Next Article in Journal
Neuropeptide S-Mediated Modulation of Prepulse Inhibition Depends on Age, Gender, Stimulus-Timing, and Attention
Next Article in Special Issue
Synthesis and Study of Antifungal Properties of New Cationic Beta-Glucan Derivatives
Previous Article in Journal
The Influence of Pharmacogenetics on the Clinical Relevance of Pharmacokinetic Drug–Drug Interactions: Drug–Gene, Drug–Gene–Gene and Drug–Drug–Gene Interactions
Previous Article in Special Issue
The Repurposing of Acetylsalicylic Acid as a Photosensitiser to Inactivate the Growth of Cryptococcal Cells
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Drug Repurposing in Medical Mycology: Identification of Compounds as Potential Antifungals to Overcome the Emergence of Multidrug-Resistant Fungi

1
Aix Marseille Univ, IRD, APHM, MEPHI, 13005 Marseille, France
2
IHU Méditerranée Infection, 13005 Marseille, France
*
Author to whom correspondence should be addressed.
The authors contributed equally.
Pharmaceuticals 2021, 14(5), 488; https://0-doi-org.brum.beds.ac.uk/10.3390/ph14050488
Submission received: 12 March 2021 / Revised: 13 May 2021 / Accepted: 18 May 2021 / Published: 20 May 2021

Abstract

:
Immunodepression, whether due to HIV infection or organ transplantation, has increased human vulnerability to fungal infections. These conditions have created an optimal environment for the emergence of opportunistic infections, which is concomitant to the increase in antifungal resistance. The use of conventional antifungal drugs as azoles and polyenes can lead to clinical failure, particularly in immunocompromised individuals. Difficulties related to treating fungal infections combined with the time required to develop new drugs, require urgent consideration of other therapeutic alternatives. Drug repurposing is one of the most promising and rapid solutions that the scientific and medical community can turn to, with low costs and safety advantages. To treat life-threatening resistant fungal infections, drug repurposing has led to the consideration of well-known and potential molecules as a last-line therapy. The aim of this review is to provide a summary of current antifungal compounds and their main resistance mechanisms, following by an overview of the antifungal activity of non-traditional antimicrobial drugs. We provide their eventual mechanisms of action and the synergistic combinations that improve the activity of current antifungal treatments. Finally, we discuss drug repurposing for the main emerging multidrug resistant (MDR) fungus, including the Candida auris, Aspergillus or Cryptococcus species.

1. Introduction

Most public health organizations, including the World Health Organization, do not have a fungal infection surveillance program, despite the fact that invasive fungal infections present a high mortality rate worldwide, often exceeding 50% [1,2]. Fungal infections have long been poorly documented and recognized, perhaps due to the need to treat other severe and serious bacterial and viral infections. However, mycoses should no longer be ignored.
The signs and symptoms of fungal infection appear during antibiotic therapy, particularly due to opportunistic fungal agents. In particular, invasive fungal infections affect patients with compromised immune systems, such as those with hematologic malignancies, HIV infection, chemotherapy treatments, etc. [3,4]. In addition, other factors such as the ageing of the population, which is susceptible to these opportunistic infections and improvements in diagnostic methods have led to their increasing prevalence in hospitals. A significant number of fungal agents, including yeast and yeast-like species such as Candida spp., Cryptococcus spp. and molds such as the Aspergillus species, complicate clinical management, with a variety of symptoms, prevalence and clinical outcomes [5]. Moreover, fungal infections can also occur in healthy people, so it is difficult to control their spread [6]. Therefore, greater consideration should be given to monitoring fungal infections [7,8,9].
To deal with these fungal infections, there are only four main therapeutic classes currently used in clinical practice, namely polyenes, azoles, echinocandins and flucytosine. Although these drugs remain active, they display several limitations that complicate their routine use including off-target toxicity, drugs interaction, clinical failure and long-term treatment [10]. Furthermore, the emerging resistance to antifungals and the poor clinical response of many isolates to antifungal therapy make this an even greater public-health concern. For example, previous exposure to an antifungal agent such as fluconazole has been shown to increase the risk of fluconazole-resistant Candida infections in immunocompromised patients [11]. Antifungal resistance remains a critical global problem, although it may vary depending on the species, geography and available therapeutic alternatives [5]. Some species are known to be more resistant than others, leading to treatment failure; these include Candida pathogens (Candida glabrata, Candida krusei, Candida lusitaniae and the very newly emerging yeast: Candida auris), some cryptococcal species and opportunistic Aspergillus or Fusarium species associated with immunocompromised hosts [5]. The aforementioned species commonly exhibit high intrinsic antifungal resistance profiles, sometimes to different classes of antifungals as is the case with almost all Fusarium spp. to triazoles, 5-fluorocytosine and echinocandins [5].
Despite great efforts made internationally to deal with antibiotic resistance by reducing the inappropriate consumption of antimicrobials, this does not yet extend to the use of antifungal agents. In reality, antifungal resistance may arise from fungicide use in agriculture, as described in a recent study in 2017, where plant bulbs were found to be positive for triazole-resistant Aspergillus fumigatus, rather than from clinical use [12]. Furthermore, antifungal therapies are mostly given to immunocompromised patients or those in intensive care units (ICUs), when treatment is unavoidable, rather than in preventive use in community medicine. Demers et al. described the impact of the heterogeneity of a single gene (MRR1), found in different C. lusitaniae subpopulations in an azole-naïve cystic fibrosis patient, on the level of fluconazole resistance, highlighting other indirect factors involved, such as the host immune system and coinfecting bacteria [13]. Therefore, reducing the consumption of antifungals may not be the only solution to improving the difficult management of invasive fungal diseases.
The development of new antifungal drugs represents a major challenge for the pharmaceutical industry, since fungi are eukaryotic organisms and have a close evolutionary relationship with their human hosts [14]. In addition, the pharmaceutical industry is no longer interested in developing and marketing new antimicrobials, including antifungals [1,15]. Indeed, the drug development process remains very expensive, time-consuming and risky due to many factors [16]. Given the difficulties of treating invasive fungal infections, consideration should be made to use efficient alternative strategies to implement immediate and appropriate measures. New targets involving enzymes and other metabolic pathways, or new formulations/generations are under development to broaden the spectrum of antifungal activities and to potentially overcome current resistances [17,18,19,20]. New antifungals can also be found in natural products or plant extracts, inspired by traditional medicine and aromatherapy [21,22]. Indeed, various essential oils showed in vitro efficiency against clinical yeast or fungi [23,24,25]. Preclinical studies to determine human toxicities, pharmacodynamics and clinical trials must continue [21,25]. At the same time, the strategy of drug repurposing (also called drug repositioning) consists of identifying drugs known to be effective for another indication than that for which they are marketed [15]. This strategy has gained in popularity in recent years and has already proven to be effective, particularly in oncology, cardiology and Alzheimer’s disease. These reinvestigated drugs have already completed preclinical trials, main human toxicities are well known, and the research and development process can be, considerably, reduced allowing for lower investment costs and faster potential clinical use [15,26]. However, the setting up of this strategy is not a straightforward matter as it does not exclude carrying out further clinical trials before the simple repositioning of a given drug, but it must also initially deal with the intellectual property rights, regulatory/authority process, license grant, pricing, patient’s acceptance, marketing strategy and commercialization in order to avoid any failure at the development stage [27].
The aim of this review is to report on the non-antifungal drugs that may be active against the most common emerging multidrug-resistant (MDR) fungi in human pathology. We first summarize the antifungal compounds currently used for clinical therapies and their main mechanisms of resistance and then report on alternative drugs used to treat fungal infections that have been reported in the literature so far. Finally, we address different drugs that have a potential for repurposing to treat the main difficult to treat fungi.

2. Current Antifungal Agents and Their Mechanisms of Resistance

Since their first and progressive discovery in the mid-20th century, systemic antifungals have improved the management of many invasive fungal infections. The spectrum of antifungals and their mechanisms of action are diverse, since they act on different structures in the fungal cell; we describe them below (Figure 1).
The first antifungal agents, polyenes, were developed in 1950s for clinical use and have a large spectrum of activity against yeast and filamentous fungi. Polyenes are macrocyclic organic molecules from a soil actinomycete, Streptomyces nodosus [28]. The two most clinically relevant members of this class are topical nystatin and intravenous amphotericin B. Both bind to a sterol moiety, ergosterol, on the fungal cell membrane, disrupting cell permeability and leading to cellular lysis. Nystatin is not effective against dermatophytes but is effective against the Candida species [29]. Liposomal amphotericin B has been developed to allow the administration of higher doses with less nephrotoxicity (fewer side effects) to mammalian cells.
Resistance to polyenes is mainly related to changes in the lipid structure of the membrane and subsequently a modification in its fluidity and absorbency. The principal altered effects concern enzymes that are involved in the synthesis of ergosterol. Deficiencies in the ERG2 and ERG3 genes, which code for the isomerase of C-8 sterol and delta-5,6-desaturase, induce modifications in membrane sterols. The quantity of ergosterol is modified, consequently affecting polyene activity. Boosted activity of catalase, an antioxidant enzyme that decreases oxidative injury, represents another polyene-resistance mechanism. However, intrinsic amphotericin B resistance frequently described in some Aspergillus spp. strains not only include the alteration of the ergosterol pathway but the signaling pathways, such as those described in A. terreus [30] or increased enzymatic activity of the peroxidase and superoxide dismutase in A. flavus [31] (Table 1).
Discovered in 1980s, azoles now represent the best conventional antifungal agents for medical treatment. Azole compounds have had a major impact on the treatment of invasive fungal infections over the last 35 years. The use of the first available azoles, imidazoles including ketoconazole, miconazole and clotrimazole, was primarily restricted to the treatment of superficial fungal infections. These compounds were substituted by the first-generation of triazoles, such as fluconazole and itraconazole, to broaden the range of application. Later, the search for new antifungals was intensified to overcome some efficiency limitations and to prevent emerging resistant pathogens. The second generation of triazoles (voriconazole, posaconazole, efinaconazole and isavuconazole) were developed with an extended spectrum of activity [32,33]. The final target of these drugs is ergosterol from the cell membrane. They inhibit lanosterol-14α-demethylase in mitochondria, which interferes with the synthesis of the membrane ergosterol. Each azole has a different affinity in its inhibition of lanosterol, which may explains the differences in spectrum of activity among azole agents [34], but all have a strong inhibitor effect on the CYP450 enzyme system, which is responsible for many drug interactions [35]. The global HIV epidemic led to the widespread and significant use of fluconazole to treat oro-oesophageal candidiasis in HIV-infected patients and fluconazole-resistant Candida strains were later widely reported in these patients [34,36].
Four mechanisms of resistance have been demonstrated so far: (i) activation of the efflux pumps due to an overexpression of membrane-associated transporters encoded by the gene families of transporters (CDR and MDR) reduces azole plasmatic concentrations [30,34]; (ii) azole agents cannot, qualitatively, bind to their enzymatic target to target changes induced by a mutation in ERG11 gene, which encodes for the lanosterol-14α-demethylase [33,34]; (iii) some strains induce the overexpression of ERG11 as a compensatory mechanism and increase the intracellular concentration of this protein. Thus, an increasing concentration of a given azole agent is needed to remain efficient. This resistance mechanism involves quantitative changes by upregulating the target enzyme [30,34]; (iv) some strains, with a mutation of the ERG3 gene, developed a pathway bypassing fungal membrane biosynthesis by replacing ergosterol with 14α-methyl-fecosterol and preventing the accumulation of a toxic product: 14ɑ-methyl-3,6-diol. This alternative route of ergosterol biosynthesis maintains both the function of the fungal membrane and the resistance to azoles [33] (Table 1).
Mannan, chitin and α- and β-glucans are the main compounds of the fungal cell wall. Drugs belonging to the echinocandins target one of these components and act as antifungals, inhibiting β-(1,3)-glucan synthetase leading to a depletion of β-(1,3)-glucan, an essential component for the structuring and function of the cell wall [33]. Echinocandins (caspofungin, anidulafungin and micafungin) are semisynthetic cyclic lipopeptides derived from natural products. They present a reserve supplement to the arsenal of drugs available to treat invasive fungal diseases with a fungicidal action against the Candida species and a fungistatic activity against the Aspergillus species [37].
Candida resistance to echinocandins has been related to several mutations in a hot-spot region of the FKS1 gene, which encodes for a subunit of echinocandin target, resulting in a lower affinity between the antifungal and its target [5,30] (Table 1).
A pyrimidine analogue, 5-fluorocytosine (5-FC), was developed in 1957 as an antimetabolite. Without any potential use as an anticancer treatment, it was used to treat fungal infections. Once 5-fluorocytosine enters the fungal cell, enzymes convert it in compounds to be incorporated in the synthesized RNA. This disrupts the protein synthesis of the affected fungi. It is also converted as a potent inhibitor of thymidylate synthase, which interferes in DNA synthesis and nuclear division [33]. It is always combined with other drugs, in association with azoles or amphotericin B, due to the high prevalence of intrinsic resistance in many fungal species and to the rapid development of resistance in yeast [33,34]. The primary resistance was about 10% in the Candida albicans strains [30]. Resistance to 5-FC may be due to mutations in FUR1 or in the genes FCY1 and FCY2 leading to defects in flucytosine metabolism [30] (Table 1).
Finally, it is notable that, besides the limited number of available antimycotic agents and the burden of antifungal resistance (described above), toxicities and drug–drug interactions of antifungals, the treatment failure due to clinical resistance is frequently reported in invasive mycoses [34]. In fact, in vitro antifungal susceptibility testing does not guarantee the success of in vivo treatment. Many factors affecting the infected patient, the responsible fungal strain and the prescribed antifungal agent, which have been well reviewed elsewhere [34], may explain this clinical resistance.

3. Non-Antifungal Drugs Identified as Having a Potential Antifungal Activity against Invasive Fungal Strains

Given the rapid evolution of resistance to antifungal drugs and the high prevalence of mycoses in clinical settings due to the increasing number of human immunodeficiency cases and/or as the result of improved fungal diagnosis, there is an urgent need to improve the efficacy of current treatments and to develop new therapeutic strategies. It is worth considering innovative approaches [5] or, simply, associations of existing drugs, which could be a promising approach to extending the use of current antifungal agents. Indeed, drug combination resulting in a synergistic activity has the potential to impede the evolution of drug resistance through employing several mechanisms or targets [15]. However, this should not be considered as the only “miracle” solution because interactions can be indifferent and combinations can be unsuccessful [38].
Drug repositioning or repurposing allows for new indications for previously approved drugs that are already marketed for other medical reasons. This approach offers many benefits over de novo drug development. Previously established pharmacokinetic and pharmacodynamic profiles and toxicity data allow for faster and cheaper development of repositioned molecules. Consequently, clinical use may be considered to overcome the rapid emergence of resistant fungi and outbreaks [15,39]. In this paper, we reported on molecules that have been found to be active in vitro or even in vivo against fungal agents, according to their initial therapeutic class (Figure 1 and Table 2).

3.1. Antimicrobials Apart from Antifungals

Polymyxins including colistin and polymyxin B are peptide antibiotics and target the outer membrane of Gram-negative bacteria [40]. Tested against MDR yeasts and molds, polymyxins showed an antifungal activity with minimum inhibitory concentrations (MICs) ranging from 16 to 128 µg/mL [41]. In particular, a fungicidal effect was described with colistin against C. albicans, Cryptococcus neoformans and Rhodotorula mucilaginosa. The mechanism of action was similar to the bacterial mechanism, by inducing membrane damage to MDR-C. albicans, as observed under fluorescent microscopy [41]. Using a checkerboard microdilution assay, synergistic activity was revealed with colistin-amphotericin B and colistin-itraconazole, against MDR C. albicans and Lichtheimia corymbifera strains [41]. The colistin-azoles combination has also been reported more recently in strains showing low susceptibilities to fluconazole. Bibi M. et al. confirm that colistin binds to the lipids of fungal membranes and works in relation to an ergosterol depletion level due to the previous action of azoles [42].
Table 2. Drugs with reported in vitro antifungal activities. NR: not reported.
Table 2. Drugs with reported in vitro antifungal activities. NR: not reported.
DrugFirst IndicationAntifungal ActivityActivity RangeAntifungal Mechanism of ActionReferences
AntimicrobialsPolymyxins
Colistin
Polymyxin B
Gram-negative bacterial infectionsC. albicans16–128 µg/mLMembrane damages on Candida albicans[41]
C. neoformans
R. mucilaginosa
S. apiospermum
L. prolificans
F. oxysporum
F. solani
R. oryzae
RibavirinHepatitis CC. albicans0.37–3.02 µg/mLDisruption of vacuoles function of C. albicans strains[41,43]
C. tropicalis
C. parapsilosis
OxyclozanideAnimal
parasitosis
C. albicans16–32 µg/mLUncoupling the mitochondrial electron transport from phosphorylation and changing the mitochondrial membrane potential[44]
ChloroquineMalariaC. neoformans3.19 µg/mLIron deprivation[45]
(10 µM)
C. albicans31.2–250 µg/mLInhibition of ergosterol biosynthesis[46]
S. cerevisiaeNRGrowth inhibition via blocking thiamine transportation[47]
MebendazoleHelminthiasis C. neoformans92.5 ng/mLMorphological alterations by reducing capsular dimension[48]
C. gatti(0.3125 µM)
Anti-inflammatoryAuranofinRheumatoid
arthritis
C. albicans0.25–16 µg/mLAction on reactive-oxygen-mediated cell death[49,50]
A. fumigatus
S. apiospermum
L. prolificans
C. neoformans
AspirinInflammation Cryptococcus spp.1–10 mg/mLStress induction via ROS-mediated damage[51,52]
IbuprofenCandida spp.
TheophyllineAsthma, COPDCandida spp.1.4–1.8 mg/mLMembrane damages by ionic and ergosterol modifications[53]
AntipsychoticsHaloperidolPsychosis C. albicans<4 µg/mL Possible action on GPCRs, mediators of signals across the cell membrane[54,55]
TrifluperidolC. neoformans
SertralineDepressionC. neoformans2–6 µg/mL Inhibition of protein synthesis [56,57]
Lomentospora prolificans
Scedosporium spp., Fusarium spp.8–32 μg/mL
Paecilomyces spp., Alternaria spp. and Curvularia spp.
ChlorpromazineSchizophrenia Candida spp.1–16 µg/mL Possible modifications of membrane[58,59]
C. neoformans
Filamentous fungi: Aspergillus spp., Scedosporium spp., Pseudallescheria spp. and
Zygomycetes
AnticancersTamoxifenBreast cancer Candida spp.
C. neoformans
8–64 µg/mL Prevention of proteins calmodulin from binding to calcineurin, cell lysis and alteration of fungal development[60,61,62]
ToremifeneDisturb the cell wall integrity via interaction with Ccr1
OthersDisulfiramAlcoholism Candida spp.1–16 µg/mLChelating metals
Inhibition of multidrug transporter implicated in drug resistance
[63,64]
C. neoformans
Aspergillus spp.
Other non-antifungal agents with bacterial activity have demonstrated a synergy with antifungals, as described by Rossato et al. [65]. For example, erythromycin with amphotericin B showed no toxic in vivo effect and could be a promising novel combination in invasive antifungal therapy [66].
Ribavirin, a purine nucleoside analogue, displays a broad-spectrum activity against many RNA and DNA viruses. Ribavirin is used to treat hepatitis C virus in combination with interferon-α [67]. Tournu et al. identified ribavirin as a potential C. albicans disrupting agent of vacuole, which is essential to yeast pathogenicity [43]. Based on this, we demonstrated the fungistatic activity of ribavirin against MDR C. albicans and fungicidal activity against C. parapsilosis. MICs largely ranged from 1.56 to 12.5 μmol/L (0.37–3.02 µg/mL) against C. albicans, C. parapsilosis and C. tropicalis. Synergistic activity was also reported when the antiviral agent was combined with either amphotericin B, fluconazole or itraconazole, against MDR C. albicans and was thus proposed to be further investigated for clinical use [68].
In terms of antiparasitic (anthelmintic) drugs, the activity of oxyclozanide, a halogenated salicylanilide, was demonstrated against C. albicans isolates, including ones that were resistant to azole and echinocandin. This anthelmintic agent seems to alter/disturb the mitochondrial oxidative phosphorylation function and thus its ability to use the nonfermentable carbon sources by disrupting the mitochondrial membrane potential [44]. Oxyclozanide is widely used as an antiparasitic veterinary drug against the liver fluke Fasciola hepatica [69], and has been studied for antibacterial properties against colistin-resistant Gram-negative bacilli infections [70]. Pic et al. showed that oxyclozanide inhibited, at 58% and 99%, the growth of C. albicans at a concentration of 10 and 100 μM, respectively. These concentrations are comparable to the therapeutic dose used for an ovine weighing 45 kg [71], however its repurposing strategy in human therapies requires further and careful data on pharmacokinetics/pharmacodynamics and/or in vivo therapeutic assays, as this drug has not yet been used in humans.
The first and main antimalarial drug, chloroquine, is able to be used in different indications as it has many effects on inflammatory responses, metabolic process, the immune system and infections [72,73]. Weber et al. described that chloroquine treatment of macrophages infected with cryptococcal cells led to the formation of iron complexes inducing the death of C. neoformans [45]. Chloroquine has also been shown to inhibit thiamine transporters in the yeast Saccharomyces cerevisiae, linked to glucose metabolism [47]. It can also damage fungal morphogenesis due to an anormal synthesis of ergosterol in drug-resistant C. albicans strains [74].
Benzimidazoles such as mebendazole, albendazole, flubendazole and triclabendazole are broad-spectrum anthelmintic drugs. Joffe et al. demonstrated the efficacy of benzimidazoles in inhibiting the growth of C. neoformans, especially mebendazole and flubendazole [48]. Mebendazole has been suggested to be repurposed as an anticryptococcal drug because it can efficiently penetrate the blood–brain barrier in animal models [48,75].

3.2. Anti-Inflammatory Drugs

Recently, Ogundeji et al. reported that aspirin and ibuprofen can control the growth of cryptococcal cells, with a high susceptibility of C. neoformans strains. Ibuprofen had a greater inhibitory effect than aspirin on all 10 Cryptococcus spp. tested strains at various drug concentrations [51]. The effects of ibuprofen seem to be dose-dependent; at high concentration (10 mg/mL), Candida cells are killed whereas at lower concentration (5 mg/mL), the drug was fungistatic [52]. In addition, synergistic outcomes were observed between ibuprofen and fluconazole or amphotericin B in Cryptococcus spp. and Candida spp., with fractional inhibitory concentration (FIC) < 0.5 [51,52,76].
Auranofin inhibits several inflammatory pathways and has been used since 1985 as an antirheumatic drug. It has already been found to be effective to treat bacterial infections [77] but Wiederhold et al. showed that auranofin also displayed an activity against various yeast and molds such as A. fumigatus, Scedosporium apiospermum and Lomentospora prolificans [49]. Although auranofin MICs were sometimes higher than those of the reference treatment (i.e., fluconazole), these concentrations could be easily achieved in patients’ blood treated with the usual therapeutic dose [49]. Therefore, auranofin, along with its activity on biofilms [78], could be used as a promising antifungal treatment [54,66]. However, due to its immunosuppressive action, this drug should be carefully investigated prior to administration as an antifungal agent, given that fungi-infected patients are commonly immunocompromised.
Finally, theophylline is generally used for asthma or chronic obstructive pulmonary disease (COPD) and has been proposed to treat candidiasis as it had effects on cell membrane integrity [53].

3.3. Antipsychotic Drugs

Oral antipsychotic drugs are often used in routine clinical practice, so their side effects and toxicity are now well known. Recently, haloperidol and trifluperidol were described for their antifungal activity against C. albicans (with MICs values < 4 µg/mL) or against C. neoformans [55,79]. The authors demonstrated that the two antipsychotics had a similar effect to fluconazole, acting on the yeast membrane [54]. The combination of a haloperidol-derivative with the antifungal posaconazole displayed a 16-fold reduction in MIC values for both the azole agent and the antipsychotic, compared with each drug alone (from >32 and >128 to 2 and 8 µg/mL respectively) with a fractional inhibitory concentration index (FICI) of 0.13 [79]. Strong synergies were also observed against C. glabrata and Aspergillus terreus [79].
Sertraline, the most frequently prescribed antidepressant, has been reported to be fungicidal against C. neoformans, with MICs ranging between 2 and 6 µg/mL [56,80,81] but also against emerging fungi [57,82]. Moreover, in vivo sertraline antifungal activity was tested, in murine models of cryptococcosis, showing a reduction in the fungal burden [56]. The combination of this compound and azoles or amphotericin B showed efficiency against various Cryptococcus spp. strains [83,84]. Of the 53 tested isolates, 31 were affected by the synergistic combination sertraline-fluconazole (FICI ≤ 0.5) [83]. However, sertraline has shown some antagonist effects with fluconazole against Candida strains [56].
Despite many in vitro, in vivo and human studies showing the efficiency of sertraline as antifungal, disparities appear in the clinicals studies [80,85]. In 2019, a clinical trial of 486 participants testing sertraline as an adjunctive treatment to IV amphotericin B and oral fluconazole to treat HIV-associated cryptococcal meningitis did not significatively improve survival, but did reveal a similar fungal clearance rate between groups [85]. In 2020, another randomized trial testing sertraline as pre-emptive therapy was stopped without a final conclusion due to the severe side effects of sertraline that were observed [86].
Chlorpromazine and trifluoperazine, dopamine antagonists, are used to treat schizophrenia. They are largely reported as having antibacterial and antifungal effects [58,59,87] and Vitale et al. confirmed the in vitro antifungal activity against difficult-to-treat filamentous fungi such as Aspergillus species (A. fumigatus, A. ustus and A. terreus), zygomycetes (Absidia corymbifera, Rhizopus oryzae and R. microspores) and Scedosporium species (S. apiospermum and S. prolificans) [58]. Phenothiazines including chlorpromazine appear to act on the fungal membrane but this needs further consideration [58]. The use of the combination must be advised to avoid resistance, to have greater efficacy and a less toxic effect: for example, chlorpromazine-amphotericin B against C. neoformans [84] or Candida species [59] demonstrated an interesting synergism profile.

3.4. Anticancer Drugs

Anticancer drugs may represent an important source of potential repurposing drugs. Indeed, as they often work on the basic metabolism pathways of eucaryotic and human cells such as the DNA replicating pathway, yeast cells may also be affected. Various anticancer drugs have been seen to be efficient in vitro against yeast growth [54,88]. Butts et al. demonstrated the fungicidal activity of tamoxifen and toremifene, two estrogenic receptor antagonists, against C. neoformans within macrophages, where the main pathogenesis of this organism happens [60]. Tamoxifen is described as an inhibitor of calmodulin. In this way, tamoxifen-treated yeasts showed cell lysis and an alteration of fungal development [61,89]. Recently, the interaction between tamoxifen and its target Ccr1 has also been described as causing the disruption of cell wall integrity [62].
However, administration of anticancer molecules to patients who are often immunocompromised should only be taken after careful consideration due to the numerous side effects that these compounds can cause, including immunosuppression [90]. A bypass solution via a synergistic association could circumvent this limitation. Interestingly, both compounds were synergistic in vitro with amphotericin B and fluconazole against cryptococcal cells [60]. In addition, in vivo candidiasis was cured, in a murine model, by the administration of 200 mg/kg of body weight per day of tamoxifen [61]. Since then, a randomized phase II clinical trial of tamoxifen as an adjuvant to the gold standard therapy for cryptococcal meningitis is in progress [91].
Antifolates as inhibitors of purine synthesis have also been reported to be effective agents against yeast development by reducing the quantity of ergosterol [65].

3.5. Other Approved Drugs

Finally, we reported on a few other well-known and old drugs that could be repurposed as their indications, doses and side effects are known and thus can be easily managed.
Fluvastatin, rosuvastatin, atorvastatin and simvastatin are statins used to lower the synthesis of cholesterol by inhibiting HMG-CoA. Statins may be an important adjuvant for the treatment of fungal infections because their efficiency has been reported against fungi, even against azole-resistant yeasts [65,92]. Indeed, Macreadie et al. demonstrated a strong inhibition of the growth of Candida spp. (with the exception of C. krusei on YEPD media containing 100 µM of statins) and A. fumigatus [93]. On the fungal cell, statins might work on the pathway of mevalonate synthesis causing a decrease in ergosterol quantity of the cell membrane [93]. Statins showed in vitro synergy with various azoles and may display beneficial outcomes on candidiasis according to one cohort study in 2013 [65,81,94].
Disulfiram is an alcohol antagonist drug that has been used in clinical practice for many years. Khan et al. reported its antifungal potential in 2007 and provided an MIC range from 1 to 16 μg/mL for fluconazole-sensitive and resistant yeast strains, with an inhibiting effect on biofilm formation [95,96]. It also had a fungicidal activity on Aspergillus spp. [95]. Earlier, Shukla et al. demonstrated that disulfiram could reverse Cdr1p-mediated drug resistance so it could be used in combination to sensitize resistant strains [63,64]. Broadly, side effects of disulfiram are uncommon, contraindications include pregnancy and unstable cardiovascular disease and close hepatic monitoring is required [97]. Given all this information, in vitro, in vivo and clinical trials should be pursued to fully justify the repositioning of this molecule.

4. Some Emerging Multidrug-Resistant Fungi and Their Compounds with Repurpose Potential Identified through Phenotypic Screening

Resistance to at least one class of antifungal agents is a concern of most existing fungal species. However, some major pathogens have a relatively high resistance rate and constitute a serious public health burden, especially C. albicans, Cryptococcus spp. and Aspergillus spp. In addition, other emerging and potentially life-threatening pathogens are increasingly being reported [98]. Some are still not well characterized and are opportunistic and MDR, such as C. auris [99], Scedosporium spp. and Fusarium spp. [100]. We present a non-exhaustive list of compounds below that could be effective as an alternative therapeutic strategy, depending on these MDR species (Figure 2).

4.1. C. albicans Biofilms

The pathogenicity of the Candida species resides in their ability to form biofilms, thereby protecting them from external elements such as antifungal agents or the host’s immune system. The common implantation of a Candida biofilm on a medical device such as a catheter can lead to candidemia and severe systemic infections. In the absence or failure of treatment, the health and economic consequences are significant [101]. Some drugs alone or in combination with a currently used antifungal are effective in vitro to reduce or inhibit the formation of a biofilm [102,103].
As reported above, chloroquine reverts the azole resistance in biofilms [46], and aspirin [104], disulfiram [96] or auranofin [78] decrease in vitro the biofilm formation of many Candida species. Quinacrine is an antimalarial, which was used during WWII but remains available to treat giardiasis or cutaneous leishmaniasis. It has been effective at preventing and treating C. albicans biofilms with MIC ranges of 64–256 µg/mL [93]. In addition, both amphotericin B and caspofungin have synergy with quinacrine with an FICI equal to 0.37 and 0.31, respectively [105]. Benzodiazepines such as midazolam or diazepam have also been proposed to be repurposed against the biofilm formation of yeast, due to interaction with its virulence factors [106,107]. Anti-inflammatory compounds appear to be very efficient, such as celecoxib, etodolac, meloxicam, etc. [81]. In addition, flufenamic acid (a nonsteroidal anti-inflammatory drug, NSAID) showed an excellent action in the prevention and treatment of biofilms from echinocandin-resistant strains [108]. The hypnotic agent, etomidate, works against the biofilm of fluconazole-resistant Candida spp. strains [109]. The antiseptic alexidine [103] and the finasteride used to treat the prostatic hyperplasia [110] were highly active in vitro at preventing C. albicans biofilms. Nile et al. highlighted the cholinergic receptor, which slowed down the biofilm-mediated virulence of C. albicans while having a boosting effect on the host immune response using pilocarpine [111] or tropicamide in a study by Machado et al. [112]. The antipsychotic drug aripiprazole was as effective as azoles by acting on the early formation of the pseudo hyphal [113]. As mentioned, many drugs from different therapeutic classes can be effective as an antibiofilm agent. In vivo studies first, followed by clinical trials remain important to confirm those in vitro options.

4.2. C. auris

C. auris is considered to be a serious global health threat due to its association with nosocomial invasive infections, its high mortality rate and its multidrug resistant profile [114]. High rates of antifungal resistance have been reported for fluconazole and amphotericin B, which is considered to be an intrinsic resistance, but some acquired resistances to echinocandins have been described in some countries [115,116]. The need for effective agents must be a priority in order to address future outbreaks, which, given the current situation, are highly likely [117]. To quickly identify candidate drugs, screening molecule libraries can offer various solutions [49,66,118]. Cheng et al. found six novel anti-C. auris compounds among more than 4300 approved drugs with 13 possible different drug associations [118]. The amebicide iodoquinol and leishmanicide miltefosine were reported in those screenings as being potential repositionable compounds due to their inhibition of C. auris growth [49]. Sulodictil [66], ebselen [66], antiemetic aprepitant [119] and lopinavir [120] were active against C. auris, either alone or in association with currently used antifungals. Finally, most of the drugs described above as being effective on yeast are also effective on C. auris, such as sertraline [121], oxyclonazide [44], colistin [112], etc. Indeed, these molecules are generally active via novel targets, which are still naïve in antifungal treatments and therefore in yeast adaptation.

4.3. Aspergillus Species

Invasive aspergillosis threatens the lives of millions of immunocompromised patients every year, with a mortality rate of 50/60% [98]. These species come from the environment and some patients are infected via contaminated foods [122,123]. Used for agricultural production, certain fungicides display a similar mechanism as azoles used in clinical routine. The evolution of resistance by selective pressure is attributed to their widespread use and there is an urgent need to find alternative strategies to current antifungals [124,125,126]. As a result, different drug screenings have been performed and have found that clozafimine, tacrolimus, cyclosporin [127], haloperidol [79], disulfiram [95], chlorpromazine [58] and auranofin [122] as reported above, could be repurposed as an anti-Aspergillus spp. treatment. Interesting synergies were also obtained between celecoxib and primaquine and paromomycin and β-escin with a FICI of <0.27 or <0.38, respectively, against A. fumigatus growth [128].

4.4. Cryptococcus Species

Occurring predominantly during the course of an immunocompromised patient’s disease, cryptococcosis causes pneumonia or meningoencephalitis, due to an inhalation of Cryptococcus cells from the environment [129]. In addition to their worldwide spread, treatment of these infections remains challenging for clinicians who deal with only three classes of antifungal agents, as echinocandins are not effective against them [89]. Several molecules were found to be active against Cryptococcus cell growth, as mentioned above, such as auranofin, aspirin, ibuprofen [122], tamoxifen [129], mebendazole [48], sertraline [86,121] and disulfiram [95], and thioridazine, amiodarone [122], miltefosine [130] and calcium channel blockers such as nifedipine [131] have also been described. The antiparasitic drug, flubendazole, was very active, with MICs ranging from 0.039 to 0.156 μg/mL, even against fluconazole-resistant strains [131]. Despite this, the perfect anticryptococcal treatment would exhibit a low toxicity for polymedicated, immunosuppressed patients would be well-distributed around the body to eradicate all cryptococcal niches including the cerebrospinal fluid [129].

4.5. Other Non-Aspergillus Molds

More rarely, other MDR fungal pathogens can cause aggressive and disseminated infections associated with poor prognosis such as Hyalohyphomycetes, including the genera Fusarium, Scedosporium, etc., or the Mucormycetes with Mucor and Rhizomucor group. Early and effective treatment is required to prevent the progression of the infection and to limit outbreaks [132,133]. However, this infection remains difficult to treat because the aforementioned fungi are resistant to most current antifungal drugs (Table 1) [132,134]. We previously screened a library of 1280 drugs against six of these filamentous fungi including Fusarium, Scedosporium, Rhizopus and Lichtheimia species [135]. The main hits found were antifungals, antiseptics and some antineoplastics against a few strains, and polymyxins [41], disulfiram [95], auranofin [122] and ribavirin [68], as mentioned above. These fungi have fewer hits and further investigations are warranted in order not to reach fatal therapeutic impasses. Some combinations of antibiotics and antifungals, however, displayed strong synergies [65].

5. Further Assessment and Prioritization of Repurpose Potential

It is important to highlight in this effort that preliminary in vitro studies, however, encouraging, do not automatically imply repurpose potential. In addition, not all compounds identified above possess equal repurpose potential while others may be altogether inappropriate for the clinical use in question. Given that safety data for marketed drugs and advanced clinical candidates are available, further assessment should be undertaken to evaluate and prioritize each molecule with respect to its repurpose potential. For example, effective auranofin concentrations that inhibited yeasts and molds’ growth, ranged from 0.25 to 16 µg/mL [50], but human pharmacokinetics displayed a Cmax of 0.025 μg/mL after 6 mg per day [136]. Aspirin on the other hand was effective against yeasts [44,69] in accordance with human doses and used for a long time now. For analgesic dosages, the Cmax was reported on average at 50 µM (9 mg/L) and were higher for anti-inflammatory use [137]. Another example, quinacrine alone, was effective against C. albicans biofilms with MICs between 64 and 256 µg/mL [63]. However, quinacrine pharmacokinetics from an intrapleural dose displayed a Cmax below 1 µg/mL for a 600 mg dose [138] and the therapeutic dose is initially 100 mg, which makes the use of this molecule in monotherapy inappropriate. In spite of that, its use in synergic combinations led to a reduce in the initial dose, from 64 µg/mL alone to 4 µg/mL with caspofungine [105] and approaching acceptable human concentrations. At the current stage, the use of quinacrine as a monotherapy does not reach the reported human concentrations, but the study of its combinations with other molecules is still interesting in view of the reported synergisms. Plasma concentrations of synergistic combinations of azoles or amphotericin B and colistin can be above some previously reported MICs, where human doses of colistin have resulted in serum concentrations up to 32 µg/mL [41]. In such cases, the renal condition of the patient must be considered, although nebulized doses of colistin could avoid this toxicity [139]. Therefore, an analysis process must be applied to all molecules before their repurposing after in vitro, in vivo and other assays. In summary, strict and careful analysis is essential before administration, concerning the consistency between the effective dose and the serum concentration/toxicity dose, but also covering all other parameters related or not to the molecule, such as potential drug interactions, bioavailability of the molecule, strain susceptibility, patient condition and consent, approval of the patient’s care panel, etc.

6. Conclusions

The emergence of new mycotic agents and the increase in antifungal resistance has led to the need to find new and/or alternative drugs. We are currently seeing that non-traditional antimicrobial agents, previously prescribed to treat non-infectious conditions, may display antimicrobial properties, and it would be a worthwhile investment to further explore these compounds before being repurposed. Otherwise, these reported drugs could serve as a starting point for the reinnovation of a new molecule. We noted compounds ranging from anti-inflammatory to antipsychotic drugs, which have been documented to control fungal growth and to be repositionable. However, their clinical application may be limited to treat life-threatening fungi, due to drug toxicity, especially with anticancer drugs that generally target eukaryotic organisms or with drugs inducing an immunosuppressive state. Susceptibility testing on each fungus, careful analysis of pharmacokinetic/pharmacodynamic (PK/PD) data and human achievable and tolerable concentrations must be confirmed before administration to assess the benefit/risk. All effective treatments should be taken into consideration as a last-line therapy, even if they have side effects, if they could save the patient [140]. Fortunately, in in vitro assays some of these compounds act synergistically with currently used antifungal agents. Thus, combinations enable the use of low concentrations with the advantage of minimizing any possible undesired physiological effects. This also enhances the efficiency of traditional antimicrobial drugs, which are fungistatic when used alone under normal conditions.
Assuming that obstacles may be overcome, drug repurposing is a promising alternative strategy into which further research and clinical trials are essential to combat the increase in invasive fungal infections.

Author Contributions

L.P., H.Y., J.-M.R. and F.B. drafted and revised the manuscript. All authors read and approved the final manuscript.

Funding

This work was supported by the French Government under the “Investissements d’avenir” (Investments for the Future) program managed by the Agence Nationale de la Recherche (ANR, fr: National Agency for Research), (reference: Méditerranée Infection 10-IAHU-03). This work was supported by Région Provence-Alpes-Côte d’Azur and European funding (FEDER (Fonds européen de développement régional) PRIMMI (Plateformes de Recherche et d’Innovation Mutualisées Méditerranée Infection)).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

No new data were created or analyzed in this study. Data sharing is not applicable to this article.

Acknowledgments

We thank our designer Juan Manuel Osorio Rozo for Figure 1 and Figure 2 (https://www.behance.net/juanmanuelosoriorozo, accessed on 18 May 2021).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Brown, G.D.; Denning, D.W.; Gow, N.A.R.; Levitz, S.M.; Netea, M.G.; White, T.C. Hidden Killers: Human Fungal Infections. Sci. Transl. Med. 2012, 4, 165rv13. [Google Scholar] [CrossRef] [Green Version]
  2. Denning, D.W. Calling upon all public health mycologists: To accompany the country burden papers from 14 countries. Eur. J. Clin. Microbiol. Infect. Dis. 2017, 36, 923–924. [Google Scholar] [CrossRef]
  3. Livio, P.; Morena, C.; Anna, C.; Massimo, O.; Luana, F.; Bruno, M.; Domenico, P.; Marco, P.; Alessandro, B.; Anna, C.; et al. The epidemiology of fungal infections in patients with hematologic malignancies: The SEIFEM-2004 study. Haematologica 2006, 91, 1068–1075. [Google Scholar]
  4. Armstrong-James, D.; Meintjes, G.; Brown, G.D. A neglected epidemic: Fungal infections in HIV/AIDS. Trends Microbiol. 2014, 22, 120–127. [Google Scholar] [CrossRef] [PubMed]
  5. Geddes-McAlister, J.; Shapiro, R.S. New pathogens, new tricks: Emerging, drug-resistant fungal pathogens and future prospects for antifungal therapeutics. Ann. N. Y. Acad. Sci. 2019, 1435, 57–78. [Google Scholar] [CrossRef] [PubMed]
  6. Gupta, R.; Goel, N.; Gupta, A.; Gupta, K.B.; Chaudhary, U.; Sood, S. A rare fungal infiltration of lungs in a healthy young girl. Case Rep. Pulmonol. 2011, 2011, 1–3. [Google Scholar] [CrossRef] [Green Version]
  7. Srinivasan, A.; Lopez-Ribot, J.L.; Ramasubramanian, A.K. Overcoming antifungal resistance. Drug Discov. Today Technol. 2014, 11, 65–71. [Google Scholar] [CrossRef] [Green Version]
  8. Alcazar-Fuoli, L.; Mellado, E. Current status of antifungal resistance and its impact on clinical practice. Br. J. Haematol. 2014, 166, 471–484. [Google Scholar] [CrossRef]
  9. Meurman, J.H.; Hämäläinen, P.; Janket, S.J. Oral infections in older adults. Aging Health 2006, 2, 1013–1023. [Google Scholar] [CrossRef]
  10. Perfect, J.R. The antifungal pipeline: A reality check. Nat. Rev. Drug Discov. 2017, 16, 603–616. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  11. Sanguinetti, M.; Posteraro, B.; Lass-Flörl, C. Antifungal drug resistance among Candida species: Mechanisms and clinical impact. Mycoses 2015, 58, 2–13. [Google Scholar] [CrossRef]
  12. Barber, A.E.; Riedel, J.; Sae-Ong, T.; Kang, K.; Brabetz, W.; Panagiotou, G.; Deising, H.B.; Kurzai, O. Effects of agricultural fungicide use on Aspergillus fumigatus: Abundance, antifungal susceptibility, and population structure. mBio 2020, 11. [Google Scholar] [CrossRef] [PubMed]
  13. Demers, E.G.; Biermann, A.R.; Masonjones, S.; Crocker, A.W.; Ashare, A.; Stajich, J.E.; Hogan, D.A. Evolution of drug resistance in an antifungal-naive chronic Candida lusitaniae infection. Proc. Natl. Acad. Sci. USA 2018, 115, 12040–12045. [Google Scholar] [CrossRef] [Green Version]
  14. Hall, R.A.; Noverr, M.C. Fungal interactions with the human host: Exploring the spectrum of symbiosis. Curr. Opin. Microbiol. 2017, 40, 58–64. [Google Scholar] [CrossRef]
  15. Peyclit, L.; Baron, S.A.; Rolain, J.-M. Drug repurposing to fight colistin and carbapenem-resistant bacteria. Front. Cell. Infect. Microbiol. 2019, 9, 193. [Google Scholar] [CrossRef] [PubMed]
  16. Dimasi, J.A.; Feldman, L.; Seckler, A.; Wilson, A. Trends in risks associated with new drug development: Success rates for investigational drugs. Clin. Pharmacol. Ther. 2010, 87, 272–277. [Google Scholar] [CrossRef]
  17. Houšt’, J.; Spížek, J.; Havlícek, V. Antifungal Drugs. Metabolites 2020, 10. [Google Scholar] [CrossRef] [Green Version]
  18. Wiederhold, N.P. The antifungal arsenal: Alternative drugs and future targets. Int. J. Antimicrob. Agents 2018, 51, 333–339. [Google Scholar] [CrossRef]
  19. Rauseo, A.M.; Coler-Reilly, A.; Larson, L.; Spec, A. Hope on the horizon: Novel fungal treatments in development. Open Forum Infect. Dis. 2020, 7, 1–19. [Google Scholar] [CrossRef] [Green Version]
  20. Su, H.; Han, L.; Huang, X. Potential targets for the development of new antifungal drugs. J. Antibiot. 2018, 71, 978–991. [Google Scholar] [CrossRef]
  21. Mahboubi, M. Artemisia sieberi Besser essential oil and treatment of fungal infections. Biomed. Pharmacother. 2017, 89, 1422–1430. [Google Scholar] [CrossRef]
  22. D′agostino, M.; Tesse, N.; Frippiat, J.P.; Machouart, M.; Debourgogne, A. Essential oils and their natural active compounds presenting antifungal properties. Molecules 2019, 24, 3713. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  23. Córdoba, S.; Vivot, W.; Szusz, W.; Albo, G. Antifungal activity of essential oils against Candida species isolated from clinical samples. Mycopathologia 2019, 184, 615–623. [Google Scholar] [CrossRef]
  24. Barac, A.; Donadu, M.; Usai, D.; Spiric, V.T.; Mazzarello, V.; Zanetti, S.; Aleksic, E.; Stevanovic, G.; Nikolic, N.; Rubino, S. Antifungal activity of Myrtus communis against Malassezia sp. isolated from the skin of patients with pityriasis versicolor. Infection 2018, 46, 253–257. [Google Scholar] [CrossRef]
  25. González-Burgos, E.; Gómez-Serranillos, M.P. Natural Products for Vulvovaginal Candidiasis Treatment: Evidence from Clinical Trials. Curr. Top. Med. Chem. 2018, 18, 1324–1332. [Google Scholar] [CrossRef] [PubMed]
  26. Pushpakom, S.; Iorio, F.; Eyers, P.A.; Escott, K.J.; Hopper, S.; Wells, A.; Doig, A.; Guilliams, T.; Latimer, J.; McNamee, C.; et al. Drug repurposing: Progress, challenges and recommendations. Nat. Rev. Drug Discov. 2018, 18, 41–58. [Google Scholar] [CrossRef]
  27. Gill, R.; Amberkar Mohan Babu, V.; Meena Kumari, K. Assets and liabilities of drug repositioning. Int. J. Pharma Bio Sci. 2016, 7, 47–53. [Google Scholar] [CrossRef]
  28. 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]
  29. Kuriyama, T.; Karasawa, T.; Williams, D.W. Antimicrobial Chemotherapy: Significance to Healthcare. In Biofilms in Infection Prevention and Control: A Healthcare Handbook; Elsevier Inc.: Amsterdam, The Netherlands, 2014; pp. 209–244. ISBN 9780123970435. [Google Scholar]
  30. Scorzoni, L.; de Paula e Silva, A.C.A.; Marcos, C.M.; Assato, P.A.; de Melo, W.C.M.A.; de Oliveira, H.C.; Costa-Orlandi, C.B.; Mendes-Giannini, M.J.S.; Fusco-Almeida, A.M. Antifungal therapy: New advances in the understanding and treatment of mycosis. Front. Microbiol. 2017, 8. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  31. Rudramurthy, S.M.; Paul, R.A.; Chakrabarti, A.; Mouton, J.W.; Meis, J.F. Invasive aspergillosis by Aspergillus flavus: Epidemiology, diagnosis, antifungal resistance, and management. J. Fungi 2019, 5, 55. [Google Scholar] [CrossRef] [Green Version]
  32. Allen, D.; Wilson, D.; Drew, R.; Perfect, J. Azole antifungals: 35 years of invasive fungal infection management. Expert Rev. Anti Infect. Ther. 2015, 13, 787–798. [Google Scholar] [CrossRef]
  33. Ghannoum, M.A.; Rice, L.B. Antifungal agents: Mode of action, mechanisms of resistance, and correlation of these mechanisms with bacterial resistance. Clin. Microbiol. Rev. 1999, 12, 501–517. [Google Scholar] [CrossRef] [Green Version]
  34. Kanafani, Z.A.; Perfect, J.R. Resistance to antifungal agents: Mechanisms and clinical impact. Clin. Infect. Dis. 2008, 46, 120–128. [Google Scholar] [CrossRef] [Green Version]
  35. Yu, D.T.; Peterson, J.F.; Seger, D.L.; Gerth, W.C.; Bates, D.W. Frequency of potential azole drug-drug interactions and consequences of potential fluconazole drug interactions. Pharmacoepidemiol. Drug Saf. 2005, 14, 755–767. [Google Scholar] [CrossRef] [PubMed]
  36. Canuto, M.M.; Rodero, F.G. Antifungal drug resistance to azoles and polyenes. Lancet Infect. Dis. 2002, 2, 550–563. [Google Scholar] [CrossRef]
  37. Aguilar-Zapata, D.; Petraitiene, R.; Petraitis, V. Echinocandins: The expanding antifungal armamentarium. Clin. Infect. Dis. 2015, 61, S604–S611. [Google Scholar] [CrossRef] [Green Version]
  38. Chassot, F.; Venturini, T.P.; Piasentin, F.B.; Santurio, J.M.; Svidzinski, T.I.E.; Alves, S.H. Activity of antifungal agents alone and in combination against echinocandin-susceptible and -resistant Candida parapsilosis strains. Rev. Iberoam. Micol. 2019, 36, 44–47. [Google Scholar] [CrossRef] [PubMed]
  39. Ciliberto, G.; Mancini, R.; Paggi, M.G. Drug repurposing against COVID-19: Focus on anticancer agents. J. Exp. Clin. Cancer Res. 2020, 39. [Google Scholar] [CrossRef] [PubMed]
  40. Baron, S.; Hadjadj, L.; Rolain, J.-M.; Olaitan, A.O. Molecular mechanisms of polymyxin resistance: Knowns and unknowns. Int. J. Antimicrob. Agents 2016, 48, 583–591. [Google Scholar] [CrossRef]
  41. Yousfi, H.; Ranque, S.; Rolain, J.M.; Bittar, F. In vitro polymyxin activity against clinical multidrug-resistant fungi. Antimicrob. Resist. Infect. Control 2019, 8. [Google Scholar] [CrossRef] [PubMed]
  42. Bibi, M.; Murphy, S.; Benhamou, R.I.; Rosenberg, A.; Ulman, A.; Bicanic, T.; Fridman, M.; Berman, J. Combining Colistin and Fluconazole Synergistically Increases Fungal Membrane Permeability and Antifungal Cidality. ACS Infect. Dis. 2021, 7. [Google Scholar] [CrossRef]
  43. Tournu, H.; Carroll, J.; Latimer, B.; Dragoi, A.M.; Dykes, S.; Cardelli, J.; Peters, T.L.; Eberle, K.E.; Palmer, G.E. Identification of small molecules that disrupt vacuolar function in the pathogen Candida albicans. PLoS ONE 2017, 12. [Google Scholar] [CrossRef] [Green Version]
  44. Pic, E.; Burgain, A.; Sellam, A. Repurposing the anthelminthic salicylanilide oxyclozanide against susceptible and clinical resistant Candida albicans strains. Med. Mycol. 2019, 57, 387–390. [Google Scholar] [CrossRef]
  45. Weber, S.M.; Levitz, S.M.; Harrison, T.S. Chloroquine and the fungal phagosome. Curr. Opin. Microbiol. 2000, 3, 349–353. [Google Scholar] [CrossRef]
  46. Shinde, R.B.; Raut, J.S.; Chauhan, N.M.; Karuppayil, S.M. Chloroquine sensitizes biofilms of Candida albicans to antifungal azoles. Braz. J. Infect. Dis. 2013, 17, 395–400. [Google Scholar] [CrossRef] [Green Version]
  47. Huang, Z.; Srinivasan, S.; Zhang, J.; Chen, K.; Li, Y.; Li, W.; Quiocho, F.A.; Pan, X. Discovering thiamine transporters as targets of chloroquine using a novel functional genomics strategy. PLoS Genet. 2012, 8. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  48. 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]
  49. 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. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  50. 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]
  51. 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] [Green Version]
  52. Pina-Vaz, C.; Sansonetty, F.; Rodrigues, A.G.; Martinez-De-Oliveira, J.; Fonseca, A.F.; Mårdh, P.A. Antifungal activity of ibuprofen alone and in combination with fluconazole against Candida species. J. Med. Microbiol. 2000, 49, 831–840. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  53. 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, 129, 860–875. [Google Scholar] [CrossRef]
  54. 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] [PubMed] [Green Version]
  55. 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. 2019, 6. [Google Scholar] [CrossRef]
  56. Zhai, B.; Wu, C.; Wang, L.; Sachs, M.S.; Lin, X. The antidepressant sertraline provides a promising therapeutic option for neurotropic cryptococcal infections. Antimicrob. Agents Chemother. 2012, 56, 3758–3766. [Google Scholar] [CrossRef] [Green Version]
  57. 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]
  58. Vitale, R.G.; Afeltra, J.; Meis, J.F.G.M.; Verweij, P.E. Activity and post antifungal effect of chlorpromazine and trifluopherazine against Aspergillus, Scedosporium and zygomycetes. Mycoses 2007, 50, 270–276. [Google Scholar] [CrossRef] [PubMed]
  59. Galgóczy, L.; Bácsi, A.; Homa, M.; Virágh, M.; Papp, T.; Vágvölgyi, C. In vitro antifungal activity of phenothiazines and their combination with amphotericin B against different Candida species. Mycoses 2011, 54. [Google Scholar] [CrossRef] [PubMed]
  60. 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. [Google Scholar] [CrossRef] [Green Version]
  61. Dolan, K.; Montgomery, S.; Buchheit, B.; DiDone, L.; Wellington, M.; Krysan, D.J. Antifungal activity of tamoxifen: In vitro and in vivo activities and mechanistic characterization. Antimicrob. Agents Chemother. 2009, 53, 3337–3346. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  62. Liu, Q.; Guo, X.; Jiang, G.; Wu, G.; Miao, H.; Liu, K.; Chen, S.; Sakamoto, N.; Kuno, T.; Yao, F.; et al. NADPH-cytochrome P450 reductase ccr1 is a target of tamoxifen and participates in its antifungal activity via regulating cell wall integrity in fission yeast. Antimicrob. Agents Chemother. 2020, 64. [Google Scholar] [CrossRef]
  63. Shukla, S.; Sauna, Z.E.; Prasad, R.; Ambudkar, S.V. Disulfiram is a potent modulator of multidrug transporter Cdr1p of Candida albicans. Biochem. Biophys. Res. Commun. 2004, 322, 520–525. [Google Scholar] [CrossRef]
  64. Sauna, Z.E.; Shukla, S.; Ambudkar, S.V. Disulfiram, an old drug with new potential therapeutic uses for human cancers and fungal infections. Mol. Biosyst. 2005, 1, 127–134. [Google Scholar] [CrossRef]
  65. Rossato, L.; Camargo dos Santos, M.; Vitale, R.G.; de Hoog, S.; Ishida, K. Alternative treatment of fungal infections: Synergy with non-antifungal agents. Mycoses 2020. [Google Scholar] [CrossRef]
  66. Rossi, S.A.; De Oliveira, H.C.; Agreda-Mellon, D.; Lucio, J.; Soares Mendes-Giannini, M.J.; 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. [Google Scholar] [CrossRef]
  67. Te, H.S.; Randall, G.; Jensen, D.M. Mechanism of action of ribavirin in the treatment of chronic hepatitis C. Gastroenterol. Hepatol. 2007, 3, 218–225. [Google Scholar]
  68. 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]
  69. Swan, G.E. The pharmacology of halogenated salicylanilides and their anthelmintic use in animals. J. S. Afr. Vet. Assoc. 1999, 70, 61–70. [Google Scholar] [CrossRef] [Green Version]
  70. Ayerbe-Algaba, R.; Gil-Marqués, M.L.; Miró-Canturri, A.; Parra-Millán, R.; Pachón-Ibáñez, M.E.; Jiménez-Mejías, M.E.; Pachón, J.; Smani, Y. The anthelmintic oxyclozanide restores the activity of colistin against colistin-resistant Gram-negative bacilli. Int. J. Antimicrob. Agents 2019, 54, 507–512. [Google Scholar] [CrossRef]
  71. ANSES RCP—DOUVISTOME. Available online: http://www.ircp.anmv.anses.fr/rcp.aspx?NomMedicament=DOUVISTOME (accessed on 17 February 2021).
  72. Rolain, J.M.; Colson, P.; Raoult, D. Recycling of chloroquine and its hydroxyl analogue to face bacterial, fungal and viral infections in the 21st century. Int. J. Antimicrob. Agents 2007, 30, 297–308. [Google Scholar] [CrossRef]
  73. Plantone, D.; Koudriavtseva, T. Current and Future Use of Chloroquine and Hydroxychloroquine in Infectious, Immune, Neoplastic, and Neurological Diseases: A Mini-Review. Clin. Drug Investig. 2018, 38, 653–671. [Google Scholar] [CrossRef]
  74. Shinde, R.B.; Rajput, S.B.; Raut, J.S.; Mohan Karuppayil, S. An in vitro repositioning study reveals antifungal potential of chloroquine to inhibit growth and morphogenesis in Candida albicans. J. Gen. Appl. Microbiol. 2013, 59, 167–170. [Google Scholar] [CrossRef] [Green Version]
  75. Bai, R.Y.; Staedtke, V.; Wanjiku, T.; Rudek, M.A.; Joshi, A.; Gallia, G.L.; Riggins, G.J. Brain penetration and efficacy of different mebendazole polymorphs in a mouse brain tumor model. Clin. Cancer Res. 2015, 21, 3462–3470. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  76. Scott, E.M.; Tariq, V.N.; McCrory, R.M. Demonstration of synergy with fluconazole and either ibuprofen, sodium salicylate, or propylparaben against Candida albicans in vitro. Antimicrob. Agents Chemother. 1995, 39, 2610–2614. [Google Scholar] [CrossRef] [Green Version]
  77. Torres, N.S.; Abercrombie, J.J.; Srinivasan, A.; Lopez-Ribot, J.L.; Ramasubramanian, A.K.; Leung, K.P. Screening a commercial library of pharmacologically active small molecules against Staphylococcus aureus biofilms. Antimicrob. Agents Chemother. 2016, 60, 5663–5672. [Google Scholar] [CrossRef] [Green Version]
  78. 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 spp. J. Appl. Microbiol. 2020, 128, 88–101. [Google Scholar] [CrossRef] [Green Version]
  79. 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]
  80. Rhein, J.; Nielsen, K.; Boulware, D.R.; Meya, D.B. Sertraline for HIV-associated cryptococcal meningitis—Authors’ reply. Lancet Infect. Dis. 2016, 16, 1111–1112. [Google Scholar] [CrossRef] [Green Version]
  81. Moraes, D.C.; Ferreira-Pereira, A. Insights on the anticandidal activity of non-antifungal drugs. J. Mycol. Med. 2019, 29, 253–259. [Google Scholar] [CrossRef] [PubMed]
  82. da Rosa, T.F.; Machado, C.; de, S.; Serafin, M.B.; Bottega, A.; Foletto, V.S.; Coelho, S.S.; Hörner, R. Repositioning or Redirection of Antidepressant Drugs in the Treatment of Bacterial and Fungal Infections. Am. J. Ther. 2020, 27, e528–e532. [Google Scholar] [CrossRef]
  83. Nayak, R.; Xu, J. Effects of sertraline hydrochloride and fluconazole combinations on Cryptococcus neoformans and Cryptococcus gattii. Mycology 2010, 1, 99–105. [Google Scholar] [CrossRef]
  84. Rossato, L.; Loreto, É.S.; Zanette, R.A.; Chassot, F.; Santurio, J.M.; Alves, S.H. In vitro synergistic effects of chlorpromazine and sertraline in combination with amphotericin B against Cryptococcus neoformans var. grubii. Folia Microbiol. (Praha) 2016, 61, 399–403. [Google Scholar] [CrossRef]
  85. Rhein, J.; Huppler Hullsiek, K.; Tugume, L.; Nuwagira, E.; Mpoza, E.; Evans, E.E.; Kiggundu, R.; Pastick, K.A.; Ssebambulidde, K.; Akampurira, A.; et al. Adjunctive sertraline for HIV-associated cryptococcal meningitis: A randomised, placebo-controlled, double-blind phase 3 trial. Lancet Infect. Dis. 2019, 19, 843–851. [Google Scholar] [CrossRef]
  86. Boulware, D.R.; Nalintya, E.; Rajasingham, R.; Kirumira, P.; Naluyima, R.; Turya, F.; Namanda, S.; Rutakingirwa, M.K.; Skipper, C.P.; Nikweri, Y.; et al. Adjunctive sertraline for asymptomatic cryptococcal antigenemia: A randomized clinical trial. Med. Mycol. 2020, 58, 1037–1043. [Google Scholar] [CrossRef]
  87. 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]
  88. Montoya, M.C.; Krysan, D.J. Repurposing estrogen receptor antagonists for the treatment of infectious disease. mBio 2018, 9. [Google Scholar] [CrossRef] [Green Version]
  89. Butts, A.; Martin, J.A.; DiDone, L.; Bradley, E.K.; Mutz, M.; Krysan, D.J. Structure-Activity Relationships for the Antifungal Activity of Selective Estrogen Receptor Antagonists Related to Tamoxifen. PLoS ONE 2015, 10. [Google Scholar] [CrossRef]
  90. Quezada, H.; Martínez-Vázquez, M.; López-Jácome, E.; González-Pedrajo, B.; Andrade, Á.; Fernández-Presas, A.M.; Tovar-García, A.; García-Contreras, R. Repurposed anti-cancer drugs: The future for anti-infective therapy? Expert Rev. Anti Infect. Ther. 2020, 18, 609–612. [Google Scholar] [CrossRef] [Green Version]
  91. Ngan, N.T.T.; Mai, N.T.H.; Tung, N.L.N.; Lan, N.P.H.; Tai, L.T.H.; Phu, N.H.; Chau, N.V.V.; Binh, T.Q.; Hung, L.Q.; Beardsley, J.; et al. A randomized open label trial of tamoxifen combined with amphotericin B and fluconazole for cryptococcal meningitis. Wellcome Open Res. 2019, 4. [Google Scholar] [CrossRef]
  92. de Oliveira Neto, A.S.; Souza, I.L.A.; Amorim, M.E.S.; de Freitas Souza, T.; Rocha, V.N.; do Couto, R.O.; Fabri, R.L.; de Freitas Araújo, M.G. Antifungal efficacy of atorvastatin-containing emulgel in the treatment of oral and vulvovaginal candidiasis. Med. Mycol. 2020. [Google Scholar] [CrossRef]
  93. 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]
  94. de Queiroz Ribeiro, N.; Costa, M.C.; Magalhães, T.F.F.; Carneiro, H.C.S.; Oliveira, L.V.; Fontes, A.C.L.; Santos, J.R.A.; Ferreira, G.F.; de Sousa Araujo, G.R.; Alves, V.; et al. Atorvastatin as a promising anticryptococcal agent. Int. J. Antimicrob. Agents 2017, 49, 695–702. [Google Scholar] [CrossRef] [PubMed]
  95. Khan, S.; Singhal, S.; Mathur, T.; Upadhyay, D.J.; Rattan, A. Antifungal potential of disulfiram. Jpn. J. Med. Mycol. 2007, 48, 109–113. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  96. Hao, W.; Qiao, D.; Han, Y.; Du, N.; Li, X.; Fan, Y.; Ge, X.; Zhang, H. Identification of disulfiram as a potential antifungal drug by screening small molecular libraries. J. Infect. Chemother. 2020. [Google Scholar] [CrossRef]
  97. Blanc, M.; Daeppen, J.-B. Does disulfiram still have a role in alcoholism treatment? Rev. Med. Suisse 2005, 1, 1728–1730. [Google Scholar]
  98. Talbot, G.H.; Bradley, J.; Edwards, J.E.; Gilbert, D.; Scheid, M.; Bartlett, J.G. Bad bugs need drugs: An update on the development pipeline from the Antimicrobial Availability Task Force of the Infectious Diseases Society of America. Clin. Infect. Dis. 2006, 42, 657–668. [Google Scholar] [CrossRef] [Green Version]
  99. Osei Sekyere, J. Candida auris: A systematic review and meta-analysis of current updates on an emerging multidrug-resistant pathogen. Microbiologyopen 2018, 7. [Google Scholar] [CrossRef] [Green Version]
  100. Ramirez-Garcia, A.; Pellon, A.; Rementeria, A.; Buldain, I.; Barreto-Bergter, E.; Rollin-Pinheiro, R.; De Meirelles, J.V.; Xisto, M.I.D.S.; Ranque, S.; Havlicek, V.; et al. Scedosporium and Lomentospora: An updated overview of underrated opportunists. Med. Mycol. 2018, 56, S102–S125. [Google Scholar] [CrossRef]
  101. Nobile, C.J.; Johnson, A.D. Candida albicans Biofilms and Human Disease. Annu. Rev. Microbiol. 2015, 69, 71–92. [Google Scholar] [CrossRef] [Green Version]
  102. Kathwate, G.H.; Shinde, R.B.; Mohan Karuppayil, S. Non-antifungal drugs inhibit growth, morphogenesis and biofilm formation in Candida albicans. J. Antibiot. 2021, 1–8. [Google Scholar] [CrossRef]
  103. 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. [Google Scholar] [CrossRef] [PubMed]
  104. Stepanović, S.; Vuković, D.; Ješić, M.; Ranin, L. Influence of acetylsalicylic acid (Aspirin) on biofilm production by Candida species. J. Chemother. 2004, 16, 134–138. [Google Scholar] [CrossRef] [PubMed]
  105. 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]
  106. Holanda, M.A.V.; Da Silva, C.R.; De A Neto, J.B.; Do Av Sá, L.G.; Do Nascimento, F.B.S.A.; Barroso, D.D.; Da Silva, L.J.; Cândido, T.M.; Leitão, A.C.; Barbosa, A.D.; et al. Evaluation of the antifungal activity in vitro of midazolam against fluconazole-resistant Candida spp. isolates. Future Microbiol. 2021, 16, 71–81. [Google Scholar] [CrossRef]
  107. Juvêncio da Silva, L.; Dias Barroso, F.D.; Vieira, L.S.; Carlos Mota, D.R.; da Silva Firmino, B.K.; Rocha da Silva, C.; de Farias Cabral, V.P.; Cândido, T.M.; Sá, L.G.D.A.V.; Barbosa da Silva, W.M.; et al. Diazepam’s antifungal activity in fluconazole-resistant Candida spp. and biofilm inhibition in C. albicans: Evaluation of the relationship with the proteins ALS3 and SAP5. J. Med. Microbiol. 2021. [Google Scholar] [CrossRef]
  108. Chavez-Dozal, A.A.; Jahng, M.; Rane, H.S.; Asare, K.; Kulkarny, V.V.; Bernardo, S.M.; Lee, S.A. In vitro analysis of flufenamic acid activity against Candida albicans biofilms. Int. J. Antimicrob. Agents 2014, 43, 86–91. [Google Scholar] [CrossRef] [Green Version]
  109. Do Amaral Valente Sá, L.G.; Da Silva, C.R.; De Andrade Neto, J.B.; Do Nascimento, F.B.S.A.; Barroso, F.D.D.; Da Silva, L.J.; De Farias Cabral, V.P.; Barbosa, A.D.; Silva, J.; Marinho, E.S.; et al. Antifungal activity of etomidate against growing biofilms of fluconazole-resistant Candida spp. strains, binding to mannoproteins and molecular docking with the ALS3 protein. J. Med. Microbiol. 2020, 69, 1221–1227. [Google Scholar] [CrossRef]
  110. 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] [Green Version]
  111. 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] [Green Version]
  112. Machado, C.B.; da Silva, C.R.; Barroso, F.D.; Campos, R.D.S.; Sá, L.V.; do Nascimento, F.A.; Cavalcanti, B.C.; Júnior, H.V.N.; Neto, J.A. In vitro evaluation of anti-fungal activity of tropicamide against strains of Candida spp. resistant to fluconazole in planktonic and biofilm form. J. Mycol. Med. 2021, 31. [Google Scholar] [CrossRef]
  113. 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] [PubMed]
  114. Lone, S.A.; Ahmad, A. Candida auris—the growing menace to global health. Mycoses 2019, 62, 620–637. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  115. Lockhart, S.R. Candida auris and multidrug resistance: Defining the new normal. Fungal Genet. Biol. 2019, 131. [Google Scholar] [CrossRef]
  116. Kean, R.; Ramage, G. Combined Antifungal Resistance and Biofilm Tolerance: The Global Threat of Candida auris. mSphere 2019, 4. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  117. Allaw, F.; Kara Zahreddine, N.; Ibrahim, A.; Tannous, J.; Taleb, H.; Bizri, A.R.; Dbaibo, G.; Kanj, S.S. First Candida auris Outbreak during a COVID-19 Pandemic in a Tertiary-Care Center in Lebanon. Pathogens 2021, 10, 157. [Google Scholar] [CrossRef] [PubMed]
  118. Cheng, Y.-S.; Roma, J.S.; Shen, M.; Fernandes, C.M.; Tsang, P.S.; Forbes, H.E.; Boshoff, H.; Lazzarini, C.; Del Poeta, M.; Zheng, W.; et al. Identification of Antifungal Compounds against Multidrug Resistant Candida auris Utilizing a High Throughput Drug Repurposing Screen. Antimicrob. Agents Chemother. 2021. [Google Scholar] [CrossRef]
  119. Eldesouky, H.E.; Lanman, N.A.; Hazbun, T.R.; Seleem, M.N. Aprepitant, an antiemetic agent, interferes with metal ion homeostasis of Candida auris and displays potent synergistic interactions with azole drugs. Virulence 2020, 11, 1466–1481. [Google Scholar] [CrossRef]
  120. 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. 2021, 65. [Google Scholar] [CrossRef]
  121. 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]
  122. 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. Methods Protoc. 2019, 2, 31. [Google Scholar] [CrossRef] [Green Version]
  123. Vermorel-Faure, O.; Lebeau, B.; Mallaret, M.R.; Michallet, M.; Brut, A.; Ambroise-Thomas, P.; Grillot, R. Food-related fungal infection risk in agranulocytosis. Mycological control of 273 food items offered to patients hospitalized in sterile units. Press. Med. 1993, 22, 157–160. [Google Scholar]
  124. Meis, J.F.; Chowdhary, A.; Rhodes, J.L.; Fisher, M.C.; Verweij, P.E. Clinical implications of globally emerging azole resistance in Aspergillus fumigatus. Philos. Trans. R. Soc. B Biol. Sci. 2016, 371. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  125. De Lucca, A.J. Harmful fungi in both agriculture and medicine. Rev. Iberoam. Micol. 2007, 24, 3–13. [Google Scholar] [CrossRef]
  126. Perlin, D.S.; Rautemaa-Richardson, R.; Alastruey-Izquierdo, A. The global problem of antifungal resistance: Prevalence, mechanisms, and management. Lancet Infect. Dis. 2017, 17, e383–e392. [Google Scholar] [CrossRef]
  127. Osherov, N.; Kontoyiannis, D.P. The anti- Aspergillus drug pipeline: Is the glass half full or empty? Med. Mycol. 2017, 55, 118–124. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  128. Vallières, C.; Singh, N.; Alexander, C.; Avery, S.V. Repurposing Nonantifungal Approved Drugs for Synergistic Targeting of Fungal Pathogens. ACS Infect. Dis. 2020, 6, 2950–2958. [Google Scholar] [CrossRef] [PubMed]
  129. May, R.C.; Stone, N.R.H.; Wiesner, D.L.; Bicanic, T.; Nielsen, K. Cryptococcus: From environmental saprophyte to global pathogen. Nat. Rev. Microbiol. 2016, 14, 106–117. [Google Scholar] [CrossRef]
  130. Spadari, C.D.C.; Wirth, F.; Lopes, L.B.; Ishida, K. New approaches for cryptococcosis treatment. Microorganisms 2020, 8, 613. [Google Scholar] [CrossRef]
  131. 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, 2018. [Google Scholar] [CrossRef] [Green Version]
  132. Tortorano, A.M.; Richardson, M.; Roilides, E.; van Diepeningen, A.; Caira, M.; Munoz, P.; Johnson, E.; Meletiadis, J.; Pana, Z.D.; Lackner, M.; et al. ESCMID and ECMM joint guidelines on diagnosis and management of hyalohyphomycosis: Fusarium spp., Scedosporium spp. and others. Clin. Microbiol. Infect. 2014, 20, 27–46. [Google Scholar] [CrossRef] [Green Version]
  133. Douglas, A.P.; Chen, C.-A.; Slavin, M.A. Emerging infections caused by non-Aspergillus filamentous fungi. Clin. Microbiol. Infect. 2016, 22, 670–680. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  134. Shinohara, M.M.; George, E. Scedosporium apiospermum: An emerging opportunistic pathogen that must be distinguished from Aspergillus and other hyalohyphomycetes. J. Cutan. Pathol. 2009, 36, 39–41. [Google Scholar] [CrossRef] [PubMed]
  135. 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] [PubMed]
  136. Torres, N.S.; Montelongo-Jauregui, D.; Abercrombie, J.J.; Srinivasan, A.; Lopez-Ribot, J.L.; Ramasubramanian, A.K.; Leung, K.P. Antimicrobial and antibiofilm activity of synergistic combinations of a commercially available small compound library with colistin against Pseudomonas aeruginosa. Front. Microbiol. 2018, 9. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  137. Dovizio, M.; Bruno, A.; Tacconelli, S.; Patrignani, P. Mode of action of aspirin as a chemopreventive agent. Recent Results Cancer Res. 2013, 191, 39–65. [Google Scholar] [CrossRef] [PubMed]
  138. Bjorkman, S.; Ove Elisson, L.; Gabrielsson, J. Pharmacokinetics of Quinacrine after Intrapleural Instillation in Rabbits and Man. J. Pharm. Pharmacol. 1989, 41, 160–163. [Google Scholar] [CrossRef]
  139. Steinfort, D.P.; Steinfort, C. Effect of long-term nebulized colistin on lung function and quality of life in patients with chronic bronchial sepsis. Intern. Med. J. 2007, 37, 495–498. [Google Scholar] [CrossRef]
  140. Rolain, J.M.; Baquero, F. The refusal of the Society to accept antibiotic toxicity: Missing opportunities for therapy of severe infections. Clin. Microbiol. Infect. 2016, 22, 423–427. [Google Scholar] [CrossRef] [Green Version]
Figure 1. Therapeutic targets in the fungal cell of the compounds listed in the review: known antifungals and molecules that have a potential for repurposing as antifungals.
Figure 1. Therapeutic targets in the fungal cell of the compounds listed in the review: known antifungals and molecules that have a potential for repurposing as antifungals.
Pharmaceuticals 14 00488 g001
Figure 2. Examples of approved and candidate drugs with activity against emerging MDR fungi.
Figure 2. Examples of approved and candidate drugs with activity against emerging MDR fungi.
Pharmaceuticals 14 00488 g002
Table 1. Current main antifungal agents: mechanisms of action, clinical indications, side effects and mechanisms of antifungal resistance. FCZ: fluconazole, ITZ: itraconazole, VRZ: voriconazole.
Table 1. Current main antifungal agents: mechanisms of action, clinical indications, side effects and mechanisms of antifungal resistance. FCZ: fluconazole, ITZ: itraconazole, VRZ: voriconazole.
Antifungal ClassesMechanisms of ActionClinical IndicationsSide EffectsMechanisms of ResistanceCommon Resistant Species
Polyenes
Amphotericin B
Nystatin
Ergosterol binding (membrane)
permeabilization by ion
channel formation
Cell content leakage
Invasive fungal infection
Topical Candida infections
Renal toxicity
Hypokalemia
Phlebitis
Immunoallergic reaction
Deficiencies in ERG2 and ERG3 genes
Ergosterol synthesis alteration
Modifications in membrane sterols
Changes of enzymatic activity or signaling pathways
Scedosporium spp.,
Candida lusitaniae,
Aspergillus terreus
Azoles
Fluconazole
Itraconazole
Voriconazole
Posaconazole
Efinaconazole
Isavuconazole
Inhibition of lanosterol
Ergosterol synthesis inhibition
Alteration of fungal membrane fluidity and agility
All invasive candidiasis
Cryptococcal meningitis
Aspergillus spp. infections
Digestive disturbancesCephalgias
Hepatotoxicity
Drug interactions (CYPP450)
Over expression of efflux pump’s function
ERG11 gene mutations inducing blockage in azoles binding
Up-regulation of enzyme target
Bypass pathway development by ERG3 gene mutation
FCZ: Candida krusei, Aspergillus spp., Scedosporium spp., Fusarium spp., Mucorales
ITZ: Fusarium spp.
VRZ: Mucorales
Echinocandins
Micafungin
Caspofungin
Anidulafungin
Inhibition of β-1,3-glucan synthase (β-GS)
Formation of a defective cell wall
Invasive candidiasis
Invasive aspergillosis (2nd intention)
Good overall toleranceMutations on FKS1 gene (encoding for a subunit of β-GS)
Decrease of affinity between drug and target
Cryptococcus spp.,
Fusarium spp.,
Scedosporium spp.,
Mucorales
5-fluorocytosineNucleoside analogue
Disruption of protein synthesis
Inhibition of DNA synthesis
Cryptococcosis
Invasive candidiasis if treatment failure
Always in association
Gastrointestinal troubles
Hepatotoxicity
Hematotoxicity
Mutations on FUR1 gene (encoding uracil phosphoribosyl transferase)
Mutations on FCY1 gene (encoding cytosine deaminase enzyme)
Ineffective against many filamentous fungi
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Peyclit, L.; Yousfi, H.; Rolain, J.-M.; Bittar, F. Drug Repurposing in Medical Mycology: Identification of Compounds as Potential Antifungals to Overcome the Emergence of Multidrug-Resistant Fungi. Pharmaceuticals 2021, 14, 488. https://0-doi-org.brum.beds.ac.uk/10.3390/ph14050488

AMA Style

Peyclit L, Yousfi H, Rolain J-M, Bittar F. Drug Repurposing in Medical Mycology: Identification of Compounds as Potential Antifungals to Overcome the Emergence of Multidrug-Resistant Fungi. Pharmaceuticals. 2021; 14(5):488. https://0-doi-org.brum.beds.ac.uk/10.3390/ph14050488

Chicago/Turabian Style

Peyclit, Lucie, Hanane Yousfi, Jean-Marc Rolain, and Fadi Bittar. 2021. "Drug Repurposing in Medical Mycology: Identification of Compounds as Potential Antifungals to Overcome the Emergence of Multidrug-Resistant Fungi" Pharmaceuticals 14, no. 5: 488. https://0-doi-org.brum.beds.ac.uk/10.3390/ph14050488

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