicans by sulfa antibacterial drugs. Antimicrob. Agents Chemother. 2018, 62, e00701–e00717. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  114. Liu, S.; Yue, L.; Gu, W.; Li, X.; Zhang, L.; Sun, S. Synergistic effect of fluconazole and calcium channel blockers against resistant Candida albicans. PLoS ONE 2016, 11, e0150859. [Google Scholar] [CrossRef] [PubMed]
  115. Rodrigues, A.G.; Araujo, R.; Pina-Vaz, C. Interaction of local anaesthetics with other antifungal agents against pathogenic Aspergillus. Int. J. Antimicrob. Agents 2006, 27, 339–343. [Google Scholar] [CrossRef] [PubMed]
  116. Brilhante, R.S.N.; Caetano, E.P.; Lima, R.A.C.; Castelo Branco, D.S.C.M.; Serpa, R.; Oliveira, J.S.; Monteiro, A.J.; Rocha, M.F.G.; Cordeiro, R.A.; Sidrim, J.J.C. In vitro antifungal activity of miltefosine and levamisole: Their impact on ergosterol biosynthesis and cell permeability of dimorphic fungi. J. Appl. Microbiol. 2015, 119, 962–969. [Google Scholar] [CrossRef]
  117. Brilhante, R.S.N.; Silva, M.L.Q.D.; Pereira, V.S.; de Oliveira, J.S.; Maciel, J.M.; Silva, I.N.G.D.; Garcia, L.G.S.; Guedes, G.M.D.M.; Cordeiro, R.D.A.; Pereira-Neto, W.D.A.; et al. Potassium iodide and miltefosine inhibit biofilms of Sporothrix schenckii species complex in yeast and filamentous forms. Med. Mycol. 2018, 57, 764–772. [Google Scholar] [CrossRef] [PubMed]
  118. Spadari, C.d.C.; Vila, T.; Rozental, S.; Ishida, K. Miltefosine has a postantifungal effect and induces apoptosis in Cryptococcus yeasts. Antimicrob. Agents Chemother. 2018, 62, e00312–e00318. [Google Scholar] [CrossRef] [Green Version]
  119. Wall, G.; Herrera, N.; Lopez-Ribot, J.L. Repositionable compounds with antifungal activity against multidrug resistant Candida auris identified in the medicines for malaria venture’s pathogen box. J. Fungi 2019, 5, 92. [Google Scholar] [CrossRef] [Green Version]
  120. Wu, Y.; Grossman, N.; Totten, M.; Memon, W.; Fitzgerald, A.; Ying, C.; Zhang, S.X. Antifungal susceptibility profiles and drug resistance mechanisms of clinical Lomentospora prolificans isolates. Antimicrob. Agents Chemother. 2020, 64, e00318–e00320. [Google Scholar] [CrossRef]
  121. Bulatova, N.R.; Darwish, R.M. Effect of chemosensitizers on minimum inhibitory concentrations of fluconazole in Candida albicans. Med. Prince Pract. 2008, 17, 117–121. [Google Scholar] [CrossRef]
  122. Harrison, T.S.; Griffin, G.E.; Levitz, S.M. Conditional lethality of the diprotic weak bases chloroquine and quinacrine against Cryptococcus neoformans. J. Infect. Dis. 2000, 182, 283–289. [Google Scholar] [CrossRef] [Green Version]
  123. Kulkarny, V.V.; Chavez-Dozal, A.; Rane, H.S.; Jahng, M.; Bernardo, S.M.; Parra, K.J.; Lee, S.A. Quinacrine inhibits Candida albicans growth and filamentation at neutral pH. Antimicrob. Agents Chemother. 2014, 58, 7501–7509. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  124. Siles, S.A.; Srinivasan, A.; Pierce, C.G.; Lopez-Ribot, J.L.; Ramasubramanian, A.K. High-throughput screening of a collection of known pharmacologically active small compounds for identification of Candida albicans biofilm inhibitors. Antimicrob. Agents Chemother. 2013, 57, 3681–3687. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  125. Vallières, C.; Raulo, R.; Dickinson, M.; Avery, S.V. Novel combinations of agents targeting translation that synergistically inhibit fungal pathogens. Front. Microbiol. 2018, 9. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  126. Zani, C.; Bisceglie, F.; Restivo, F.M.; Feretti, D.; Pioli, M.; Degola, F.; Montalbano, S.; Galati, S.; Pelosi, G.; Viola, G.V.C.; et al. A battery of assays as an integrated approach to evaluate fungal and mycotoxin inhibition properties and cytotoxic/genotoxic side-effects for the prioritization in the screening of thiosemicarbazone derivatives. Food Chem. Toxicol. 2017, 105, 498–505. [Google Scholar] [CrossRef]
  127. Sun, N.; Li, D.; Zhang, Y.; Killeen, K.; Groutas, W.; Calderone, R. Repurposing an inhibitor of ribosomal biogenesis with broad anti-fungal activity. Sci. Rep. 2017, 7, 17014. [Google Scholar] [CrossRef]
  128. Rogolino, D.; Gatti, A.; Carcelli, M.; Pelosi, G.; Bisceglie, F.; Restivo, F.M.; Degola, F.; Buschini, A.; Montalbano, S.; Feretti, D.; et al. Thiosemicarbazone scaffold for the design of antifungal and antiaflatoxigenic agents: Evaluation of ligands and related copper complexes. Sci. Rep. 2017, 7, 11214. [Google Scholar] [CrossRef]
  129. Degola, F.; Morcia, C.; Bisceglie, F.; Mussi, F.; Tumino, G.; Ghizzoni, R.; Pelosi, G.; Terzi, V.; Buschini, A.; Restivo, F.M.; et al. In vitro evaluation of the activity of thiosemicarbazone derivatives against mycotoxigenic fungi affecting cereals. Int. J. Food Microbiol. 2015, 200, 104–111. [Google Scholar] [CrossRef]
  130. De Sá, N.P.; Cisalpino, P.S.; Bertollo, C.M.; Santos, P.C.; Rosa, C.A.; de Souza, D.D.G.; Barbeira, P.J.S.; Alves, T.M.D.A.; Zani, C.L.; Johann, S. Thiosemicarbazone of lapachol acts on cell membrane in Paracoccidioides brasiliensis. Med. Mycol. 2018, 57, 332–339. [Google Scholar] [CrossRef]
  131. Schmidt, S.; Hogardt, M.; Demir, A.; Röger, F.; Lehrnbecher, T. Immunosuppressive compounds affect the fungal growth and viability of defined Aspergillus species. Pathogens 2019, 8, 273. [Google Scholar] [CrossRef] [Green Version]
  132. Noto, T.; Sawada, M.; Ando, K.; Koyama, K. Some biological properties of mycophenolic acid. J. Antibiot. 1969, 22, 165–169. [Google Scholar] [CrossRef] [Green Version]
  133. Kim, K.; Zilbermintz, L.; Martchenko, M. Repurposing FDA approved drugs against the human fungal pathogen, Candida albicans. Ann. Clin. Microbiol. Antimicrob. 2015, 14, 32. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  134. Bartoli, J.; Montalbano, S.; Spadola, G.; Rogolino, D.; Pelosi, G.; Bisceglie, F.; Restivo, F.M.; Degola, F.; Serra, O.; Buschini, A.; et al. Antiaflatoxigenic thiosemicarbazones as crop-protective agents: A cytotoxic and genotoxic study. J. Agric. Food Chem. 2019, 67, 10947–10953. [Google Scholar] [CrossRef] [PubMed]
  135. Banerjee, D.; Burkard, L.; Panepinto, J.C. Inhibition of nucleotide biosynthesis potentiates the antifungal activity of amphotericin B. PLoS ONE 2014, 9, e87246. [Google Scholar] [CrossRef] [PubMed]
  136. Zanette, R.A.; Jesus, F.P.K.; Pilotto, M.B.; Weiblen, C.; Pötter, L.; Ferreiro, L.; Alves, S.H.; Santurio, J.M. Micafungin alone and in combination therapy with deferasirox against Pythium insidiosum. J. Mycol. Med. 2015, 25, 91–94. [Google Scholar] [CrossRef] [PubMed]
  137. Puri, S.; Kumar, R.; Rojas, I.G.; Salvatori, O.; Edgerton, M. Iron chelator deferasirox reduces Candida albicans invasion of oral epithelial cells and infection levels in murine oropharyngeal Candidiasis. Antimicrob. Agents Chemother. 2019, 63, e02152-18. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  138. Magalhães, T.F.F.; Costa, M.C.; Holanda, R.A.; Ferreira, G.F.; Carvalho, V.S.D.; Freitas, G.J.C.; Ribeiro, N.Q.; Emídio, E.C.P.; Carmo, P.H.F.; de Brito, C.B.; et al. N-acetylcysteine reduces amphotericin B deoxycholate nephrotoxicity and improves the outcome of murine cryptococcosis. Med. Mycol. 2020, 58, 835–844. [Google Scholar] [CrossRef]
  139. Lewis, R.E.; Albert, N.D.; Kontoyiannis, D.P. Comparative pharmacodynamics of posaconazole in neutropenic murine models of invasive pulmonary aspergillosis and mucormycosis. Antimicrob. Agents Chemother. 2014, 58, 6767–6772. [Google Scholar] [CrossRef] [Green Version]
  140. Homa, M.; Galgóczy, L.; Tóth, E.; Tóth, L.; Papp, T.; Chandrasekaran, M.; Kadaikunnan, S.; Alharbi, N.S.; Vágvölgyi, C. In vitro antifungal activity of antipsychotic drugs and their combinations with conventional antifungals against Scedosporium and Pseudallescheria isolates. Med. Mycol. 2015, 53, 890–895. [Google Scholar] [CrossRef] [Green Version]
  141. You, Z.; Ran, X.; Dai, Y.; Ran, Y. Clioquinol, an alternative antimicrobial agent against common pathogenic microbe. J. Mycol. Med. 2018, 28, 492–501. [Google Scholar] [CrossRef]
  142. Thangamani, S.; Eldesouky, H.E.; Mohammad, H.; Pascuzzi, P.E.; Avramova, L.; Hazbun, T.R.; Seleem, M.N. Ebselen exerts antifungal activity by regulating glutathione (GSH) and reactive oxygen species (ROS) production in fungal cells. Biochim. Biophys. Acta Gen. Subj. 2017, 1861, 3002–3010. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  143. Su, S.; Shi, X.; Xu, W.; Li, Y.; Chen, X.; Jia, S.; Sun, S. Antifungal activity and potential mechanism of panobinostat in combination with fluconazole against Candida albicans. Front. Microbiol. 2020, 11. [Google Scholar] [CrossRef] [PubMed]
  144. Leonardelli, F.; Macedo, D.; Dudiuk, C.; Theill, L.; Cabeza, M.S.; Gamarra, S.; Garcia-Effron, G. In Vitro activity of combinations of zinc chelators with amphotericin b and posaconazole against six mucorales species. Antimicrob. Agents Chemother. 2019, 63. [Google Scholar] [CrossRef] [Green Version]
  145. Jaromin, A.; Zarnowski, R.; Piętka-Ottlik, M.; Andes, D.R.; Gubernator, J. Topical delivery of ebselen encapsulated in biopolymeric nanocapsules: Drug repurposing enhanced antifungal activity. Nanomedicine 2018, 13, 1139–1155. [Google Scholar] [CrossRef] [Green Version]
  146. Ben-Gigi, G.; Polacheck, I.; Eilam, Y. In vitro synergistic activity of ketoconazole with trifluoperazine and with chlorpromazine against medically important yeasts. Chemotherapy 1988, 34, 96–100. [Google Scholar] [CrossRef]
  147. Butts, A.; Palmer, G.E.; Rogers, P.D. Antifungal adjuvants: Preserving and extending the antifungal arsenal. Virulence 2017, 8, 198–210. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  148. Jung, E.H.; Meyers, D.J.; Bosch, J.; Casadevall, A. Novel antifungal compounds discovered in medicines for malaria venture’s malaria box. mSphere 2018, 3. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  149. Montoya, M.C.; DiDone, L.; Heier, R.F.; Meyers, M.J.; Krysan, D.J. Antifungal Phenothiazines: Optimization, characterization of mechanism, and modulation of neuroreceptor activity. ACS Infect. Dis. 2018, 4, 499–507. [Google Scholar] [CrossRef] [PubMed]
  150. Siavoshi, F.; Tavakolian, A.; Foroumadi, A.; Hosseini, N.M.; Massarrat, S.; Pedramnia, S.; Saniee, P. Comparison of the effect of non-antifungal and antifungal agents on Candida isolates from the gastrointestinal tract. Arch. Iran. Med. 2012, 15, 27–31. [Google Scholar]
  151. Singh, S.; Fatima, Z.; Ahmad, K.; Hameed, S. Repurposing of respiratory drug theophylline against Candida albicans: Mechanistic insights unveil alterations in membrane properties and metabolic fitness. J. Appl. Microbiol. 2020. [Google Scholar] [CrossRef]
  152. Rajasekharan, S.K.; Lee, J.-H.; Lee, J. Aripiprazole repurposed as an inhibitor of biofilm formation and sterol biosynthesis in multidrug-resistant Candida albicans. Int. J. Antimicrob. Agents 2019, 54, 518–523. [Google Scholar] [CrossRef]
  153. Pic, E.; Burgain, A.; Sellam, A. Repurposing the anthelminthic salicylanilide oxyclozanide against susceptible and clinical resistant Candida albicans strains. Med. Mycol. 2018, 57, 387–390. [Google Scholar] [CrossRef] [PubMed]
  154. Nile, C.; Falleni, M.; Cirasola, D.; Alghamdi, A.; Anderson, O.F.; Delaney, C.; Ramage, G.; Ottaviano, E.; Tosi, D.; Bulfamante, G.; et al. Repurposing pilocarpine hydrochloride for treatment of Candida albicans infections. mSphere 2019, 4. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  155. Montoya, M.C.; Beattie, S.; Alden, K.M.; Krysan, D.J. Derivatives of the antimalarial drug mefloquine are broad-spectrum antifungal molecules with activity against drug-resistant clinical isolates. Antimicrob. Agents Chemother. 2020, 64. [Google Scholar] [CrossRef] [PubMed]
  156. Fonseca, D.; Leal-Pinto, S.M.; Roa-Cordero, M.V.; Vargas, J.D.; Moreno-Moreno, E.M.; Macías, M.A.; Suescun, L.; Muñoz-Castro, Á.; Hurtado, J.J. Inhibition of C. albicans dimorphic switch by Cobalt(II) complexes with ligands derived from pyrazoles and dinitrobenzoate: Synthesis, characterization and biological activity. Int. J. Mol. Sci. 2019, 20, 3237. [Google Scholar] [CrossRef] [Green Version]
  157. Joffe, L.S.; Schneider, R.; Lopes, W.; Azevedo, R.; Staats, C.C.; Kmetzsch, L.; Schrank, A.; Del Poeta, M.; Vainstein, M.H.; Rodrigues, M.L. The anti-helminthic compound mebendazole has multiple antifungal effects against Cryptococcus neoformans. Front. Microbiol. 2017, 8. [Google Scholar] [CrossRef] [PubMed]
  158. Ma, Y.; Liang, S.; Zhang, Y.; Yang, D.; Wang, R. Development of anti-fungal pesticides from protein kinase inhibitor-based anticancer agents. Eur. J. Med. Chem. 2018, 148, 349–358. [Google Scholar] [CrossRef]
  159. Ngo, H.X.; Shrestha, S.K.; Garneau-Tsodikova, S. Identification of ebsulfur analogues with broad-spectrum antifungal activity. ChemMedChem 2016, 11, 1507–1516. [Google Scholar] [CrossRef] [Green Version]
  160. Veri, A.; Cowen, L.E. Progress and prospects for targeting Hsp90 to treat fungal infections. Parasitology 2014, 141, 1127–1137. [Google Scholar] [CrossRef]
  161. She, P.; Liu, Y.; Wang, Y.; Tan, F.; Luo, Z.; Wu, Y. Antibiofilm efficacy of the gold compound auranofin on dual species biofilms of Staphylococcus aureus and Candida sp. J. Appl. Microbiol. 2020, 128, 88–101. [Google Scholar] [CrossRef] [Green Version]
  162. Nobile, C.J.; Ennis, C.L.; Hartooni, N.; Johnson, A.D.; Lohse, M.B. A selective serotonin reuptake inhibitor, a proton pump inhibitor, and two calcium channel blockers inhibit Candida albicans biofilms. Microorganisms 2020, 8, 756. [Google Scholar] [CrossRef]
  163. Eldesouky, H.E.; Salama, E.A.; Lanman, N.A.; Hazbun, T.R.; Seleem, M.N. Potent synergistic interactions between lopinavir and azole antifungal drugs against emerging multidrug-resistant Candida auris. Antimicrob. Agents Chemother. 2020. [Google Scholar] [CrossRef] [PubMed]
  164. Chavez-Dozal, A.A.; Lown, L.; Jahng, M.; Walraven, C.J.; Lee, S.A. In vitro analysis of finasteride activity against Candida albicans urinary biofilm formation and filamentation. Antimicrob. Agents Chemother. 2014, 58, 5855–5862. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  165. Bernardo, S.M.; Allen, C.P.; Waller, A.; Young, S.M.; Oprea, T.; Sklar, L.A.; Lee, S.A. An automated high-throughput cell-based multiplexed flow cytometry assay to identify novel compounds to target Candida albicans virulence-related proteins. PLoS ONE 2014, 9, e110354. [Google Scholar] [CrossRef] [PubMed]
  166. Meng, L.-N.; Liu, J.-Y.; Wang, Y.-T.; Ni, S.-S.; Xiang, M.-J. The discovery of potential phosphopantetheinyl transferase Ppt2 inhibitors against drug-resistant Candida albicans. Braz. J. Microbiol. 2020. [Google Scholar] [CrossRef]
  167. Mei, Y.; Jiang, T.; Zou, Y.; Wang, Y.; Zhou, J.; Li, J.; Liu, L.; Tan, J.; Wei, L.; Li, J.; et al. FDA approved drug library screening identifies robenidine as a repositionable antifungal. Front. Microbiol. 2020, 11. [Google Scholar] [CrossRef]
  168. Li, Z.; Fu, B.; Green, C.M.; Liu, B.; Zhang, J.; Lang, Y.; Chaturvedi, S.; Belfort, M.; Liao, G.; Li, H. Cisplatin protects mice from challenge of Cryptococcus neoformans by targeting the Prp8 intein. Emerg. Microbes Infect. 2019, 8, 895–908. [Google Scholar] [CrossRef] [Green Version]
  169. Garcia, C.; Burgain, A.; Chaillot, J.; Pic, É.; Khemiri, I.; Sellam, A. A phenotypic small-molecule screen identifies halogenated salicylanilides as inhibitors of fungal morphogenesis, biofilm formation and host cell invasion. Sci. Rep. 2018, 8, 11559. [Google Scholar] [CrossRef] [Green Version]
  170. Dehdashti, S.J.; Abbott, J.; Nguyen, D.T.; McKew, J.C.; Williamson, P.R.; Zheng, W. A high-throughput screening assay for assessing the viability of Cryptococcus neoformans under nutrient starvation conditions. Anal. Bioanl. Chem. 2013, 405, 6823–6829. [Google Scholar] [CrossRef] [Green Version]
  171. Clinical and Laboratory Standards Institute (CLSI). Reference Method for Broth Dilution Antifungal Susceptibility Testing of Yeasts: Approved Standard, 2nd ed.; CLSI document M27-A3; Clinical and Laboratory Standards Institute: Wayne, PA, USA, 2008. [Google Scholar]
  172. The European Committee on Antimicrobial Susceptibility Testing. EUCAST Method for Susceptibility Testing of Yeasts (v 7.3.2 Valid from 22 April 2020). Available online: https://www.eucast.org/astoffungi/methodsinantifungalsusceptibilitytesting/susceptibility_testing_of_yeasts/ (accessed on 21 August 2020).
  173. Law, G.L.; Tisoncik-Go, J.; Korth, M.J.; Katze, M.G. Drug repurposing: A better approach for infectious disease drug discovery? Curr. Opin. Immunol. 2013, 25, 588–592. [Google Scholar] [CrossRef]
  174. Abadio, A.K.R.; Kioshima, E.S.; Teixeira, M.M.; Martins, N.F.; Maigret, B.; Felipe, M.S.S. Comparative genomics allowed the identification of drug targets against human fungal pathogens. BMC Genom. 2011, 12, 75. [Google Scholar] [CrossRef] [Green Version]
  175. Gowri, M.; Sofi Beaula, W.; Biswal, J.; Dhamodharan, P.; Saiharish, R.; Rohan prasad, S.; Pitani, R.; Kandaswamy, D.; Raghunathan, R.; Jeyakanthan, J.; et al. β-lactam substituted polycyclic fused pyrrolidine/pyrrolizidine derivatives eradicate C. albicans in an ex vivo human dentinal tubule model by inhibiting sterol 14-α demethylase and cAMP pathway. Biochim. Biophys. Acta 2016, 1860, 636–647. [Google Scholar] [CrossRef] [PubMed]
  176. OpenEye Scientific Software OMEGA; v.2.5.1; OpenEye Scientific Software: Santa Fe, NM, USA, 2019.
  177. OpenEye Scientific Software QUACPAC; v.1.6.3; OpenEye Scientific Software: Santa Fe, NM, USA, 2013.
  178. OpenEye Scientific Software OEDocking; v3.2.0; OpenEye Scientific Software: Santa Fe, NM, USA, 2017.
  179. Anandakrishnan, R.; Aguilar, B.; Onufriev, A.V. H++ 3.0: Automating pK prediction and the preparation of biomolecular structures for atomistic molecular modeling and simulations. Nucleic Acids Res. 2012, 40, W537–W541. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  180. Biasini, M.; Bienert, S.; Waterhouse, A.; Arnold, K.; Studer, G.; Schmidt, T.; Kiefer, F.; Cassarino, T.G.; Bertoni, M.; Bordoli, L.; et al. SWISS-MODEL: Modelling protein tertiary and quaternary structure using evolutionary information. Nucleic Acids Res. 2014, 42, W252–W258. [Google Scholar] [CrossRef] [PubMed]
  181. Bordoli, L.; Kiefer, F.; Arnold, K.; Benkert, P.; Battey, J.; Schwede, T. Protein structure homology modeling using SWISS-MODEL workspace. Nat. Protoc. 2009, 4, 1–13. [Google Scholar] [CrossRef]
  182. Broad Institute. Fungal Genomics. Available online: https://www.broadinstitute.org/fungal-genome-initiative (accessed on 21 August 2020).
  183. Chen, V.B.; Arendall, W.B., III; Headd, J.J.; Keedy, D.A.; Immormino, R.M.; Kapral, G.J.; Murray, L.W.; Richardson, J.S.; Richardson, D.C. MolProbity: All-atom structure validation for macromolecular crystallography. Acta Crystallograph. Sect. D 2010, 66, 12–21. [Google Scholar] [CrossRef] [Green Version]
  184. De Paula e Silva, A.C.A.; Oliveira, H.C.; Silva, J.F.; Sangalli-Leite, F.; Scorzoni, L.; Fusco-Almeida, A.M.; Mendes-Giannini, M.J.S. Microplate alamarBlue assay for Paracoccidioides susceptibility testing. J. Clin. Microbiol. 2013, 51, 1250–1252. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  185. Law, V.; Knox, C.; Djoumbou, Y.; Jewison, T.; Guo, A.C.; Liu, Y.; Maciejewski, A.; Arndt, D.; Wilson, M.; Neveu, V.; et al. DrugBank 4.0: Shedding new light on drug metabolism. Nucleic Acids Res. 2013, 42, D1091–D1097. [Google Scholar] [CrossRef] [Green Version]
  186. Li, Y.H.; Yu, C.Y.; Li, X.X.; Zhang, P.; Tang, J.; Yang, Q.; Fu, T.; Zhang, X.; Cui, X.; Tu, G.; et al. Therapeutic target database update 2018: Enriched resource for facilitating bench-to-clinic research of targeted therapeutics. Nucleic Acids Res. 2017, 46, D1121–D1127. [Google Scholar] [CrossRef]
  187. McGann, M. FRED and HYBRID docking performance on standardized datasets. J. Comput.-Aided Mol. Des. 2012, 26, 897–906. [Google Scholar] [CrossRef]
  188. Rodrigues, J.P.G.L.M.; Levitt, M.; Chopra, G. KoBaMIN: A knowledge-based minimization web server for protein structure refinement. Nucleic Acids Res. 2012, 40, W323–W328. [Google Scholar] [CrossRef] [Green Version]
  189. Rose, P.W.; Prlić, A.; Bi, C.; Bluhm, W.F.; Christie, C.H.; Dutta, S.; Green, R.K.; Goodsell, D.S.; Westbrook, J.D.; Woo, J.; et al. The RCSB Protein Data Bank: Views of structural biology for basic and applied research and education. Nucleic Acids Res. 2014, 43, D345–D356. [Google Scholar] [CrossRef] [PubMed]
  190. Wang, Y.; Coleman-Derr, D.; Chen, G.; Gu, Y.Q. OrthoVenn: A web server for genome wide comparison and annotation of orthologous clusters across multiple species. Nucleic Acids Res. 2015, 43, W78–W84. [Google Scholar] [CrossRef] [PubMed]
  191. Zhang, R.; Lin, Y. DEG 5.0, a database of essential genes in both prokaryotes and eukaryotes. Nucleic Acids Res. 2008, 37, D455–D458. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  192. Zhang, R.; Ou, H.Y.; Zhang, C.T. DEG: A database of essential genes. Nucleic Acids Res. 2004, 32, D271–D272. [Google Scholar] [CrossRef] [PubMed]
  193. European Bioinformatics Institute. CLUSTALW. Available online: http://www.ebi.ac.uk/Tools/msa/ (accessed on 21 August 2020).
  194. Hess, B.; Kutzner, C.; van der Spoel, D.; Lindahl, E. GROMACS 4:  Algorithms for highly efficient, load-balanced, and scalable molecular simulation. J. Chem. Theory Comput. 2008, 4, 435–447. [Google Scholar] [CrossRef] [Green Version]
  195. National Center for Biotechnology Information. BLAST. Available online: https://www.ncbi.nlm.nih.gov/BLAST/ (accessed on 21 August 2020).
  196. Amiri, S.; Sansom, M.S.P.; Biggin, P.C. Molecular dynamics studies of AChBP with nicotine and carbamylcholine: The role of water in the binding pocket. Protein Eng. Des. Sel. 2007, 20, 353–359. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  197. Clancy, C.J.; Nguyen, M.H. Comparison of a photometric method with standardized methods of antifungal susceptibility testing of yeasts. J. Clin. Microbiol. 1997, 35, 2878–2882. [Google Scholar] [CrossRef] [Green Version]
  198. Humphrey, W.; Dalke, A.; Schulten, K. VMD: Visual molecular dynamics. J. Mol. Graph. 1996, 14, 33–38. [Google Scholar] [CrossRef]
  199. Molecular Operating Environment (MOE); v.2013.08; Chemical Computing Group, Inc.: Montreal, QC, Canada, 2014.
  200. Oostenbrink, C.; Villa, A.; Mark, A.E.; Van Gunsteren, W.F. A biomolecular force field based on the free enthalpy of hydration and solvation: The GROMOS force-field parameter sets 53A5 and 53A6. J. Comput. Chem. 2004, 25, 1656–1676. [Google Scholar] [CrossRef]
  201. Pronk, S.; Páll, S.; Schulz, R.; Larsson, P.; Bjelkmar, P.; Apostolov, R.; Shirts, M.R.; Smith, J.C.; Kasson, P.M.; van der Spoel, D.; et al. GROMACS 4.5: A high-throughput and highly parallel open source molecular simulation toolkit. Bioinformatics 2013, 29, 845–854. [Google Scholar] [CrossRef]
  202. Rarey, M.; Kramer, B.; Lengauer, T.; Klebe, G. A fast flexible docking method using an incremental construction algorithm. J. Mol. Biol. 1996, 261, 470–489. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  203. Schüttelkopf, A.W.; van Aalten, D.M. PRODRG: A tool for high-throughput crystallography of protein-ligand complexes. Acta Crystallogr. D Biol. Crystallogr. 2004, 60, 1355–1363. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  204. Wang, H.; Dommert, F.; Holm, C. Optimizing working parameters of the smooth particle mesh Ewald algorithm in terms of accuracy and efficiency. J. Chem. Phys. 2010, 133, 034117. [Google Scholar] [CrossRef] [PubMed]
  205. Dallakyan, S.; Olson, A.J. Small-molecule library screening by docking with PyRx. Methods Mol. Biol. 2015, 1263, 243–250. [Google Scholar] [CrossRef]
  206. Morris, G.M.; Huey, R.; Lindstrom, W.; Sanner, M.F.; Belew, R.K.; Goodsell, D.S.; Olson, A.J. AutoDock4 and AutoDockTools4: Automated docking with selective receptor flexibility. J. Comput. Chem. 2009, 30, 2785–2791. [Google Scholar] [CrossRef] [Green Version]
  207. Gasteiger, J.; Rudolph, C.; Sadowski, J. Automatic generation of 3D-atomic coordinates for organic molecules. Tetrahedron. Comput. Methodol. 1990, 3, 537–547. [Google Scholar] [CrossRef]
  208. Jones, G.; Willett, P.; Glen, R.C.; Leach, A.R.; Taylor, R. Development and validation of a genetic algorithm for flexible docking. J. Mol. Biol. 1997, 267, 727–748. [Google Scholar] [CrossRef] [Green Version]
  209. Schäfer, T.; Kriege, N.; Humbeck, L.; Klein, K.; Koch, O.; Mutzel, P. Scaffold hunter: A comprehensive visual analytics framework for drug discovery. J. Cheminform. 2017, 9, 28. [Google Scholar] [CrossRef] [Green Version]
  210. Wallace, A.C.; Laskowski, R.A.; Thornton, J.M. LIGPLOT: A program to generate schematic diagrams of protein-ligand interactions. Protein Eng. 1995, 8, 127–134. [Google Scholar] [CrossRef]
  211. Meletiadis, J.; Antachopoulos, C.; Stergiopoulou, T.; Pournaras, S.; Roilides, E.; Walsh, T.J. Differential fungicidal activities of amphotericin B and voriconazole against Aspergillus species determined by microbroth methodology. Antimicrob. Agents Chemother. 2007, 51, 3329–3337. [Google Scholar] [CrossRef] [Green Version]
  212. Odds, F.C. Synergy, antagonism, and what the chequerboard puts between them. J. Antimicrob. Chemother. 2003, 52. [Google Scholar] [CrossRef] [PubMed]
  213. Alastruey-Izquierdo, A.; Melhem, M.S.; Bonfietti, L.X.; Rodriguez-Tudela, J.L. Susceptiblity test for fungi: Clinical and laboratorial correlations in medical mycology. Rev. Inst. Med. Trop São Paulo 2015, 57 (Suppl. 19), 57–64. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  214. Wu, S.; Wang, Y.; Liu, N.; Dong, G.; Sheng, C. Tackling fungal resistance by biofilm inhibitors. J. Med. Chem. 2017, 60, 2193–2211. [Google Scholar] [CrossRef] [PubMed]
  215. Sherry, L.; Rajendran, R.; Lappin, D.F.; Borghi, E.; Perdoni, F.; Falleni, M.; Tosi, D.; Smith, K.; Williams, C.; Jones, B.; et al. Biofilms formed by Candida albicans bloodstream isolates display phenotypic and transcriptional heterogeneity that are associated with resistance and pathogenicity. BMC Microbiol. 2014, 14, 182. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  216. Desai, J.V.; Mitchell, A.P.; Andes, D.R. Fungal biofilms, drug resistance, and recurrent infection. Cold Spring Harb. Perspect. Med. 2014, 4. [Google Scholar] [CrossRef] [Green Version]
  217. Gulati, M.; Lohse, M.B.; Ennis, C.L.; Gonzalez, R.E.; Perry, A.M.; Bapat, P.; Arevalo, A.V.; Rodriguez, D.L.; Nobile, C.J. In Vitro culturing and screening of Candida albicans biofilms. Curr. Prot. Microbiol. 2018, 50, e60. [Google Scholar] [CrossRef]
  218. Pierce, C.G.; Srinivasan, A.; Uppuluri, P.; Ramasubramanian, A.K.; López-Ribot, J.L. Antifungal therapy with an emphasis on biofilms. Curr. Opin. Pharmacol. 2013, 13, 726–730. [Google Scholar] [CrossRef] [Green Version]
  219. Sherry, L.; Ramage, G.; Kean, R.; Borman, A.; Johnson, E.; Richardson, M.; Rautemaa-Richardson, R. Biofilm-Forming capability of highly virulent, multidrug-resistant Candida auris. Emerg. Infect. Dis. 2017, 23, 328. [Google Scholar] [CrossRef] [Green Version]
  220. Ha, K.C.; White, T.C. Effects of azole antifungal drugs on the transition from yeast cells to hyphae in susceptible and resistant isolates of the pathogenic yeast Candida albicans. Antimicrob. Agents Chemother. 1999, 43, 763–768. [Google Scholar] [CrossRef] [Green Version]
  221. Heilman, B.J.; Tadle, A.C.; Pimentel, L.R.; Mascharak, P.K. Selective damage to hyphal form through light-induced delivery of nitric oxide to Candida albicans colonies. J. Inorg. Biochem. 2013, 123, 18–22. [Google Scholar] [CrossRef]
  222. Higgins, J.; Pinjon, E.; Oltean, H.N.; White, T.C.; Kelly, S.L.; Martel, C.M.; Sullivan, D.J.; Coleman, D.C.; Moran, G.P. Triclosan antagonizes fluconazole activity against Candida albicans. J. Dent. Res. 2012, 91, 65–70. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  223. Van de Sande, W.W.J.; Tavakol, M.; van Vianen, W.; Bakker-Woudenberg, I.A.J.M. The effects of antifungal agents to conidial and hyphal forms of Aspergillus fumigatus. Med. Mycol. 2010, 48, 48–55. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  224. Mariné, M.; Bom, V.L.P.; de Castro, P.A.; Winkelstroter, L.K.; Ramalho, L.N.; Brown, N.A.; Goldman, G.H. The development of animal infection models and antifungal efficacy assays against clinical isolates of Trichosporon asahii, T. asteroides and T. inkin. Virulence 2015, 6, 476–486. [Google Scholar] [CrossRef]
  225. Paulussen, C.; Boulet, G.A.V.; Cos, P.; Delputte, P.; Maes, L.J.R.M. Animal models of invasive aspergillosis for drug discovery. Drug Discov. Today 2014, 19, 1380–1386. [Google Scholar] [CrossRef] [PubMed]
  226. National Research Council (US) Committee to Update Science, Medicine, and Animals. Regulation of animal research. In Science, Medicine, and Animals; National Academies Press: Washington, DC, USA, 2004. Available online: https://www.ncbi.nlm.nih.gov/books/NBK24650/ (accessed on 14 November 2020).
  227. Arvanitis, M.; Glavis-Bloom, J.; Mylonakis, E. Invertebrate models of fungal infection. Biochim. Biophys. Acta Mol. Basis Dis. 2013, 1832, 1378–1383. [Google Scholar] [CrossRef] [Green Version]
  228. De Oliveira, H.C.; Monteiro, M.C.; Rossi, S.A.; Pemán, J.; Ruiz-Gaitán, A.; Mendes-Giannini, M.J.S.; Mellado, E.; Zaragoza, O. Identification of off-patent compounds that present antifungal activity against the emerging fungal pathogen Candida auris. Front. Cell. Infect. Microbiol. 2019, 9, 83. [Google Scholar] [CrossRef] [Green Version]
  229. Gamo, F.-J.; Sanz, L.M.; Vidal, J.; de Cozar, C.; Alvarez, E.; Lavandera, J.-L.; Vanderwall, D.E.; Green, D.V.S.; Kumar, V.; Hasan, S.; et al. Thousands of chemical starting points for antimalarial lead identification. Nature 2010, 465, 305–310. [Google Scholar] [CrossRef]
  230. Johnson, M.D.; Perfect, J.R. Use of antifungal combination therapy: Agents, order, and timing. Curr. Fungal Infect. Rep. 2010, 4, 87–95. [Google Scholar] [CrossRef] [Green Version]
  231. Carrillo-Muñoz, A.J.; Finquelievich, J.; Tur-Tur, C.; Eraso, E.; Jauregizar, N.; Quindós, G.; Giusiano, G. Combination antifungal therapy: A strategy for the management of invasive fungal infections. Rev. Esp. Quim. 2014, 27, 141–158. [Google Scholar]
  232. Kim, J.H.; Haff, R.P.; Faria, N.C.G.; Martins, M.D.L.; Chan, K.L.; Campbell, B.C. Targeting the mitochondrial respiratory chain of Cryptococcus through antifungal chemosensitization: A model for control of non-fermentative pathogens. Molecules 2013, 18, 8873–8894. [Google Scholar] [CrossRef]
  233. Campbell, B.C.; Chan, K.L.; Kim, J.H. Chemosensitization as a means to augment commercial antifungal agents. Front. Microbiol. 2012, 3, 79. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  234. Kim, J.H.; Chan, K.L.; Cheng, L.W. Octyl gallate as an intervention catalyst to augment antifungal efficacy of caspofungin. J—Multidiscip. Sci. J. 2018, 1, 19–28. [Google Scholar] [CrossRef] [Green Version]
  235. AHFS® Patient Medication Information™. Pioglitazone. Available online: https://medlineplus.gov/druginfo/meds/a699016.html#why (accessed on 3 September 2020).
  236. Ribeiro, N.Q.; Santos, A.P.N.; Emídio, E.C.P.; Costa, M.C.; Freitas, G.J.C.; Carmo, P.H.F.; Silva, M.F.; de Brito, C.B.; de Souza, D.G.; Paixão, T.A.; et al. Pioglitazone as an adjuvant of amphotericin B for the treatment of cryptococcosis. Int. J. Antimicrob. Agents 2019, 54, 301–308. [Google Scholar] [CrossRef] [PubMed]
  237. Kim, J.H.; Western Regional Research Center, USDA-ARS, Albany, CA, USA; Haranahalli, K.; Stony Brook University, Stony Brook, NY, USA. Personal communication, 2020.
  238. Saccharomyces Genome Database. Available online: www.yeastgenome.org (accessed on 3 September 2020).
  239. Parsons, A.B.; Brost, R.L.; Ding, H.; Li, Z.; Zhang, C.; Sheikh, B.; Brown, G.W.; Kane, P.M.; Hughes, T.R.; Boone, C. Integration of chemical-genetic and genetic interaction data links bioactive compounds to cellular target pathways. Nat. Biotechnol. 2004, 22, 62–69. [Google Scholar] [CrossRef] [PubMed]
  240. Kim, J.H.; Campbell, B.C.; Yu, J.; Mahoney, N.; Chan, K.L.; Molyneux, R.J.; Bhatnagar, D.; Cleveland, T.E. Examination of fungal stress response genes using Saccharomyces cerevisiae as a model system: Targeting genes affecting aflatoxin biosynthesis by Aspergillus flavus Link. Appl. Microbiol. Biotechnol. 2005, 67, 807–815. [Google Scholar] [CrossRef]
  241. U.S. Food and Drug Administration. Substances Added to Food. Available online: https://www.fda.gov/food/food-additives-petitions/substances-added-food-formerly-eafus (accessed on 21 July 2020).
  242. Fungicide Resistance Action Committee. Available online: https://www.frac.info (accessed on 21 July 2020).
  243. Kim, J.H.; Chan, K.L.; Cheng, L.W. Cinnamic acid analogs as intervention catalysts for overcoming antifungal tolerance. Molecules 2017, 22, 1783. [Google Scholar]
  244. Ma, C.-M.; Abe, T.; Komiyama, T.; Wang, W.; Hattori, M.; Daneshtalab, M. Synthesis, anti-fungal and 1,3-β-d-glucan synthase inhibitory activities of caffeic and quinic acid derivatives. Bioorg. Med. Chem. 2010, 18, 7009–7014. [Google Scholar] [CrossRef]
  245. Couto, N.; Wood, J.; Barber, J. The role of glutathione reductase and related enzymes on cellular redox homoeostasis network. Free Radic. Biol. Med. 2016, 95, 27–42. [Google Scholar] [CrossRef]
  246. Xue, T.; Nguyen, C.K.; Romans, A.; May, G.S. A mitogen-activated protein kinase that senses nitrogen regulates conidial germination and growth in Aspergillus fumigatus. Eukaryot. Cell 2004, 3, 557–560. [Google Scholar] [CrossRef] [Green Version]
  247. Reyes, G.; Romans, A.; Nguyen, C.K.; May, G.S. Novel mitogen-activated protein kinase MpkC of Aspergillus fumigatus is required for utilization of polyalcohol sugars. Eukaryot. Cell 2006, 5, 1934–1940. [Google Scholar] [CrossRef] [Green Version]
  248. Berman, J.; Krysan, D.J. Drug resistance and tolerance in fungi. Nat. Rev. Microbiol. 2020, 18, 319–331. [Google Scholar] [CrossRef] [PubMed]
  249. Beardsley, J.; Halliday, C.L.; Chen, S.C.-A.; Sorrell, T.C. Responding to the emergence of antifungal drug resistance: Perspectives from the bench and the bedside. Future Microbiol. 2018, 13, 1175–1191. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  250. Hernando-Amado, S.; Coque, T.M.; Baquero, F.; Martínez, J.L. Defining and combating antibiotic resistance from One Health and Global Health perspectives. Nat. Microbiol. 2019, 4, 1432–1442. [Google Scholar] [CrossRef] [PubMed]
  251. 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]
  252. Translational Informatics Division at University of New Mexico. DrugCentral. Available online: http://drugcentral.org (accessed on 3 September 2020).
  253. Clinical Trials Transformation Initiative. Aggregate Content of ClinicalTrials.gov (AACT). Available online: https://aact.ctti-clinicaltrials.org/ (accessed on 3 September 2020).
Figure 1. Yeast dilution bioassay showing differential susceptibility of S. cerevisiae slt2Δ, bck1Δ, and glr1Δ mutants to cinnamic acid analogs (0.5 mM) (adapted from [243]). Numbers 100, 10−1, 10−2, 10−3, 10−4, and 10−5 indicate the cell dilution rate for yeast spotting; growth scores 101, 102, 103, 104, 105, and 106 denote cell numbers which appeared following incubation. slt2Δ, mitogen-activated protein kinase (MAPK) mutant; bck1Δ, MAPK kinase kinase (MAPKKK) mutant; glr1Δ, glutathione reductase mutant.
Figure 1. Yeast dilution bioassay showing differential susceptibility of S. cerevisiae slt2Δ, bck1Δ, and glr1Δ mutants to cinnamic acid analogs (0.5 mM) (adapted from [243]). Numbers 100, 10−1, 10−2, 10−3, 10−4, and 10−5 indicate the cell dilution rate for yeast spotting; growth scores 101, 102, 103, 104, 105, and 106 denote cell numbers which appeared following incubation. slt2Δ, mitogen-activated protein kinase (MAPK) mutant; bck1Δ, MAPK kinase kinase (MAPKKK) mutant; glr1Δ, glutathione reductase mutant.
Antibiotics 09 00812 g001
Figure 2. (a) A. fumigatus MAPK mutant showing tolerance to the repurposed benzoic ingredient (Kim et al., unpublished observation); (b) scheme describing structural modifications of cinnamates or benzoates that could overcome the tolerance of S. cerevisiae glr1Δ or A. fumigatus MAPK mutants, respectively, to the repurposed compounds (see Figure 1 and [243] for cinnamates).
Figure 2. (a) A. fumigatus MAPK mutant showing tolerance to the repurposed benzoic ingredient (Kim et al., unpublished observation); (b) scheme describing structural modifications of cinnamates or benzoates that could overcome the tolerance of S. cerevisiae glr1Δ or A. fumigatus MAPK mutants, respectively, to the repurposed compounds (see Figure 1 and [243] for cinnamates).
Antibiotics 09 00812 g002
Table 1. Tools and procedures applicable to the in silico/computational antifungal drug repurposing.
Table 1. Tools and procedures applicable to the in silico/computational antifungal drug repurposing.
Pathogens
Steps
Paracoccidioides SpeciesCandida albicansCandida, Aspergillus, and Trychophyton SpeciesPythium insidiosumParacoccidioides SpeciesCandida auris
Mining and compilation of pathogen genome dataBroad Institute Fungal Genomics Database (https://www.broadinstitute.org/fungal-genome-initiative)The Basic Local Alignment Search Tool (BLASTP), National Center for Biotechnology Information (http://www.ncbi.nlm.nih.gov/BLAST/);
CLUSTALW, European
Bioinformatics Institute (http://www.ebi.ac.uk/Tools/msa/)
MOE 2014.09 (PDB search module);
Clustal Omega tool, UniProt server (http://www.ebi.ac.uk/Tools/msa/clustalo/)
UniProt database (https://www.uniprot.org);
Protein Data Bank (https://www.rcsb.org);
BLASTP (blast 2.2.28_ program) (https://www.ncbi.nlm.nih.gov)
--
Identification of orthologs in pathogen isolatesOrthoVenn server -----
Identification of homologs in the drug–target databasesDrugBank; Therapeutic targets database (TTD)---MDL Drug Data Report (MDDR) (BIOVIA and Thomson Reuters);
DrugBank (https://www.drugbank.ca/);
TargetMol provider (http://targetmol.com/)
-
Selection of essential targets in the Saccharomyces cerevisiae systemDatabase of Essential Genes (DEG) -----
Homology modelingSWISS-MODEL server;
Protein Data Bank
(PDB);
KoBaMIN server;
HCC server; MolProbity
SWISS-MODEL server;
Ramachandran plot, SAVES
(http://services.mbi.ucla.edu/PROCHECK)
AMBER99 (with S. cerevisiae chorismate mutase; PDB entry code 4CSM) as a templatePubChem server (http://pubchem.ncbi.nlm.nih.gov/) (3D structure);
PreADMET server (https://preadmet.bmdrc.kr/introduction/) (3D structure)
Homology modeling for Candida albicans thioredoxin reductase [174]Cytochrome P450 51 (CYP51) modeling [175]
Ligand preparation and Molecular dockingOMEGA v.3.0.0.1 software by OpenEye Scientific [176];
QUACPAC v.1.7.0.2 software by OpenEye Scientific [177];
OEDocking suite v.3.2.0 by OpenEye Scientific [178];
FRED program with the ChemGauss4 score function in the OEDocking suite
Protein Preparation Wizard of Schrodinger’s suite 8.5 (Chimeric 1EA1);
Docking between internal ligand (fluconazole) and 1EA1 in Schrodinger’s suite 8.5;
GROMACS 5.0 for molecular dynamics simulations with
GROMOS9643a1 force field;
LIGPREP and
MAESTRO (fluvastatin 3D
structure);
MATLAB version R2015b (plotting and calculations);
Dundee Prodrug 2.5;
GLIDE 5.0
MOE 2014.09 (Chemical
Computing Group Inc., Sherbooke St, Montreal, QC, Canada) software;
Leadit 2.1.2 (BioSolveIT GmbH, Germany) software;
Gromacs 4.5.5 (Molecular dynamics);
PRODRG2 and GROMOS 53A6 force field (Topology);
Particle Mesh Ewald (PME) method and linear constraint solver (LINCS);
Xmgrace and VMD software
AutoDock Vina program; PyRx suite open-source software version 0.9.7;
AutoDock Vina program;
Discovery Studio Visualizer version 17.2.0 (Dassault Systemes Biovia Corp.)
Scaffold Hunter program (http://scaffoldhunter.sourceforge.net/) (selection of top ten best compounds from each database);
GOLD software (docking simulation against Candida albicans thioredoxin reductase);
CORINA (three-dimensional models);
Visual Molecular Dynamics (VMD) program (http://www.ks.uiuc.edu/Research/vmd/);
LigPlot program (https://www.ebi.ac.uk/thornton-srv/software/LIGPLOT/)
Protein preparation wizard (C. albicans CYP51);
LigPrep module of the Schrodinger suite (sertraline);
Schrodinger suite (LLC, New York, NY) (induced fit molecular docking analysis);
Glide module (extra precision feature for sertraline)
Experimental validation in the target pathogenModified Clinical and Laboratory Standards Institute (CLSI) protocolMicrotiter bioassay3-(4,5-dimethylthiazol-2-yl)-2,5-diphenyltetrazolium bromide (MTT) microdilution assayAgar- and broth-based assayCLSI M27-A3 [171]
Microdilution (killing kinetics)
References [15,176,177,178,179,180,181,182,183,184,185,186,187,188,189,190,191,192][18,193,194,195][16,196,197,198,199,200,201,202,203,204][17,205,206][20,171,174,207,208,209,210][21,175]
Table 2. Examples of repurposed drugs negatively affecting biofilm formation, hyphal filamentation, or yeast growth in fungi.
Table 2. Examples of repurposed drugs negatively affecting biofilm formation, hyphal filamentation, or yeast growth in fungi.
CompoundsFungiEffectsRef.
Haloperidol or
benzocyclane
derivative
C. albicans,
C. glabrata,
C. neoformans
Inhibition of filamentation, melanin production, and
biofilm formation
[78]
AripiprazoleC. albicansInhibition of biofilm formation and hyphal filamentation[152]
Alexidine dihydrochlorideC. albicans,
C. auris,
A. fumigatus
Antibiofilm activity[83]
MefloquineC. albicans,
C. neoformans,
Inhibition of the expression of virulence traits: filamentation in C. albicans and capsule formation/melanization in C. neoformans[155]
Pentamidine, bifonazole,
econazole, alexidine,
cetylpyridinium chloride,
otilonium bromide,
benzethonium chloride,
niclosamide, disulfiram,
temsirolimus
C. neoformansInhibition of spore germination and yeast growth[100]
Sulfonamide drugsC. albicansInhibition of biofilm[113]
MiltefosineC. albicans,
C. auris,
C. dubliniensis,
C. glabrata,
C. krusei,
C. parapsilosis, C. tropicalis,
Sporothrix schenckii
Inhibition of both planktonic growth and biofilm formation; inhibition of Coccidioides posadasii filamentous phase and Histoplasma capsulatum filamentous/yeast phases[116,117,119]
MebendazoleC. neoformansAntifungal activity against phagocytized C. neoformans: affected biofilms and reduced capsular dimensions[157]
QuinacrineC. albicansInhibition of biofilm and
inhibition of planktonic growth (alkaline pH) and filamentation
[123]
Auranofin,
pyrvinium pamoate, benzbromarone
C. albicansInhibition of biofilm formation[124]
FinasterideC. albicansInhibition of urinary biofilm formation and filamentation[164]
AuranofinC. albicans,
Staphylococcus aureus
Inhibition of C. albicans and S. aureus (bacterium) mono- and dual biofilm formation[161]
PanobinostatC. albicansInhibition of biofilm, hyphal, and planktonic growth
[143]
RobenidineA. fumigatus,
C. albicans,
C. neoformans,
S. cerevisiae
Inhibition of yeast cell growth, filamentation, and biofilm formation[167]
bis-Biguanide alexidine dihydrochlorideC. albicansAntifungal and antibiofilm activity[83]
Halogenated salicylanilide,
niclosamide
C. albicans,
C. auris
Antifilamentation and antibiofilm activities [169]
Arachidonic acid C. albicans,
C. parapsilosis,
C. glabrata,
C. tropicalis
Antibiofilm activity[88]
Aspirin, ibuprofenC. albicans,
Trichosporon asahii
Antibiofilm and antiplanktonic activity[60,62]
NortriptylineC. utilis,
C. krusei,
C. glabrata
Antihyphal and antibiofilm activity[74]
QuinineC. albicansAntifungal synergy with bicarbonate or hygromycin against biofilm[125]
Table 3. Examples of animal models used in drug repurposing (see also Table S1, Supplementary Materials).
Table 3. Examples of animal models used in drug repurposing (see also Table S1, Supplementary Materials).
Drug/Compound RepurposedAnimal ModelFungiEffectReference
RaltegravirBALB/c mice,
male, 6 weeks old
Paracoccidioides speciesReduction of the fungal burden, decreased alterations in the lung structure of mice (1 mg/kg of raltegravir)[20]
ThioridazineMurine J774 phagocytesC. neoformansDecreased the intracellular burden of C. neoformans (2.7-fold at a concentration 16-fold below the MIC (2 μg/mL))[75]
BeauvericinSpecific pathogen-free female ICR (Crl: CD-1) miceC. albicans or C. parapsilosisReduction in tissue damage and inflammatory cell infiltration in
kidneys (0.5 mg/kg beauvericin and KET 1)
[85]
BeauvericinBALB/c (inbred) mice, female, 7 weeks oldC. albicansBeauvericin (4 mg/kg) and FLU 1 (0.5 mg/kg) combination exhibited a therapeutic benefit[84]
PentamidineC57BL/6J, female, 8 to 10 weeks oldC. neoformansMice treated prophylactically with pentamidine (for 3 days prior to infection) resulted in a 2-fold-lower fungal burden than the control;
minimized lung fungal burden in spore-mediated infections of mice
[100]
DeferasiroxC57BL/6 mice, female, 4 to 6 weeks old;
immunosuppression model of murine oropharyngeal candidiasis
C. albicansPreventive deferasirox treatment significantly reduced the fungal burden in tongue tissue [137]
N-AcetylcysteineC57/BL6 mice, female, 6 to 8 weeks oldC. gattiiN-Acetylcysteine + AMB 1 achieved higher survival than the control and reduced morbidity in murine-induced cryptococcosis; reduced fungal burden in lungs/brain and lower concentrations of proinflammatory cytokines in the lungs[138]
CisplatinBALB/c mice, female, 4 to 6 weeks old
C. neoformansCisplatin significantly inhibited C. neoformans growth in a mouse model[168]
PanobinostatGalleria mellonella larvaeC. albicansPanobinostat and FLU combination enhanced
survival rate of G. mellonella
[143]
Pilocarpine hydrochlorideG. mellonella larvaeC. albicansPilocarpine hydrochloride
protects G. mellonella larvae from C. albicans-induced mortality in a dose-dependent manner
[154]
PitavastatinCaenorhabditis elegans animal modelC. albicansPitavastatin–FLU combination reduced the biofilm formation of
Candida species and the fungal burdens in a C. elegans infection model
[77]
1 Drug abbreviations: amphotericin B (AMB), fluconazole (FLU), and ketoconazole (KET).
Table 4. Summary of the drug/compound libraries used in the antifungal drug repurposing (see also Table S1, Supplementary Materials).
Table 4. Summary of the drug/compound libraries used in the antifungal drug repurposing (see also Table S1, Supplementary Materials).
Drug, Compound Libraries SourcesFungi TestedReferences
MDL Drug Data Report (MDDR),
DrugBank, TargetMol databases or library (L4200)
BIOVIA and Thomson Reuters https://www.drugbank.ca/, http://targetmol.com/C. albicans,
Paracoccidioides species
[20,167]
Prestwick Chemical
Library
Prestwick Chemical (Illkirch, France) http://www.prestwickchemical.com/Candida species,
Cryptococcus species,
Saccharomyces cerevisiae,
Aspergillus fumigatus,
Fusarium oxysporum,
Fusarium solani,
Lichtheimia species,
Lomentospora prolificans,
Paecilomyces variotii,
Rhizopus arrhizus,
Scedosporium apiospermum
[23,25,66,75,83,94,96,124,165,228]
Pharmakon1600 drug libraryMicroSource Discovery Systems (Gaylordsville, CT, USA)
http://www.msdiscovery.com/
C. albicans,
C. glabrata
C. auris
[68,77,162]
Library of Pharmacologically Active Compounds (LOPAC1280)Sigma-Aldrich
(St. Louis, MO, USA)
https://www.sigmaaldrich.com/
Exserohilum rostratum[101]
Medicines for Malaria Venture (MMV) Malaria Box Medicines for Malaria Venture
(Geneva, Switzerland)
https://www.mmv.org/
[229]
C. albicans,
C. gatti,
C. neoformans,
L. prolificans,
Fonsecaea pedrosoi,
Fonsecaea monophora,
Fonsecaea nubica,
Cladophialophora carrionii,
Phialophora verrucosa,
Rhinocladiela similis,
Exophiala jeanselmei var. heteromorpha,
Exophiala dermatitidis
[91,148]
Screen-Well Enzo library
of 640 compounds
Enzo Life Sciences (Farmingdale, NY, USA)
https://www.enzolifesciences.com/
Aspergillus species,
Candida,
Cryptococcus deuterogattii,
Saccharomyces
[110]
L1300 Selleck library,
1018 United States Food and Drug Administration (FDA)-approved Selleck library
Selleck Chemicals
https://www.selleckchem.com/
A. fumigatus,
C. gattii,
C. glabrata,
C. neoformans,
Trichophyton rubrum
[73,100]
Pathogen Box® chemical libraryMedicines for Malaria Venture
(Geneva, Switzerland)
https://www.mmv.org/
C. auris[119]
United States National Institutes of Health/National Cancer Institute (NIH/NCI) compound
library
Developmental Therapeutics Program of the NIH/NCI (Rockville, MD, USA)
https://dtp.cancer.gov/
A. fumigatus,
Candida species,
C. neoformans
[127]
1547 or 1581 FDA-approved
drug library
Johns Hopkins, USA
Johns Hopkins Clinical Compound Library (JHCCL) version 1.0
C. albicans,
C. auris,
C. krusei,
C. parapsilosis,
C. tropicalis
[133,163]
678 Maybridge collectionThermo Fisher Scientific (Waltham, MA, USA)
https://www.thermofisher.com/us/en/home/chemicals/maybridge.html
C. albicans,
C. auris
[169]
Table 5. Summary of the drugs and compounds exerting synergism during co-application (see also Table S1, Supplementary Materials).
Table 5. Summary of the drugs and compounds exerting synergism during co-application (see also Table S1, Supplementary Materials).
Repurposed Drugs/Compounds Co-Applied.Conventional Antifungal Drugs with Synergism
N-Acetylcysteine, alexidine dihydrochloride, amiodarone, arachidonic acid, aspirin, beauvericin, bis-biguanide alexidine dihydrochloride, benzocyclane, bromperidol derivative, chenodiol, chlorcyclizine, clomiphene, cloperastine, colistin, l-cycloserine, deferasirox, drospirenon, ebselen, erythromycin, glimepiride, ibuprofen, idoxifene, lovastatin, methylene-idoxifene, miltefosine
nisoldipine, nortriptyline, panobinostat, perhexiline, pitavastatin, polymyxin b, promazine, pyrvinium pamoate, quinacrine, quinine,
ribavirin, riluzole, sertraline, suloctidil, tamoxifen, thioridazine, thiosemicarbazone, toremifene, trifluoperazine
Azoles
(clotrimazole, fluconazole, isavuconazole, itraconazole, ketoconazole, miconazole, posaconazole, voriconazole),
echinocandins
(caspofungin, anidulafungin, micafungin),
polyene (AMB),
allylamine (terbinafine)
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Kim, J.H.; Cheng, L.W.; Chan, K.L.; Tam, C.C.; Mahoney, N.; Friedman, M.; Shilman, M.M.; Land, K.M. Antifungal Drug Repurposing. Antibiotics 2020, 9, 812. https://doi.org/10.3390/antibiotics9110812

AMA Style

Kim JH, Cheng LW, Chan KL, Tam CC, Mahoney N, Friedman M, Shilman MM, Land KM. Antifungal Drug Repurposing. Antibiotics. 2020; 9(11):812. https://doi.org/10.3390/antibiotics9110812

Chicago/Turabian Style

Kim, Jong H., Luisa W. Cheng, Kathleen L. Chan, Christina C. Tam, Noreen Mahoney, Mendel Friedman, Mikhail Martchenko Shilman, and Kirkwood M. Land. 2020. "Antifungal Drug Repurposing" Antibiotics 9, no. 11: 812. https://doi.org/10.3390/antibiotics9110812

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