Next Article in Journal
Distribution of Phototrophic Purple Nonsulfur Bacteria in Massive Blooms in Coastal and Wastewater Ditch Environments
Next Article in Special Issue
Candida albicans Adaptation on Simulated Human Body Fluids under Different pH
Previous Article in Journal
Genome Sequencing of Pantoea agglomerans C1 Provides Insights into Molecular and Genetic Mechanisms of Plant Growth-Promotion and Tolerance to Heavy Metals
Previous Article in Special Issue
Recurrent Vulvovaginal Candidiasis: An Immunological Perspective
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Candida albicans Antifungal Resistance and Tolerance in Bloodstream Infections: The Triad Yeast-Host-Antifungal

by
Sofia Costa-de-Oliveira
1,2,* and
Acácio G. Rodrigues
1,2,3
1
Division of Microbiology, Department of Pathology, Faculty of Medicine, University of Porto, Al. Hernâni Monteiro, 4200-319 Porto, Portugal
2
Center for Research in Health Technologies and Information Systems (CINTESIS), R. Dr. Plácido da Costa, 4200-450 Porto, Portugal
3
Burn Unit, São João Hospital Center, Al. Hernâni Monteiro, 4200-319 Porto, Portugal
*
Author to whom correspondence should be addressed.
Microorganisms 2020, 8(2), 154; https://doi.org/10.3390/microorganisms8020154
Submission received: 1 December 2019 / Revised: 10 January 2020 / Accepted: 16 January 2020 / Published: 22 January 2020
(This article belongs to the Special Issue Candida albicans Virulence Factors and Its Pathogenecity)

Abstract

:
Candida albicans represents the most frequent isolated yeast from bloodstream infections. Despite the remarkable progress in diagnostic and therapeutic approaches, these infections continue to be a critical challenge in intensive care units worldwide. The economic cost of bloodstream fungal infections and its associated mortality, especially in debilitated patients, remains unacceptably high. Candida albicans is a highly adaptable microorganism, being able to develop resistance following prolonged exposure to antifungals. Formation of biofilms, which diminish the accessibility of the antifungal, selection of spontaneous mutations that increase expression or decreased susceptibility of the target, altered chromosome abnormalities, overexpression of multidrug efflux pumps and the ability to escape host immune defenses are some of the factors that can contribute to antifungal tolerance and resistance. The knowledge of the antifungal resistance mechanisms can allow the design of alternative therapeutically options in order to modulate or revert the resistance. We have focused this review on the main factors that are involved in antifungal resistance and tolerance in patients with C. albicans bloodstream infections.

Graphical Abstract

1. Introduction

Candida albicans coexists in humans as commensal without damage to the host, colonizing several body locations like the skin, genital tract, and gastro-intestinal tract [1]. Nevertheless, as an opportunistic pathogen, whenever the immune status of the host or its microbiota becomes disturbed, it can cause extensive mucosal colonization and local and/or systemic disease [2,3]. Along the years, in parallel to the advance of medical procedures, the incidence of bloodstream Candida infections increased as well as the associated mortality rate, being C. albicans the most frequent yeast isolated from patient biological samples [4,5,6,7,8,9].
The intensive care unit (ICU) setting provides C. albicans the opportunity for development of infection. Colonization of the skin and mucous membranes and the alteration or disruption of natural host barriers, like wounds, surgery, and the insertion of indwelling intravascular catheters are the main predisposing factors for Candida infections [6,10,11].
One of the main factors that contribute to the high mortality rate associated with Candidaemia is the difficulty in diagnosis, due to the nonspecific clinical symptoms of systemic fungal infection and delayed laboratorial detection methods, as well as the subsequent delay in initiation of adequate antifungal therapy [12,13]. Unlike antibacterial drugs, the array of available antifungals is somewhat scarcer. Azoles, polyenes, and echinocandins are the three main antifungal classes, being the last considered first-line therapy in many hospitals for the treatment of invasive candidiasis [14,15,16,17].
With the increase of clinical and/or microbiological antifungal resistance or tolerance, susceptibility tests play an ever-increasing role in the selection of antifungal drugs. Notably, correlation between in vitro susceptibility and treatment success is not always straightforward [18]. The in vivo conditions are significantly different of in vitro, in particular the microorganisms are often under the effect of both antifungal and non-antifungal drugs, as is the typical case of critical care patients [19,20,21]. Besides, C. albicans has particular traits and tricks that makes this yeast a true challenge for clinicians and researchers.
This review highlights the multiple attributes of C. albicans that may influence and promote antifungal resistance and tolerance.

2. Antifungal Drugs: Mechanisms of Action and Resistance

The battery of clinical antifungal agents available is somewhat limited, in contrast to antibacterial drugs. They arise from the number of drug targets in fungi, and its similarity to human eukaryotic cells. Nevertheless, pursuit for new cell targets, within the genomic era, has increased exponentially. Patients under long term antifungal prophylaxis or antifungal treatment display favorable conditions for the emergence of antifungal resistance [22].
Three types of antifungal resistance have been described: primary or intrinsic, exhibited before antifungal exposure, secondary or acquired, and clinical resistance [23]. Secondary or acquired resistance develops following exposure to an antifungal agent and can be either reversible, due to transient adaptation, or persistent because of one or several genetic alterations [23].
The main antifungals used in invasive candidiasis treatment, as well as the main mechanisms of action and resistance are summarized in Table 1 and detailed in the next subsections. Clinical resistance will be address in Section 3.

2.1. Polyenes

Polyenes belong to a class of natural compounds with a heterocyclic amphipathic molecule (one hydrophilic charged side of the molecule and one hydrophobic, uncharged side). They target ergosterol in the fungal membrane by inserting into the lipid bilayers and creating pores that disrupt plasma membrane integrity, allowing small molecules to diffuse across the membrane, resulting in cell death [24]. Nystatin and amphotericin B belong to this group. Nystatin has a spectrum of activity slightly narrower than that of amphotericin B but is active against a number of species yeasts and molds [25]. The topical use of nystatin is considered the most common route of administration and plays an important role in the prophylaxis of oral and systemic candidiasis in full-term and premature newborns, infants, and immunocompromised patients [26,27]. Amphotericin B is still considered the gold standard in the treatment of most fungal infections, especially severe invasive infections [28]. However, amphotericin is toxic to mammalian cells, particularly causing nephrotoxicity. To overcome its toxicity, a variety of reformulated versions has been introduced. Lipid formulations of amphotericin B are better tolerated than amphotericin B deoxycolate [29]. Although having broad-spectrum activity against most fungi, lipid formulations are very expensive, limiting the use to second-line or salvage therapy.
Resistance to amphotericin B is quite unusual and most often results from mutations in the ERG3 gene (which encodes a C-5 sterol desaturase, an enzyme involved in ergosterol biosynthesis), which lower the concentration of ergosterol in the fungal membrane [30]. Resistance to amphotericin B may also be mediated by increasing catalase activity, with decreasing susceptibility to oxidative damage [31]. Among C. albicans isolates, amphotericin resistance is still very rare [32,33,34].

2.2. Pyrimidine Analogues

5-Flucytosine is the only representative of this class of antifungals. It acts through conversion to 5-fluorouracil by a cytosine deaminase, which is incorporated into DNA and RNA, inhibiting cellular function and division [24]. Since most filamentous fungi lack cytosine deaminase, the spectrum of flucytosine is restricted to pathogenic yeasts. 5-Flucytosine should be used in combination with other antifungal agents namely amphotericin B, rather than in monotherapy, because resistance develops at high frequency [24]. Resistance among Candida correlates with mutations in the enzyme uracil phosphoribosyltransferase (Fur1p) that turns, unable the conversion of 5-fluorouracil to 5-fluorouridine monophosphate [35].

2.3. Triazoles

Triazoles are the largest class of antifungal drugs in clinical use and have been deployed for approximately three decades. They are heterocyclic synthetic compounds that inhibit the fungal cytochrome P450 14α-lanosterol demethylase, encoded by the ERG11 gene (also known as CYP51) which catalyzes the late step of ergosterol biosynthesis. These drugs bind through a nitrogen group in their five-membered azole ring to the heme group in the target protein and block demethylation of the C-14 of lanosterol, leading to the substitution of methylated sterols in the membrane. Inhibition of this enzyme results in decreased membrane ergosterol content and accumulation of toxic methylated intermediates, with resultant disruption of fungal cell membrane function, growth inhibition, and, in some cases, cell death [35,36,37]. Triazole antifungal activity is generally fungistatic against Candida spp., but fungicidal against Aspergillus.
The triazoles include fluconazole, itraconazole, voriconazole, posaconazole, and isavuconazole. Given its excellent safety and low cost profile and the proven efficacy for the treatment of invasive candidiasis, fluconazole remains one of the most commonly used antifungal agents [38]. Voriconazole is a second-generation triazole that is active against all Candida species and has a broad spectrum of activity and, like itraconazole, is also fungicidal against some isolates of filamentous fungi [39]. Posaconazole differs in structure from the compact triazoles (fluconazole and voriconazole) in part by its extended side chain (a feature held in common with itraconazole); however, it displays a dioxolane ring altered to a tetrahydrofuran [36,40]. The structural differences between the azoles might seem small, but they dictate its antifungal potency and spectrum, bioavailability, drug interaction, and toxic potential. Posaconazole is currently only available as oral formulation, and it must be taken with food or a nutritional supplement, somewhat limiting its usefulness. The drug is well tolerated, with an overall safety profile comparable to that of fluconazole [40]. Isavuconazole is an expanded-spectrum triazole with excellent in vitro activity against yeasts and molds [41,42,43]. It is consider, as well as voriconazole, a gold standard drug for invasive aspergillosis in patients with underlying hematological malignancies [44].
The major mechanism responsible for high level of azole resistance is the overexpression of cell membrane efflux pumps [45,46]. Two classes of pumps are responsible for lowering the accumulation of azoles inside the yeast cell by actively translocating compounds across cell membrane: ABC pumps and the major facilitator (MF) transporters [1,11,13,47,48,49,50,51,52,53,54,55].
The ABC pumps, also called ATP-binding cassette, use the hydrolysis of ATP as an energy source. They have low specificity since they accept as substrates azoles but also a wide range of compounds [56]. The most frequently encountered triazole resistance mechanism among clinical isolates of C. albicans is the upregulation or overexpression of mainly CDR1 and CDR2 genes [57,58,59,60]. Their expression is regulated by the zinc finger transcription factor Tac1, which binds to the drug response element (DRE) found in their promoter [61]. Loss of heterozygosity of specific genomic regions, the increase of chromosome copy number or chromosome aneuploidies have been associated with azole resistance [62,63,64]. CDR expression is increase by gain-of-function mutations in Tac1p, with high level of fluconazole resistance occurring when this mutation is couple with loss of heterozygosity [62]. Interestingly, TAC1 is located in the left arm of chromosome 5 (Chr5), the same chromosome where mating-type-locus (MTL) is located [65]. C. albicans exhibits two MTL alleles, MTLa and MTLα, and the loss of heterozigoty at MTL locus is frequently associated with homozygosity at the TAC1 and ERG11 loci. Some authors described this homozygosity to be related to antifungal resistance [62,63,66,67]. However, others demonstrated homozygosity at MTL to be infrequent among clinical isolates and that it does not influence directly antifungal resistance [64,68,69].
The second main class of multidrug transporters involved in azole resistance is the MF class. MDR1 gene is involved specifically in resistance to fluconazole rather than to other azoles and uses the proton motive force of the membrane as an energy source [70,71]. The multidrug resistant regulator, Mrr1, is the transcription factor that controls the expression and is upregulated with MDR1 in drug resistant clinical isolates [55,71]. The gain-of-function in the transcription factor Mrr1p, followed by loss of heterozygosity, represents the main cause of MDR1 overexpression in fluconazole resistant C. albicans strains [72].
Another mechanism that operates in order to overcome the effect of the drug in the yeast cell is the alteration of the target enzyme Erg11, where at least 12 mutations have been associated with azole resistance, avoiding the binding of the drug to the target [73,74]. In C. albicans, point mutations in ERG11 resulting in amino acid substitutions (G464S) have been associated with fluconazole resistance [75]. Upregulation of ERG11 due to the amplification of the copy number of the gene is another approach used by the fungal cell in order to overcome antifungal action [76]. ERG11 overexpression can be achieved through mutations in the transcription factor Upc2 in C. albicans [77,78,79]. This transcription factor binds to the azole-responsive enhancer element (ARE) in the ERG11 promoter [80]. Upc2 also binds to two distinct regions on its own promoter to autoregulate expression during azole exposure [81].

2.4. Echinocandins

Echinocandins are at present considered the first-line therapy for Candida invasive infections [17]. These compounds are fungicidal in vitro against yeasts. Three agents are presently available for clinical use: caspofungin, micafungin, and anidulafungin. They inhibit β-(1,3) glucan synthase, an enzyme complex that is located in the plasma membrane of fungal cells [24,82,83,84]. This enzyme has a minimum of two subunits, Fks1, the catalytic subunit, and Rho, a GTP-binding protein that regulate the activity of the glucan synthase [84]. They are responsible for the production of β-(1,3) glucan which is essential for fungi as it represents one of the major components of the fungal cell wall [84]. The safety profile of echinocandins is excellent, with few reported adverse events and drug interactions. Despite their considerably greater cost, echinocandins are replacing fluconazole as the antifungal of choice in the Intensive Care Unit (ICU) setting [38].
Recent studies have shown that echinocandins are efficacious and safe, supporting the recommendation as a first-line option in case of bloodstream infections [17].
Echinocandin resistance in Candida spp. has been attributed to mutations in the FKS1 gene, the catalytic subunit of β-(1,3)-glucan synthase, and in a lesser extent in FKS2, resulting in amino acid substitutions in conserved regions hot spot 1 (HS1) and hot spot 2 (HS2) [15]. These mutations turn the mutant enzyme 50- to 3000-fold less sensitive to the drug, being amino acid changes at Ser645 as the most pronounced resistant phenotype [85,86]. Acquired mutations in FKS1 and FKS2 genes have been predominantly found at position 645 (Serine), S645F (serine to phenylalanine), S645P (serine to proline), and S645Y (serine to tyrosine), and have now been identified in a wide range of Candida clinical isolates [84,85,87]. In C. albicans mutations encoding an FKS1, HS1 alteration S645P, S629P, S654P, F641S, and F641I and in FKS1, HS2 alteration R1361G are common [88,89]. Nevertheless, the prevalence of Fks mutations in geographically distinct clinical isolates remains low [86]. A survey of C. albicans and C. glabrata bloodstream isolates in Switzerland showed that echinocandin resistance remained at a low level despite a significant increase in echinocandin use and was mainly associated with individual pre-echinocandin exposure of prolonged duration [89]. Hot spot mutations are more likely to confer resistance to caspofungin than to anidulafungin or micafungin. Such fact suggests that caspofungin could be less potent than the other two drugs [86,90]. However, these differences in echinocandin potency are abolished in the presence of human serum and therefore cross-resistance is likely to occur in vivo [91,92].

3. Factors Contributing to Candida albicans Clinical Resistance

For clinicians, the three main issues of concern about antifungal resistance are: how commonly does it occurs, how easy it is to induce through inappropriate usage of antifungal agents, and how often does it result in treatment failure. Clinical resistance may be defined as the persistence or progression of an infection despite appropriate antifungal therapy with an in vitro susceptibility of the organism [23,93]. Given this definition, it is reasonable to affirm that clinical resistance and antifungal tolerance are intrinsically related. The ability of the pathogen to tolerate drug concentrations above the minimal inhibitory concentration (MIC) values is defined as antifungal tolerance, which may promote the acquisition of antifungal resistance [93].
C. albicans has been the subject of extensive research in order to unveil the mechanisms governing fungal virulence and drug tolerance. Many factors may contribute to clinical resistance and to the discrepancy between the laboratory susceptibility pattern and the clinical outcome (Figure 1). The antifungal efficacy lives in the Bermuda triangle that encompasses patient, drug, and yeast factors that ultimately are responsible for a poor clinical outcome. Patient pharmacogenomics, which can influence drug absorption, distribution and metabolism, its immunological status, and the underlying disease are additional important factors to be considered when managing individual patients [94]. Clinical resistance mechanism involving the ability of C. albicans to better tolerate and survive high concentrations of antifungal drugs will be the focus of this section.

3.1. The Tolerance Pathways

Under stressing conditions like antifungal exposure, Candida cells may exploit several cellular responses, such as development of mutations, overexpression of multidrug efflux pumps, modulation of the cAMP protein kinase A (PKA) or Ca2+-calmodulin-calcineurin pathways [95].
The stress responses mediating triazole resistance most often involve the cyclic AMP (cAMP)-protein kinase A (PKA) signaling pathway [96,97]. The cAMP-PKA pathway in C. albicans is likely required to facilitate the recovery process and resume growth after various stress conditions, like fluconazole exposure. CDC35, encoding the adenyl cyclase enzyme, and CAP, the cAMP-associated protein, are involved in azole tolerance. Disruption of either gene results in hypersusceptibility to azoles that can be partially reverse by the addition of cAMP [96].
Calcineurin is a heterodimeric phosphatase that is involved in calcium-dependent signaling and regulation of diverse cellular processes [35,98,99]. In C. albicans, it is involved in virulence, membrane stress response, and is required for the survival in the presence of antifungal drugs, specifically fluconazole [100]. Jia et al. found that calcium-activated-calcineurin, through its target Rta2p and the transcriptional factor Crz1, dramatically reduced the efficacy of fluconazole against C. albicans, both in vitro and in vivo [101].
The molecular chaperone heat shock protein 90 (Hsp90) stabilizes calcineurin, enabling calcineurin-dependent stress responses that are required to survive the exposure to fluconazole and echinocandins in C. albicans [102]. Cell wall integrity signaling mediated via protein kinase C (PKC), the protein phosphatase calcineurin, and Hsp90, is very important in enabling echinocandin drug tolerance and compensatory mechanisms such as upregulation of chitin synthesis [102]. Echinocandin treatment may trigger cell wall salvage mechanisms producing physiological alterations that decrease the susceptibility to these antifungal agents [103]. The inhibition of the β-(1,3)-glucan synthesis leads to a compensatory increase in chitin synthesis mediated by the PKC cell wall integrity MAP kinase, Ca2+-calcineurin and high osmolarity glycerol response (HOG) signaling pathways [104]. Such an increase in chitin content is responsible for the paradoxical growth (PG) or “eagle effect” and occurs most frequently with caspofungin than with anidulafungin and micafungin, most frequently at concentrations well above the MIC level or of sub-MIC level [103,105,106,107,108]. It was also demonstrated that high-chitin C. albicans cells are less susceptible to caspofungin [109]. After two hours of caspofungin exposure, chitin content increase significantly, especially in C. parapsilosis, C. tropicalis, and C. albicans, strains that showed PG in microdilution assays [110]. Although some authors state that the PG is eliminated by human serum thus being unlikely to occur in vivo, others have shown that the effect can in fact occur in vivo after exposure to echinocandin treatment [111,112].
A recent study implicated the C. albicans transcription factor Cas5, a key transcriptional regulator of cell wall stress response, in governing echinocandin tolerance and an attractive target for antifungal development [113].
Yang and co-workers found mechanisms of caspofungin tolerance in C. albicans that may be involved in earlier tolerance development, which involve rearrangements of chromosome 5 (Ch5) [114,115]. Changes of expression of three genes residing on the right arm of Ch5: CHT2, implicated in cell remodeling, and CNB1 and MID1, which belong to the calcineurin stress response-signaling pathway were found [115]. Multiple genes may be regulated and involved in order to increase the amount of chitin like downregulation of CHT2, PGA4, and CSU51 on the monosomic Ch5, as well as upregulation of CHS2 and CHS3 for the chitin synthases encoded outside Ch5 [116].
For the immune system of infected hosts, both chitin and β-glucan act as pathogen-associated molecular patterns (PAMPs), especially β-glucans that are expose on the cell surface and its recognition is mediated by Dectin-1 [117,118]. Echinocandin promotes the efficiency of phagocyte killing, as the inhibition of β-glucan synthase results in a pathogen that is more recognizable by host cells [118]. Thus, this will increase the dectin-1-mediated inflammatory response of macrophages to C. albicans because of the exposure of the normally concealed branched glucan polymers [117,118]. However, the inhibition of glucan production in the cell wall may trigger the salvage pathway and subsequent increase of chitin content, influencing the dectin-1 receptor recognition, and generating a decrease in inflammatory response [119]. C. albicans cells with high chitin content stimulated a lower level of cytokine response than the ones expressing normal chitin levels [118].

3.2. Cell Plasticity

The success of C. albicans as a pathogen depends largely on its ability to generate diversity not only at the genetic level but also at the morphological and physiological level [120]. C. albicans is considered pleomorphic due to its ability to switch from yeast to hyphal or pseudohyphal form [120,121]. Morphogenic changes are coupled to biofilm formation, which plays an important role in virulence (Figure 1). C. albicans can produce biofilm on medical implants, like indwelling vascular catheters, and its formation acts as a physical barrier, protecting the underlying cells of antifungal drugs, hence lowering the available drug concentration [122,123]. Candida biofilm infections can have devastating consequences, progressing to bloodstream invasive infections since cells are usually resistant to antifungal drugs and to the host immune system. Biofilm cell development follows closely an intricate gene regulatory network genes and complex transcriptional factors like Efg1, Tec1, Bcr1, Brg1, Ndt80, and Rob1, which are involved in cell surface regulation, hyphal formation, and development and virulence expression [124,125].
Candida biofilm cell communities strike back antifungal action through multifactorial mechanisms [126,127,128]. Efflux pumps such as the ATP binding cassette transporters (CDR1 and CDR2) and major facilitator transporter (MDR1) are involved in azole resistance by C. albicans biofilm, especially at the early stages of its formation rather than in mature biolfilms [127,129]. The extracellular matrix of biofilm, a polymeric material that promotes adherence and protects cells from hostile environments, is also a major contributor to antifungal resistance in C. albicans. Azoles and conventional amphotericin B are ineffective against C. albicans biofilms [130,131]. In contrast, liposomal amphotericin B and echinocandins have showed to exhibit antifungal activity upon C. albicans biofilms [131,132,133]. Since the main mechanism of action is to inhibit β-(1,3)-glucan synthesis, echinocandins are able to abolish the excess of extracellular matrix production, turning the antifungal effect more effective.
The mitochondrial respiratory pathway can regulate the metabolic behavior contributing to fitness and flexibility of Candida organisms in response to external challenges [134,135,136]. The existence of an alternative respiratory pathway, alternative oxidase (AOX), present in some Candida species, especially C. albicans has been implicated in reduced susceptibility to azoles and resistance to oxidative stress [137,138]. Although the inhibition of this alternative respiratory pathway might seem an attractive strategy to combat Candida infections, some authors state that this inhibition may result in different outcomes in turns of other drug susceptibilities, in which the link between mitochondrial respiration and cell wall regulation may vary among species [139,140].

3.3. Effect of Concomitant Therapy

Antibacterial drugs administration in high-risk patients is one of the major factors that contributes to the development of candidiasis. They cause an imbalance in the fungal microbiome, increasing colonization and proliferation of yeasts [147]. Concomitant medications administered to patients can influence the pharmacodynamics of the antifungals [19,146]. Fluoroquinolones antagonize fluconazole activity against C. albicans strains, whether rifampicin can induce the expression of MDR1 pumps [145,146]. While the effect of other medications, some of them lifesaving in the case of critical care patients remains to be elucidated, it has been demonstrated in vitro that albumin and propofol impair the antifungal susceptibility profile [19,20,21].

4. Strategies to Overcome Antifungal Resistance and Tolerance

The knowledge encompassing the mechanism of antifungal resistance brought by the genomic era supports the development of novel therapeutic strategies in order to bypass antifungal drug resistance. The principal cell mechanism of antifungal resistance is the active transport of drugs out of the cell by efflux pumps, a mechanism expressed not only by yeasts but also by human cells [52,53,54,55,59,94]. The main strategy to reduce efflux impact involves the maintenance of a high antifungal concentration inside the cell, at its site of action. The simplest approach would be the use of antifungals that are not substrate of efflux pumps, like amphotericin B or echinocandins, which given its respective hydrophobicity and size, do not interact with the efflux pump [83,148].
The second approach would be the development of inhibitors or chemosensitizers of efflux, impairing the target, activity, by blocking access to the binding site, or even the efflux pump transcription.
In humans, one of the factors that is responsible for the failure of cancer therapy are ATP-dependent drug efflux pumps, such as P-glycoprotein (P-gp) [149]. P-gp substrates such as FK506 [150,151,152] or cyclosporine A (CsA) are immunosuppressors that are able to inhibit efflux [153]. They act similarly in C. albicans, inhibiting the calcineurin-mediated azole tolerance by binding to small, abundant, conserved binding proteins called immunophilins. CsA binds to cyclophilin A (Cyp1p) and FK506 to FKBP12, forming protein-drug complexes that inhibit calcineurin [52,100,154]. By inhibiting calcineurin, these compounds act synergistically with azoles [100,155,156]. While FK506 and CsA chemosensitize C. albicans cells to azoles, turning the azoles fungicidal, they are simultaneously immunosuppressive drugs, which turns its administration problematic to immunosuppressed patients. Nevertheless, the inhibition of calcineurin-mediated azole tolerance is still a potential therapeutic approach [157]. Non-immunosuppressive analogs could inhibit fungal calcineurin by exploiting structural differences between the human and the fungal targets [157].
Ibuprofen ([2-(4-isobutylphenyl)-propionic acid]) has been described to act synergistically with pyrazinamide [158], fluconazole [59,159,160], and amphotericin B [161] in fungi. In C. albicans expressing CDR efflux pumps, ibuprofen induced azole intracellular accumulation, changing the resistant phenotype to susceptible [59,160]. In a murine model of systemic infection, ibuprofen acts synergically with fluconazole against a fluconazole-resistant strain, drastically reducing the fungal burden and morbidity [162]. This potent anti-inflammatory, non-steroidal drug might play important role in future therapeutic strategies.
Kaempferol, a flavonoid, exhibited a synergistic effect in C. albicans fluconazole resistant strains, by decreasing CDR1, CDR2, and MDR1 expression [163].
Another helpful strategy would be the design of inhibitors that could act indirectly on efflux, de-energizing the ATP or H+ dependent transporter, by lowering the cytoplasmic ATP concentration or by depleting the electrochemical potential of the plasma membrane, respectively [56,164]. However, altering ATP and membrane potential could compromise other cellular metabolic activities. Alternatively, the promotion of antifungal uptake could also be a strategy to overcome antifungal resistance due to efflux. Dubikovskaya et al. showed that the inclusion of multiple arginine residues (octaarginine (R8)) in human anticancer drugs enhances the delivery to its intracellular targets [165], an approach that has already been tried in yeasts [164].
Nikkomycin Z is a peptidyl nucleoside that functions as a substrate analogue and inhibits chitin synthase at its catalytic site [166]. To prevent chitin increase, Stevens combined in vitro nikkomycin Z with caspofungin and a synergistic effect was obtained [167]. In A. fumigatus, it was reported that chitin cell wall content was not affected by nikkomycin Z treatment but was markedly affected by caspofungin plus nikkomycin Z [117,168]. In addition, the combination of the two antifungals led to changes in cell wall chitin and β-glucan content. This synergistic effect was also observed in Alternaria infectoria [169]. In the case of C. albicans biofilm, the combination of echinocandins with nikkomycin Z demonstrated to be synergic, causing an extended cells death [170,171].

5. Conclusions

The last 30 years of medical advances led to a significant increase of life-threatening Candida infections. The high incidence and mortality rates associated with invasive candidiasis have remained unchanged for more than a decade despite the advances in the field of antifungal therapy. Such infections could be treated more effectively if faster and more specific diagnostic and therapeutic alternatives were available. Preventive strategies targeting patients with a high-risk profile, the development of new diagnostic tools for early identification of fungal species, its susceptibility profile, also including innate resistant species or those that are more prone to develop multidrug resistance, especially in patients submitted to long-term therapy are of utmost importance.
Given the association between antifungal exposure and the development of resistance, prophylaxis must be selectively restricted to high-risk patients. Apparently, there is no antifungal that is immune to the development of acquired resistance. It is essential that laboratories start performing routinely in vitro susceptibility testing especially in isolates from invasive infections, recovered from patients receiving antifungal prophylaxis and in strains isolated from patients who do not respond promptly to therapy.
Despite the yet low in vitro resistance of C. albicans to antifungal drugs, this yeast represents a challenge to clinicians and must not be underestimated.
The medical complexity of patients taken together with the intricate cellular mechanism involved in drug resistance and tolerance makes the pursuit of more effective solutions mandatory.

Author Contributions

S.C.-d.-O. and A.G.R. wrote and revised the manuscript. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by Astellas and Grupo de Infeção e Sepsis from Centro Hospitalar São João, Porto Portugal (Clinical Mycology Research grant).

Conflicts of Interest

The authors declare no conflict of interests.

References

  1. Calderone, R. (Ed.) Candida and Candidiasis; ASM Press: Washington, DC, USA, 2002. [Google Scholar]
  2. Cannon, R.D.; Holmes, A.R.; Mason, A.B.; Monk, B.C. Oral Candida: Clearance, colonization, or candidiasis? J. Dent. Res. 1995, 74, 1152–1161. [Google Scholar] [CrossRef] [PubMed]
  3. Casadevall, A.; Pirofski, L.A. Accidental virulence, cryptic pathogenesis, martians, lost hosts, and the pathogenicity of environmental microbes. Eukaryot. Cell 2007, 6, 2169–2174. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  4. Segal, B.H.; Almyroudis, N.G.; Battiwalla, M.; Herbrecht, R.; Perfect, J.R.; Walsh, T.J.; Wingard, J.R. Prevention and early treatment of invasive fungal infection in patients with cancer and neutropenia and in stem cell transplant recipients in the era of newer broad-spectrum antifungal agents and diagnostic adjuncts. Clin. Infect. Dis. Off. Publ. Infect. Dis. Soc. Am. 2007, 44, 402–409. [Google Scholar] [CrossRef] [PubMed]
  5. Pfaller, M.A.; Diekema, D.J. Epidemiology of invasive candidiasis: A persistent public health problem. Clin. Microbiol. Rev. 2007, 20, 133–163. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  6. Costa-de-Oliveira, S.; Pina-Vaz, C.; Mendonca, D.; Goncalves Rodrigues, A. A first Portuguese epidemiological survey of fungaemia in a university hospital. Eur. J. Clin. Microbiol. Infect. Dis. Off. Publ. Eur. Soc. Clin. Microbiol. 2008, 27, 365–374. [Google Scholar] [CrossRef]
  7. Montagna, M.T.; Lovero, G.; Borghi, E.; Amato, G.; Andreoni, S.; Campion, L.; Lo Cascio, G.; Lombardi, G.; Luzzaro, F.; Manso, E.; et al. Candidemia in intensive care unit: A nationwide prospective observational survey (GISIA-3 study) and review of the European literature from 2000 through 2013. Eur. Rev. Med. Pharmacol. Sci. 2014, 18, 661–674. [Google Scholar]
  8. Quindos, G. Epidemiology of candidaemia and invasive candidiasis. A changing face. Rev. Iberoam. Micol. 2014, 31, 42–48. [Google Scholar] [CrossRef]
  9. Pappas, P.G.; Lionakis, M.S.; Arendrup, M.C.; Ostrosky-Zeichner, L.; Kullberg, B.J. Invasive candidiasis. Nat. Rev. Dis. Primers 2018, 4, 18026. [Google Scholar] [CrossRef]
  10. Dimopoulos, G.; Karabinis, A.; Samonis, G.; Falagas, M.E. Candidemia in immunocompromised and immunocompetent critically ill patients: A prospective comparative study. Eur. J. Clin. Microbiol. Infect. Dis. Off. Publ. Eur. Soc. Clin. Microbiol. 2007, 26, 377–384. [Google Scholar] [CrossRef]
  11. Blumberg, H.M.; Jarvis, W.R.; Soucie, J.M.; Edwards, J.E.; Patterson, J.E.; Pfaller, M.A.; Rangel-Frausto, M.S.; Rinaldi, M.G.; Saiman, L.; Wiblin, R.T.; et al. Risk factors for candidal bloodstream infections in surgical intensive care unit patients: The NEMIS prospective multicenter study. The National Epidemiology of Mycosis Survey. Clin. Infect. Dis. Off. Publ. Infect. Dis. Soc. Am. 2001, 33, 177–186. [Google Scholar] [CrossRef] [Green Version]
  12. Nolla-Salas, J.; Sitges-Serra, A.; Leon-Gil, C.; Martinez-Gonzalez, J.; Leon-Regidor, M.A.; Ibanez-Lucia, P.; Torres-Rodriguez, J.M. Candidemia in non-neutropenic critically ill patients: Analysis of prognostic factors and assessment of systemic antifungal therapy. Study Group of Fungal Infection in the ICU. Intensive Care Med. 1997, 23, 23–30. [Google Scholar] [CrossRef] [PubMed]
  13. Garey, K.W.; Rege, M.; Pai, M.P.; Mingo, D.E.; Suda, K.J.; Turpin, R.S.; Bearden, D.T. Time to initiation of fluconazole therapy impacts mortality in patients with candidemia: A multi-institutional study. Clin. Infect. Dis. Off. Publ. Infect. Dis. Soc. Am. 2006, 43, 25–31. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  14. Pappas, P.G.; Kauffman, C.A.; Andes, D.; Benjamin, D.K., Jr.; Calandra, T.F.; Edwards, J.E., Jr.; Filler, S.G.; Fisher, J.F.; Kullberg, B.J.; Ostrosky-Zeichner, L.; et al. Clinical practice guidelines for the management of candidiasis: 2009 update by the Infectious Diseases Society of America. Clin. Infect. Dis. Off. Publ. Infect. Dis. Soc. Am. 2009, 48, 503–535. [Google Scholar] [CrossRef] [Green Version]
  15. Park, S.; Kelly, R.; Kahn, J.N.; Robles, J.; Hsu, M.J.; Register, E.; Li, W.; Vyas, V.; Fan, H.; Abruzzo, G.; et al. Specific substitutions in the echinocandin target Fks1p account for reduced susceptibility of rare laboratory and clinical Candida sp. isolates. Antimicrob. Agents Chemother. 2005, 49, 3264–3273. [Google Scholar] [CrossRef] [Green Version]
  16. Wiederhold, N.P.; Lewis, R.E. The echinocandin antifungals: An overview of the pharmacology, spectrum and clinical efficacy. Expert Opin. Investig. Drugs 2003, 12, 1313–1333. [Google Scholar] [CrossRef] [PubMed]
  17. Pappas, P.G.; Kauffman, C.A.; Andes, D.R.; Clancy, C.J.; Marr, K.A.; Ostrosky-Zeichner, L.; Reboli, A.C.; Schuster, M.G.; Vazquez, J.A.; Walsh, T.J.; et al. Clinical Practice Guideline for the Management of Candidiasis: 2016 Update by the Infectious Diseases Society of America. Clin. Infect. Dis. 2016, 62, e1–e50. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  18. Hospenthal, D.R.; Murray, C.K.; Rinaldi, M.G. The role of antifungal susceptibility testing in the therapy of candidiasis. Diagn. Microbiol. Infect. Dis. 2004, 48, 153–160. [Google Scholar] [CrossRef] [PubMed]
  19. Costa-de-Oliveira, S.; Araujo, R.; Silva-Dias, A.; Pina-Vaz, C.; Rodrigues, A.G. Propofol lipidic infusion promotes resistance to antifungals by reducing drug input into the fungal cell. BMC Microbiol. 2008, 8, 9. [Google Scholar] [CrossRef] [Green Version]
  20. Rodrigues, A.G.; Araujo, R.; Pina-Vaz, C. Human albumin promotes germination, hyphal growth and antifungal resistance by Aspergillus fumigatus. Med. Mycol. 2005, 43, 711–717. [Google Scholar] [CrossRef] [Green Version]
  21. Pedrosa, A.F.; Rodrigues, A.G. Candidemia in burn patients: Figures and facts. J. Trauma Acute Care Surg. 2011, 70, 498–506. [Google Scholar] [CrossRef]
  22. Brion, L.P.; Uko, S.E.; Goldman, D.L. Risk of resistance associated with fluconazole prophylaxis: Systematic review. J. Infect. 2007, 54, 521–529. [Google Scholar] [CrossRef] [PubMed]
  23. Kanafani, Z.A.; Perfect, J.R. Antimicrobial resistance: Resistance to antifungal agents: Mechanisms and clinical impact. Clin. Infect. Dis. 2008, 46, 120–128. [Google Scholar] [CrossRef] [Green Version]
  24. Peman, J.; Canton, E.; Espinel-Ingroff, A. Antifungal drug resistance mechanisms. Expert Rev. Anti-Infect. Ther. 2009, 7, 453–460. [Google Scholar] [CrossRef] [PubMed]
  25. Chandrasekar, P. Management of invasive fungal infections: A role for polyenes. J. Antimicrob. Chemother. 2011, 66, 457–465. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  26. Howell, A.; Isaacs, D.; Halliday, R.; Australasian Study Group For Neonatal Infections. Oral nystatin prophylaxis and neonatal fungal infections. Arch. Dis. Child. Fetal Neonatal Ed. 2009, 94, F429–F433. [Google Scholar] [CrossRef] [PubMed]
  27. Gotzsche, P.C.; Johansen, H.K. Nystatin prophylaxis and treatment in severely immunodepressed patients. Cochrane Database Syst. Rev. 2002. [Google Scholar] [CrossRef]
  28. Lacerda, J.F.; Oliveira, C.M. Diagnosis and treatment of invasive fungal infections focus on liposomal amphotericin B. Clin. Drug Investig. 2013, 33, S5–S14. [Google Scholar] [CrossRef]
  29. Wingard, J.R.; White, M.H.; Anaissie, E.; Raffalli, J.; Goodman, J.; Arrieta, A. A randomized, double-blind comparative trial evaluating the safety of liposomal amphotericin B versus amphotericin B lipid complex in the empirical treatment of febrile neutropenia. L Amph/ABLC Collaborative Study Group. Clin. Infect. Dis. 2000, 31, 1155–1163. [Google Scholar] [CrossRef] [Green Version]
  30. Kelly, S.L.; Lamb, D.C.; Kelly, D.E.; Manning, N.J.; Loeffler, J.; Hebart, H.; Schumacher, U.; Einsele, H. Resistance to fluconazole and cross-resistance to amphotericin B in Candida albicans from AIDS patients caused by defective sterol delta5,6-desaturation. FEBS Lett. 1997, 400, 80–82. [Google Scholar] [CrossRef] [Green Version]
  31. Sokol-Anderson, M.L.; Brajtburg, J.; Medoff, G. Amphotericin B-induced oxidative damage and killing of Candida albicans. J. Infect. Dis. 1986, 154, 76–83. [Google Scholar] [CrossRef]
  32. Castanheira, M.; Deshpande, L.M.; Davis, A.P.; Rhomberg, P.R.; Pfaller, M.A. Monitoring Antifungal Resistance in a Global Collection of Invasive Yeasts and Molds: Application of CLSI Epidemiological Cutoff Values and Whole-Genome Sequencing Analysis for Detection of Azole Resistance in Candida albicans. Antimicrob. Agents Chemother. 2017, 61. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  33. Arendrup, M.C. Update on antifungal resistance in Aspergillus and Candida. Clin. Microbiol. Infect. 2014, 20, 42–48. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  34. Faria-Ramos, I.; Neves-Maia, J.; Ricardo, E.; Santos-Antunes, J.; Silva, A.T.; Costa-de-Oliveira, S.; Canton, E.; Rodrigues, A.G.; Pina-Vaz, C. Species distribution and in vitro antifungal susceptibility profiles of yeast isolates from invasive infections during a Portuguese multicenter survey. Eur. J. Clin. Microbiol. Infect. Dis. 2014, 33, 2241–2247. [Google Scholar] [CrossRef] [PubMed]
  35. Akins, R.A. An update on antifungal targets and mechanisms of resistance in Candida albicans. Med. Mycol. 2005, 43, 285–318. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  36. Xiao, L.; Madison, V.; Chau, A.S.; Loebenberg, D.; Palermo, R.E.; McNicholas, P.M. Three-dimensional models of wild-type and mutated forms of cytochrome P450 14alpha-sterol demethylases from Aspergillus fumigatus and Candida albicans provide insights into posaconazole binding. Antimicrob. Agents Chemother. 2004, 48, 568–574. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  37. Odds, F.C.; Brown, A.J.; Gow, N.A. Antifungal agents: Mechanisms of action. Trends Microbiol. 2003, 11, 272–279. [Google Scholar] [CrossRef]
  38. Meyer, E.; Schwab, F.; Gastmeier, P.; Ruden, H.; Heininger, A. Antifungal use in intensive care units. J. Antimicrob. Chemother. 2007, 60, 619–624. [Google Scholar] [CrossRef]
  39. Espinel-Ingroff, A.; Diekema, D.J.; Fothergill, A.; Johnson, E.; Pelaez, T.; Pfaller, M.A.; Rinaldi, M.G.; Canton, E.; Turnidge, J. Wild-type MIC distributions and epidemiological cutoff values for the triazoles and six Aspergillus spp. for the CLSI broth microdilution method (M38-A2 document). J. Clin. Microbiol. 2010, 48, 3251–3257. [Google Scholar] [CrossRef] [Green Version]
  40. Nagappan, V.; Deresinski, S. Reviews of anti-infective agents: Posaconazole: A broad-spectrum triazole antifungal agent. Clin. Infect. Dis. 2007, 45, 1610–1617. [Google Scholar] [CrossRef] [Green Version]
  41. Pfaller, M.A.; Messer, S.A.; Rhomberg, P.R.; Jones, R.N.; Castanheira, M. In vitro activities of isavuconazole and comparator antifungal agents tested against a global collection of opportunistic yeasts and molds. J. Clin. Microbiol. 2013, 51, 2608–2616. [Google Scholar] [CrossRef] [Green Version]
  42. Seyedmousavi, S.; Verweij, P.E.; Mouton, J.W. Isavuconazole, a broad-spectrum triazole for the treatment of systemic fungal diseases. Expert Rev. Anti-Infect. Ther. 2015, 13, 9–27. [Google Scholar] [CrossRef] [PubMed]
  43. Thompson, G.R., 3rd; Wiederhold, N.P.; Sutton, D.A.; Fothergill, A.; Patterson, T.F. In vitro activity of isavuconazole against Trichosporon, Rhodotorula, Geotrichum, Saccharomyces and Pichia species. J. Antimicrob. Chemother. 2009, 64, 79–83. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  44. Ullmann, A.J.; Aguado, J.M.; Arikan-Akdagli, S.; Denning, D.W.; Groll, A.H.; Lagrou, K.; Lass-Florl, C.; Lewis, R.E.; Munoz, P.; Verweij, P.E.; et al. Diagnosis and management of Aspergillus diseases: Executive summary of the 2017 ESCMID-ECMM-ERS guideline. Clin. Microbiol. Infect. 2018, 24, e1–e38. [Google Scholar] [CrossRef] [PubMed]
  45. Holmes, A.R.; Lin, Y.H.; Niimi, K.; Lamping, E.; Keniya, M.; Niimi, M.; Tanabe, K.; Monk, B.C.; Cannon, R.D. ABC transporter Cdr1p contributes more than Cdr2p does to fluconazole efflux in fluconazole-resistant Candida albicans clinical isolates. Antimicrob. Agents Chemother. 2008, 52, 3851–3862. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  46. Rogers, P.D.; Barker, K.S. Genome-wide expression profile analysis reveals coordinately regulated genes associated with stepwise acquisition of azole resistance in Candida albicans clinical isolates. Antimicrob. Agents Chemother. 2003, 47, 1220–1227. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  47. Rodriguez-Tudela, J.L.; Alcazar-Fuoli, L.; Cuesta, I.; Alastruey-Izquierdo, A.; Monzon, A.; Mellado, E.; Cuenca-Estrella, M. Clinical relevance of resistance to antifungals. Int. J. Antimicrob. Agents 2008, 32, S111–S113. [Google Scholar] [CrossRef]
  48. Clark, T.A.; Hajjeh, R.A. Recent trends in the epidemiology of invasive mycoses. Curr. Opin. Infect. Dis. 2002, 15, 569–574. [Google Scholar] [CrossRef]
  49. Ricardo, E.; Silva, A.P.; Goncalves, T.; Costa de Oliveira, S.; Granato, C.; Martins, J.; Rodrigues, A.G.; Pina-Vaz, C. Candida krusei reservoir in a neutropaenia unit: Molecular evidence of a foe? Clin. Microbiol. Infect. Off. Publ. Eur. Soc. Clin. Microbiol. Infect. Dis. 2011, 17, 259–263. [Google Scholar] [CrossRef] [Green Version]
  50. Cole, G.T.; Halawa, A.A.; Anaissie, E.J. The role of the gastrointestinal tract in hematogenous candidiasis: From the laboratory to the bedside. Clin. Infect. Dis. Off. Publ. Infect. Dis. Soc. Am. 1996, 22, S73–S88. [Google Scholar] [CrossRef] [Green Version]
  51. Sanglard, D.; Ischer, F.; Calabrese, D.; Majcherczyk, P.A.; Bille, J. The ATP binding cassette transporter gene CgCDR1 from Candida glabrata is involved in the resistance of clinical isolates to azole antifungal agents. Antimicrob. Agents Chemother. 1999, 43, 2753–2765. [Google Scholar] [CrossRef] [Green Version]
  52. Cowen, L.E. The evolution of fungal drug resistance: Modulating the trajectory from genotype to phenotype. Nat. Rev. Microbiol. 2008, 6, 187–198. [Google Scholar] [CrossRef] [PubMed]
  53. Sanglard, D.; Coste, A.; Ferrari, S. Antifungal drug resistance mechanisms in fungal pathogens from the perspective of transcriptional gene regulation. FEMS Yeast Res. 2009, 9, 1029–1050. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  54. Borst, P.; Evers, R.; Kool, M.; Wijnholds, J. A family of drug transporters: The multidrug resistance-associated proteins. J. Natl. Cancer Inst. 2000, 92, 1295–1302. [Google Scholar] [CrossRef]
  55. Morschhauser, J. Regulation of multidrug resistance in pathogenic fungi. Fungal Genet. Biol. 2010, 47, 94–106. [Google Scholar] [CrossRef] [PubMed]
  56. Cannon, R.D.; Lamping, E.; Holmes, A.R.; Niimi, K.; Baret, P.V.; Keniya, M.V.; Tanabe, K.; Niimi, M.; Goffeau, A.; Monk, B.C. Efflux-mediated antifungal drug resistance. Clin. Microbiol. Rev. 2009, 22, 291–321. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  57. Franz, R.; Kelly, S.L.; Lamb, D.C.; Kelly, D.E.; Ruhnke, M.; Morschhauser, J. Multiple molecular mechanisms contribute to a stepwise development of fluconazole resistance in clinical Candida albicans strains. Antimicrob. Agents Chemother. 1998, 42, 3065–3072. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  58. Perea, S.; Lopez-Ribot, J.L.; Kirkpatrick, W.R.; McAtee, R.K.; Santillan, R.A.; Martinez, M.; Calabrese, D.; Sanglard, D.; Patterson, T.F. Prevalence of molecular mechanisms of resistance to azole antifungal agents in Candida albicans strains displaying high-level fluconazole resistance isolated from human immunodeficiency virus-infected patients. Antimicrob. Agents Chemother. 2001, 45, 2676–2684. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  59. Ricardo, E.; Costa-de-Oliveira, S.; Dias, A.S.; Guerra, J.; Rodrigues, A.G.; Pina-Vaz, C. Ibuprofen reverts antifungal resistance on Candida albicans showing overexpression of CDR genes. FEMS Yeast Res. 2009, 9, 618–625. [Google Scholar] [CrossRef] [Green Version]
  60. Prasad, R.; Banerjee, A.; Khandelwal, N.K.; Dhamgaye, S. The ABCs of Candida albicans Multidrug Transporter Cdr1. Eukaryot. Cell 2015, 14, 1154–1164. [Google Scholar] [CrossRef] [Green Version]
  61. Coste, A.T.; Karababa, M.; Ischer, F.; Bille, J.; Sanglard, D. TAC1, transcriptional activator of CDR genes, is a new transcription factor involved in the regulation of Candida albicans ABC transporters CDR1 and CDR2. Eukaryot. Cell 2004, 3, 1639–1652. [Google Scholar] [CrossRef] [Green Version]
  62. Coste, A.; Turner, V.; Ischer, F.; Morschhauser, J.; Forche, A.; Selmecki, A.; Berman, J.; Bille, J.; Sanglard, D. A mutation in Tac1p, a transcription factor regulating CDR1 and CDR2, is coupled with loss of heterozygosity at chromosome 5 to mediate antifungal resistance in Candida albicans. Genetics 2006, 172, 2139–2156. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  63. Selmecki, A.; Forche, A.; Berman, J. Aneuploidy and isochromosome formation in drug-resistant Candida albicans. Science 2006, 313, 367–370. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  64. Legrand, M.; Lephart, P.; Forche, A.; Mueller, F.M.; Walsh, T.; Magee, P.T.; Magee, B.B. Homozygosity at the MTL locus in clinical strains of Candida albicans: Karyotypic rearrangements and tetraploid formation. Mol. Microbiol. 2004, 52, 1451–1462. [Google Scholar] [CrossRef] [PubMed]
  65. Hull, C.M.; Johnson, A.D. Identification of a mating type-like locus in the asexual pathogenic yeast Candida albicans. Science 1999, 285, 1271–1275. [Google Scholar] [CrossRef] [PubMed]
  66. Coste, A.; Selmecki, A.; Forche, A.; Diogo, D.; Bougnoux, M.E.; D’ENFERT, C.; Berman, J.; Sanglard, D. Genotypic evolution of azole resistance mechanisms in sequential Candida albicans isolates. Eukaryot. Cell 2007, 6, 1889–1904. [Google Scholar] [CrossRef] [Green Version]
  67. Rustad, T.R.; Stevens, D.A.; Pfaller, M.A.; White, T.C. Homozygosity at the Candida albicans MTL locus associated with azole resistance. Microbiology 2002, 148, 1061–1072. [Google Scholar] [CrossRef] [Green Version]
  68. Lockhart, S.R.; Pujol, C.; Daniels, K.J.; Miller, M.G.; Johnson, A.D.; Pfaller, M.A.; Soll, D.R. In Candida albicans, white-opaque switchers are homozygous for mating type. Genetics 2002, 162, 737–745. [Google Scholar]
  69. Pujol, C.; Messer, S.A.; Pfaller, M.; Soll, D.R. Drug resistance is not directly affected by mating type locus zygosity in Candida albicans. Antimicrob. Agents Chemother. 2003, 47, 1207–1212. [Google Scholar] [CrossRef] [Green Version]
  70. White, T.C.; Holleman, S.; Dy, F.; Mirels, L.F.; Stevens, D.A. Resistance mechanisms in clinical isolates of Candida albicans. Antimicrob. Agents Chemother. 2002, 46, 1704–1713. [Google Scholar] [CrossRef] [Green Version]
  71. Morschhauser, J.; Barker, K.S.; Liu, T.T.; Bla, B.W.J.; Homayouni, R.; Rogers, P.D. The transcription factor Mrr1p controls expression of the MDR1 efflux pump and mediates multidrug resistance in Candida albicans. PLoS Pathog. 2007, 3, e164. [Google Scholar] [CrossRef] [Green Version]
  72. Dunkel, N.; Blass, J.; Rogers, P.D.; Morschhauser, J. Mutations in the multi-drug resistance regulator MRR1, followed by loss of heterozygosity, are the main cause of MDR1 overexpression in fluconazole-resistant Candida albicans strains. Mol. Microbiol. 2008, 69, 827–840. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  73. Sanglard, D.; Ischer, F.; Koymans, L.; Bille, J. Amino acid substitutions in the cytochrome P-450 lanosterol 14alpha-demethylase (CYP51A1) from azole-resistant Candida albicans clinical isolates contribute to resistance to azole antifungal agents. Antimicrob. Agents Chemother. 1998, 42, 241–253. [Google Scholar] [PubMed]
  74. Sanglard, D.; Ischer, F.; Parkinson, T.; Falconer, D.; Bille, J. Candida albicans mutations in the ergosterol biosynthetic pathway and resistance to several antifungal agents. Antimicrob. Agents Chemother. 2003, 47, 2404–2412. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  75. Martinez, M.; Lopez-Ribot, J.L.; Kirkpatrick, W.R.; Bachmann, S.P.; Perea, S.; Ruesga, M.T.; Patterson, T.F. Heterogeneous mechanisms of azole resistance in Candida albicans clinical isolates from an HIV-infected patient on continuous fluconazole therapy for oropharyngeal candidosis. J. Antimicrob. Chemother. 2002, 49, 515–524. [Google Scholar] [CrossRef] [Green Version]
  76. Selmecki, A.; Gerami-Nejad, M.; Paulson, C.; Forche, A.; Berman, J. An isochromosome confers drug resistance in vivo by amplification of two genes, ERG11 and TAC1. Mol. Microbiol. 2008, 68, 624–641. [Google Scholar] [CrossRef]
  77. Silver, P.M.; Oliver, B.G.; White, T.C. Role of Candida albicans transcription factor Upc2p in drug resistance and sterol metabolism. Eukaryot. Cell 2004, 3, 1391–1397. [Google Scholar] [CrossRef] [Green Version]
  78. Dunkel, N.; Liu, T.T.; Barker, K.S.; Homayouni, R.; Morschhauser, J.; Rogers, P.D. A gain-of-function mutation in the transcription factor Upc2p causes upregulation of ergosterol biosynthesis genes and increased fluconazole resistance in a clinical Candida albicans isolate. Eukaryot. Cell 2008, 7, 1180–1190. [Google Scholar] [CrossRef] [Green Version]
  79. Heilmann, C.J.; Schneider, S.; Barker, K.S.; Rogers, P.D.; Morschhauser, J. An A643T mutation in the transcription factor Upc2p causes constitutive ERG11 upregulation and increased fluconazole resistance in Candida albicans. Antimicrob. Agents Chemother. 2010, 54, 353–359. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  80. Oliver, B.G.; Song, J.L.; Choiniere, J.H.; White, T.C. cis-Acting elements within the Candida albicans ERG11 promoter mediate the azole response through transcription factor Upc2p. Eukaryot. Cell 2007, 6, 2231–2239. [Google Scholar] [CrossRef] [Green Version]
  81. Hoot, S.J.; Smith, A.R.; Brown, R.P.; White, T.C. An A643V amino acid substitution in Upc2p contributes to azole resistance in well-characterized clinical isolates of Candida albicans. Antimicrob. Agents Chemother. 2011, 55, 940–942. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  82. Agarwal, A.K.; Rogers, P.D.; Baerson, S.R.; Jacob, M.R.; Barker, K.S.; Cleary, J.D.; Walker, L.A.; Nagle, D.G.; Clark, A.M. Genome-wide expression profiling of the response to polyene, pyrimidine, azole, and echinocandin antifungal agents in Saccharomyces cerevisiae. J. Biol. Chem. 2003, 278, 34998–35015. [Google Scholar] [CrossRef] [Green Version]
  83. Denning, D.W. Echinocandin antifungal drugs. Lancet 2003, 362, 1142–1151. [Google Scholar] [CrossRef]
  84. Perlin, D.S. Resistance to echinocandin-class antifungal drugs. Drug Resist. Updates 2007, 10, 121–130. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  85. Garcia-Effron, G.; Park, S.; Perlin, D.S. Correlating echinocandin MIC and kinetic inhibition of fks1 mutant glucan synthases for Candida albicans: Implications for interpretive breakpoints. Antimicrob. Agents Chemother. 2009, 53, 112–122. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  86. Katiyar, S.; Pfaller, M.; Edlind, T. Candida albicans and Candida glabrata clinical isolates exhibiting reduced echinocandin susceptibility. Antimicrob. Agents Chemother. 2006, 50, 2892–2894. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  87. Garcia-Effron, G.; Chua, D.J.; Tomada, J.R.; DiPersio, J.; Perlin, D.S.; Ghannoum, M.; Bonilla, H. Novel FKS mutations associated with echinocandin resistance in Candida species. Antimicrob. Agents Chemother. 2010, 54, 2225–2227. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  88. Pfaller, M.A.; Diekema, D.J.; Turnidge, J.D.; Castanheira, M.; Jones, R.N. Twenty Years of the SENTRY Antifungal Surveillance Program: Results for Candida Species from 1997–2016. Open Forum Infect. Dis. 2019, 6, S79–S94. [Google Scholar] [CrossRef] [Green Version]
  89. Kritikos, A.; Neofytos, D.; Khanna, N.; Schreiber, P.W.; Boggian, K.; Bille, J.; Schrenzel, J.; Muhlethaler, K.; Zbinden, R.; Bruderer, T.; et al. Accuracy of Sensititre YeastOne echinocandins epidemiological cut-off values for identification of FKS mutant Candida albicans and Candida glabrata: A ten year national survey of the Fungal Infection Network of Switzerland (FUNGINOS). Clin. Microbiol. Infect. 2018, 24, e1211–e1214. [Google Scholar] [CrossRef] [PubMed]
  90. Garcia-Effron, G.; Lee, S.; Park, S.; Cleary, J.D.; Perlin, D.S. Effect of Candida glabrata FKS1 and FKS2 mutations on echinocandin sensitivity and kinetics of 1,3-beta-D-glucan synthase: Implication for the existing susceptibility breakpoint. Antimicrob. Agents Chemother. 2009, 53, 3690–3699. [Google Scholar] [CrossRef] [Green Version]
  91. Wiederhold, N.P.; Najvar, L.K.; Bocanegra, R.; Molina, D.; Olivo, M.; Graybill, J.R. In vivo efficacy of anidulafungin and caspofungin against Candida glabrata and association with in vitro potency in the presence of sera. Antimicrob. Agents Chemother. 2007, 51, 1616–1620. [Google Scholar] [CrossRef] [Green Version]
  92. Paderu, P.; Garcia-Effron, G.; Balashov, S.; Delmas, G.; Park, S.; Perlin, D.S. Serum differentially alters the antifungal properties of echinocandin drugs. Antimicrob. Agents Chemother. 2007, 51, 2253–2256. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  93. Delarze, E.; Sanglard, D. Defining the frontiers between antifungal resistance, tolerance and the concept of persistence. Drug Resist. Updates 2015, 23, 12–19. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  94. Meletiadis, J.; Chanock, S.; Walsh, T.J. Human pharmacogenomic variations and their implications for antifungal efficacy. Clin. Microbiol. Rev. 2006, 19, 763–787. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  95. Cowen, L.E.; Steinbach, W.J. Stress, drugs, and evolution: The role of cellular signaling in fungal drug resistance. Eukaryot. Cell 2008, 7, 747–764. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  96. Jain, P.; Akula, I.; Edlind, T. Cyclic AMP signaling pathway modulates susceptibility of candida species and Saccharomyces cerevisiae to antifungal azoles and other sterol biosynthesis inhibitors. Antimicrob. Agents Chemother. 2003, 47, 3195–3201. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  97. Maidan, M.M.; De Rop, L.; Serneels, J.; Exler, S.; Rupp, S.; Tournu, H.; Thevelein, J.M.; Van Dijck, P. The G protein-coupled receptor Gpr1 and the Galpha protein Gpa2 act through the cAMP-protein kinase A pathway to induce morphogenesis in Candida albicans. Mol. Biol. Cell 2005, 16, 1971–1986. [Google Scholar] [CrossRef] [Green Version]
  98. Juvvadi, P.R.; Lee, S.C.; Heitman, J.; Steinbach, W.J. Calcineurin in fungal virulence and drug resistance: Prospects for harnessing targeted inhibition of calcineurin for an antifungal therapeutic approach. Virulence 2017, 8, 186–197. [Google Scholar] [CrossRef] [Green Version]
  99. Onyewu, C.; Wormley, F.L., Jr.; Perfect, J.R.; Heitman, J. The calcineurin target, Crz1, functions in azole tolerance but is not required for virulence of Candida albicans. Infect. Immun. 2004, 72, 7330–7333. [Google Scholar] [CrossRef] [Green Version]
  100. Sanglard, D.; Ischer, F.; Marchetti, O.; Entenza, J.; Bille, J. Calcineurin A of Candida albicans: Involvement in antifungal tolerance, cell morphogenesis and virulence. Mol. Microbiol. 2003, 48, 959–976. [Google Scholar] [CrossRef] [PubMed]
  101. Jia, Y.; Tang, R.J.; Wang, L.; Zhang, X.; Wang, Y.; Jia, X.M.; Jiang, Y.Y. Calcium-activated-calcineurin reduces the In vitro and In vivo sensitivity of fluconazole to Candida albicans via Rta2p. PLoS ONE 2012, 7, e48369. [Google Scholar] [CrossRef] [Green Version]
  102. Singh, S.D.; Robbins, N.; Zaas, A.K.; Schell, W.A.; Perfect, J.R.; Cowen, L.E. Hsp90 governs echinocandin resistance in the pathogenic yeast Candida albicans via calcineurin. PLoS Pathog. 2009, 5, e1000532. [Google Scholar] [CrossRef] [PubMed]
  103. Walker, L.A.; Munro, C.A.; de Bruijn, I.; Lenardon, M.D.; McKinnon, A.; Gow, N.A. Stimulation of chitin synthesis rescues Candida albicans from echinocandins. PLoS Pathog. 2008, 4, e1000040. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  104. Munro, C.A.; Selvaggini, S.; de Bruijn, I.; Walker, L.; Lenardon, M.D.; Gerssen, B.; Milne, S.; Brown, A.J.; Gow, N.A. The PKC, HOG and Ca2+ signalling pathways co-ordinately regulate chitin synthesis in Candida albicans. Mol. Microbiol. 2007, 63, 1399–1413. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  105. Lenardon, M.D.; Munro, C.A.; Gow, N.A. Chitin synthesis and fungal pathogenesis. Curr. Opin. Microbiol. 2010, 13, 416–423. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  106. Stevens, D.A.; Ichinomiya, M.; Koshi, Y.; Horiuchi, H. Escape of Candida from caspofungin inhibition at concentrations above the MIC (paradoxical effect) accomplished by increased cell wall chitin; evidence for beta-1,6-glucan synthesis inhibition by caspofungin. Antimicrob. Agents Chemother. 2006, 50, 3160–3161. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  107. Jacobsen, M.D.; Whyte, J.A.; Odds, F.C. Candida albicans and Candida dubliniensis respond differently to echinocandin antifungal agents in vitro. Antimicrob. Agents Chemother. 2007, 51, 1882–1884. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  108. Fleischhacker, M.; Radecke, C.; Schulz, B.; Ruhnke, M. Paradoxical growth effects of the echinocandins caspofungin and micafungin, but not of anidulafungin, on clinical isolates of Candida albicans and C. dubliniensis. Eur. J. Clin. Microbiol. Infect. Dis. 2008, 27, 127–131. [Google Scholar] [CrossRef] [PubMed]
  109. Walker, L.A.; Gow, N.A.; Munro, C.A. Elevated chitin content reduces the susceptibility of Candida species to caspofungin. Antimicrob. Agents Chemother. 2013, 57, 146–154. [Google Scholar] [CrossRef] [Green Version]
  110. Costa-de-Oliveira, S.; Silva, A.P.; Miranda, I.M.; Salvador, A.; Azevedo, M.M.; Munro, C.A.; Rodrigues, A.G.; Pina-Vaz, C. Determination of chitin content in fungal cell wall: An alternative flow cytometric method. Cytom. Part 2013, 83, 324–328. [Google Scholar] [CrossRef] [Green Version]
  111. Lee, K.K.; Maccallum, D.M.; Jacobsen, M.D.; Walker, L.A.; Odds, F.C.; Gow, N.A.; Munro, C.A. Elevated Cell Wall Chitin in Candida albicans Confers Echinocandin Resistance In Vivo. Antimicrob. Agents Chemother. 2012, 56, 208–217. [Google Scholar] [CrossRef] [Green Version]
  112. Shields, R.K.; Nguyen, M.H.; Du, C.; Press, E.; Cheng, S.; Clancy, C.J. Paradoxical effect of caspofungin against Candida bloodstream isolates is mediated by multiple pathways but eliminated in human serum. Antimicrob. Agents Chemother. 2011, 55, 2641–2647. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  113. Xie, J.L.; Qin, L.; Miao, Z.; Grys, B.T.; Diaz, J.C.; Ting, K.; Krieger, J.R.; Tong, J.; Tan, K.; Leach, M.D.; et al. The Candida albicans transcription factor Cas5 couples stress responses, drug resistance and cell cycle regulation. Nat. Commun. 2017, 8, 499. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  114. Yang, F.; Kravets, A.; Bethlendy, G.; Welle, S.; Rustchenko, E. Chromosome 5 monosomy of Candida albicans controls susceptibility to various toxic agents, including major antifungals. Antimicrob. Agents Chemother. 2013, 57, 5026–5036. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  115. Yang, F.; Zhang, L.; Wakabayashi, H.; Myers, J.; Jiang, Y.; Cao, Y.; Jimenez-Ortigosa, C.; Perlin, D.S.; Rustchenko, E. Tolerance to Caspofungin in Candida albicans Is Associated with at Least Three Distinctive Mechanisms That Govern Expression of FKS Genes and Cell Wall Remodeling. Antimicrob. Agents Chemother. 2017, 61. [Google Scholar] [CrossRef] [Green Version]
  116. Suwunnakorn, S.; Wakabayashi, H.; Rustchenko, E. Chromosome 5 of Human Pathogen Candida albicans Carries Multiple Genes for Negative Control of Caspofungin and Anidulafungin Susceptibility. Antimicrob. Agents Chemother. 2016, 60, 7457–7467. [Google Scholar] [CrossRef] [Green Version]
  117. Lewis, R.E.; Liao, G.; Young, K.; Douglas, C.; Kontoyiannis, D.P. Macrophage reporter cell assay for screening immunopharmacological activity of cell wall-active antifungals. Antimicrob. Agents Chemother. 2014, 58, 1738–1743. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  118. Mora-Montes, H.M.; Netea, M.G.; Ferwerda, G.; Lenardon, M.D.; Brown, G.D.; Mistry, A.R.; Kullberg, B.J.; O’Callaghan, C.A.; Sheth, C.C.; Odds, F.C.; et al. Recognition and blocking of innate immunity cells by Candida albicans chitin. Infect. Immun. 2011, 79, 1961–1970. [Google Scholar] [CrossRef] [Green Version]
  119. Marakalala, M.J.; Vautier, S.; Potrykus, J.; Walker, L.A.; Shepardson, K.M.; Hopke, A.; Mora-Montes, H.M.; Kerrigan, A.; Netea, M.G.; Murray, G.I.; et al. Differential adaptation of Candida albicans in vivo modulates immune recognition by dectin-1. PLoS Pathog. 2013, 9, e1003315. [Google Scholar] [CrossRef]
  120. Miranda, I.; Rocha, R.; Santos, M.C.; Mateus, D.D.; Moura, G.R.; Carreto, L.; Santos, M.A. A genetic code alteration is a phenotype diversity generator in the human pathogen Candida albicans. PLoS ONE 2007, 2, e996. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  121. Silva, R.M.; Paredes, J.A.; Moura, G.R.; Manadas, B.; Lima-Costa, T.; Rocha, R.; Miranda, I.; Gomes, A.C.; Koerkamp, M.J.; Perrot, M.; et al. Critical roles for a genetic code alteration in the evolution of the genus Candida. EMBO J. 2007, 26, 4555–4565. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  122. D’ENFERT, C. Hidden killers: Persistence of opportunistic fungal pathogens in the human host. Curr. Opin. Microbiol. 2009, 12, 358–364. [Google Scholar] [CrossRef]
  123. Silva-Dias, A.; Miranda, I.M.; Branco, J.; Monteiro-Soares, M.; Pina-Vaz, C.; Rodrigues, A.G. Adhesion, biofilm formation, cell surface hydrophobicity, and antifungal planktonic susceptibility: Relationship among Candida spp. Front. Microbiol. 2015, 6, 205. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  124. Araujo, D.; Henriques, M.; Silva, S. Portrait of Candida Species Biofilm Regulatory Network Genes. Trends Microbiol. 2017, 25, 62–75. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  125. Nobile, C.J.; Fox, E.P.; Nett, J.E.; Sorrells, T.R.; Mitrovich, Q.M.; Hernday, A.D.; Tuch, B.B.; Andes, D.R.; Johnson, A.D. A recently evolved transcriptional network controls biofilm development in Candida albicans. Cell 2012, 148, 126–138. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  126. Andes, D.; Nett, J.; Oschel, P.; Albrecht, R.; Marchillo, K.; Pitula, A. Development and characterization of an in vivo central venous catheter Candida albicans biofilm model. Infect. Immun. 2004, 72, 6023–6031. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  127. Mukherjee, P.K.; Chandra, J.; Kuhn, D.M.; Ghannoum, M.A. Mechanism of fluconazole resistance in Candida albicans biofilms: Phase-specific role of efflux pumps and membrane sterols. Infect. Immun. 2003, 71, 4333–4340. [Google Scholar] [CrossRef] [Green Version]
  128. Ramage, G.; Rajendran, R.; Sherry, L.; Williams, C. Fungal biofilm resistance. Int. J. Microbiol. 2012, 2012, 528521. [Google Scholar] [CrossRef]
  129. Ramage, G.; Bachmann, S.; Patterson, T.F.; Wickes, B.L.; Lopez-Ribot, J.L. Investigation of multidrug efflux pumps in relation to fluconazole resistance in Candida albicans biofilms. J. Antimicrob. Chemother. 2002, 49, 973–980. [Google Scholar] [CrossRef] [Green Version]
  130. Wang, J.F.; Xue, Y.; Zhu, X.B.; Fan, H. Efficacy and safety of echinocandins versus triazoles for the prophylaxis and treatment of fungal infections: A Meta-analysis of RCTs. Eur. J. Clin. Microbiol. Infect. Dis. 2015, 34, 651–659. [Google Scholar] [CrossRef] [Green Version]
  131. Katragkou, A.; Roilides, E.; Walsh, T.J. Role of Echinocandins in Fungal Biofilm-Related Disease: Vascular Catheter-Related Infections, Immunomodulation, and Mucosal Surfaces. Clin. Infect. Dis. 2015, 61, S622–S629. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  132. Seidler, M.; Salvenmoser, S.; Muller, F.M. Liposomal amphotericin B eradicates Candida albicans biofilm in a continuous catheter flow model. FEMS Yeast Res. 2010, 10, 492–495. [Google Scholar] [CrossRef] [PubMed]
  133. Larkin, E.L.; Dharmaiah, S.; Ghannoum, M.A. Biofilms and beyond: Expanding echinocandin utility. J. Antimicrob. Chemother. 2018, 73, i73–i81. [Google Scholar] [CrossRef] [PubMed]
  134. Brun, S.; Aubry, C.; Lima, O.; Filmon, R.; Berges, T.; Chabasse, D.; Bouchara, J.P. Relationships between respiration and susceptibility to azole antifungals in Candida glabrata. Antimicrob. Agents Chemother. 2003, 47, 847–853. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  135. Chamilos, G.; Lewis, R.E.; Kontoyiannis, D.P. Inhibition of Candida parapsilosis mitochondrial respiratory pathways enhances susceptibility to caspofungin. Antimicrob. Agents Chemother. 2006, 50, 744–747. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  136. Costa-de-Oliveira, S.; Sampaio-Marques, B.; Barbosa, M.; Ricardo, E.; Pina-Vaz, C.; Ludovico, P.; Rodrigues, A.G. An alternative respiratory pathway on Candida krusei: Implications on susceptibility profile and oxidative stress. FEMS Yeast Res. 2012, 12, 423–429. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  137. Yan, L.; Li, M.; Cao, Y.; Gao, P.; Wang, Y.; Jiang, Y. The alternative oxidase of Candida albicans causes reduced fluconazole susceptibility. J. Antimicrob. Chemother. 2009, 64, 764–773. [Google Scholar] [CrossRef] [Green Version]
  138. Cannon, R.D.; Lamping, E.; Holmes, A.R.; Niimi, K.; Tanabe, K.; Niimi, M.; Monk, B.C. Candida albicans drug resistance another way to cope with stress. Microbiology 2007, 153, 3211–3217. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  139. Duvenage, L.; Munro, C.A.; Gourlay, C.W. The potential of respiration inhibition as a new approach to combat human fungal pathogens. Curr. Genet. 2019, 65, 1–7. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  140. Duvenage, L.; Walker, L.A.; Bojarczuk, A.; Johnston, S.A.; MacCallum, D.M.; Munro, C.A.; Gourlay, C.W. Inhibition of Classical and Alternative Modes of Respiration in Candida albicans Leads to Cell Wall Remodeling and Increased Macrophage Recognition. MBio 2019, 10. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  141. Eggimann, P.; Garbino, J.; Pittet, D. Management of Candida species infections in critically ill patients. Lancet Infect. Dis. 2003, 3, 772–785. [Google Scholar] [CrossRef]
  142. Hachem, R.; Hanna, H.; Kontoyiannis, D.; Jiang, Y.; Raad, I. The changing epidemiology of invasive candidiasis: Candida glabrata and Candida krusei as the leading causes of candidemia in hematologic malignancy. Cancer 2008, 112, 2493–2499. [Google Scholar] [CrossRef] [PubMed]
  143. Goldani, L.Z.; Craven, D.E.; Sugar, A.M. Central venous catheter infection with Rhodotorula minuta in a patient with AIDS taking suppressive doses of fluconazole. J. Med. Vet. Mycol. Bi-Mon. Publ. Int. Soc. Hum. Anim. Mycol. 1995, 33, 267–270. [Google Scholar]
  144. Cutler, J.E. Putative virulence factors of Candida albicans. Annu. Rev. Microbiol. 1991, 45, 187–218. [Google Scholar] [CrossRef] [PubMed]
  145. Vogel, M.; Hartmann, T.; Koberle, M.; Treiber, M.; Autenrieth, I.B.; Schumacher, U.K. Rifampicin induces MDR1 expression in Candida albicans. J. Antimicrob. Chemother. 2008, 61, 541–547. [Google Scholar] [CrossRef] [PubMed]
  146. Stergiopoulou, T.; Meletiadis, J.; Sein, T.; Papaioannidou, P.; Tsiouris, I.; Roilides, E.; Walsh, T.J. Comparative pharmacodynamic interaction analysis between ciprofloxacin, moxifloxacin and levofloxacin and antifungal agents against Candida albicans and Aspergillus fumigatus. J. Antimicrob. Chemother. 2009, 63, 343–348. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  147. Nogueira, F.; Sharghi, S.; Kuchler, K.; Lion, T. Pathogenetic Impact of Bacterial-Fungal Interactions. Microorganisms 2019, 7, 459. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  148. Maki, N.; Moitra, K.; Silver, C.; Ghosh, P.; Chattopadhyay, A.; Dey, S. Modulator-induced interference in functional cross talk between the substrate and the ATP sites of human P-glycoprotein. Biochemistry 2006, 45, 2739–2751. [Google Scholar] [CrossRef] [PubMed]
  149. Modok, S.; Mellor, H.R.; Callaghan, R. Modulation of multidrug resistance efflux pump activity to overcome chemoresistance in cancer. Curr. Opin. Pharmacol. 2006, 6, 350–354. [Google Scholar] [CrossRef]
  150. Tsujimura, S.; Saito, K.; Nawata, M.; Nakayamada, S.; Tanaka, Y. Overcoming drug resistance induced by P-glycoprotein on lymphocytes in patients with refractory rheumatoid arthritis. Ann. Rheum. Dis. 2008, 67, 380–388. [Google Scholar] [CrossRef]
  151. Nim, S.; Rawal, M.K.; Prasad, R. FK520 interacts with the discrete intrahelical amino acids of multidrug transporter Cdr1 protein and acts as antagonist to selectively chemosensitize azole-resistant clinical isolates of Candida albicans. FEMS Yeast Res. 2014, 14, 624–632. [Google Scholar] [CrossRef] [Green Version]
  152. Tanabe, K.; Bonus, M.; Tomiyama, S.; Miyoshi, K.; Nagi, M.; Niimi, K.; Chindamporn, A.; Gohlke, H.; Schmitt, L.; Cannon, R.D.; et al. FK506 Resistance of Saccharomyces cerevisiae Pdr5 and Candida albicans Cdr1 Involves Mutations in the Transmembrane Domains and Extracellular Loops. Antimicrob. Agents Chemother. 2019, 63. [Google Scholar] [CrossRef] [Green Version]
  153. Twentyman, P.R. Cyclosporins as drug resistance modifiers. Biochem. Pharmacol. 1992, 43, 109–117. [Google Scholar] [CrossRef]
  154. Kawamura, A.; Su, M.S. Interaction of FKBP12-FK506 with calcineurin A at the B subunit-binding domain. J. Biol. Chem. 1995, 270, 15463–15466. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  155. Sun, S.; Li, Y.; Guo, Q.; Shi, C.; Yu, J.; Ma, L. In vitro interactions between tacrolimus and azoles against Candida albicans determined by different methods. Antimicrob. Agents Chemother. 2008, 52, 409–417. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  156. Marchetti, O.; Entenza, J.M.; Sanglard, D.; Bille, J.; Glauser, M.P.; Moreillon, P. Fluconazole plus cyclosporine: A fungicidal combination effective against experimental endocarditis due to Candida albicans. Antimicrob. Agents Chemother. 2000, 44, 2932–2938. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  157. Steinbach, W.J.; Reedy, J.L.; Cramer, R.A., Jr.; Perfect, J.R.; Heitman, J. Harnessing calcineurin as a novel anti-infective agent against invasive fungal infections. Nat. Rev. Microbiol. 2007, 5, 418–430. [Google Scholar] [CrossRef] [PubMed]
  158. Byrne, S.T.; Denkin, S.M.; Zhang, Y. Aspirin and ibuprofen enhance pyrazinamide treatment of murine tuberculosis. J. Antimicrob. Chemother. 2007, 59, 313–316. [Google Scholar] [CrossRef]
  159. Arai, R.; Sugita, T.; Nishikawa, A. Reassessment of the in vitro synergistic effect of fluconazole with the non-steroidal anti-inflammatory agent ibuprofen against Candida albicans. Mycoses 2005, 48, 38–41. [Google Scholar] [CrossRef]
  160. Pina-Vaz, C.; Rodrigues, A.G.; Costa-de-Oliveira, S.; Ricardo, E.; Mardh, P.A. Potent synergic effect between ibuprofen and azoles on Candida resulting from blockade of efflux pumps as determined by FUN-1 staining and flow cytometry. J. Antimicrob. Chemother. 2005, 56, 678–685. [Google Scholar] [CrossRef] [Green Version]
  161. Venturini, T.P.; Rossato, L.; Spader, T.B.; Tronco-Alves, G.R.; Azevedo, M.I.; Weiler, C.B.; Santurio, J.M.; Alves, S.H. In vitro synergisms obtained by amphotericin B and voriconazole associated with non-antifungal agents against Fusarium spp. Diagn. Microbiol. Infect. Dis. 2011, 71, 126–130. [Google Scholar] [CrossRef] [Green Version]
  162. Costa-de-Oliveira, S.; Miranda, I.M.; Silva-Dias, A.; Silva, A.P.; Rodrigues, A.G.; Pina-Vaz, C. Ibuprofen potentiates the in vivo antifungal activity of fluconazole against Candida albicans murine infection. Antimicrob. Agents Chemother. 2015, 59, 4289–4292. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  163. Shao, J.; Zhang, M.; Wang, T.; Li, Y.; Wang, C. The roles of CDR1, CDR2, and MDR1 in kaempferol-induced suppression with fluconazole-resistant Candida albicans. Pharm. Biol. 2016, 54, 984–992. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  164. Monk, B.C.; Niimi, K.; Lin, S.; Knight, A.; Kardos, T.B.; Cannon, R.D.; Parshot, R.; King, A.; Lun, D.; Harding, D.R. Surface-active fungicidal D-peptide inhibitors of the plasma membrane proton pump that block azole resistance. Antimicrob. Agents Chemother. 2005, 49, 57–70. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  165. Dubikovskaya, E.A.; Thorne, S.H.; Pillow, T.H.; Contag, C.H.; Wender, P.A. Overcoming multidrug resistance of small-molecule therapeutics through conjugation with releasable octaarginine transporters. Proc. Natl. Acad. Sci. USA 2008, 105, 12128–12133. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  166. Ganesan, L.T.; Manavathu, E.K.; Cutright, J.L.; Alangaden, G.J.; Chandrasekar, P.H. In-vitro activity of nikkomycin Z alone and in combination with polyenes, triazoles or echinocandins against Aspergillus fumigatus. Clin. Microbiol. Infect. 2004, 10, 961–966. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  167. Stevens, D.A. Drug interaction studies of a glucan synthase inhibitor (LY 303366) and a chitin synthase inhibitor (Nikkomycin Z) for inhibition and killing of fungal pathogens. Antimicrob. Agents Chemother. 2000, 44, 2547–2548. [Google Scholar] [CrossRef] [Green Version]
  168. Verwer, P.E.; van Duijn, M.L.; Tavakol, M.; Bakker-Woudenberg, I.A.; van de Sande, W.W. Reshuffling of Aspergillus fumigatus cell wall components chitin and beta-glucan under the influence of caspofungin or nikkomycin Z alone or in combination. Antimicrob. Agents Chemother. 2012, 56, 1595–1598. [Google Scholar] [CrossRef] [Green Version]
  169. Fernandes, C.; Anjos, J.; Walker, L.A.; Silva, B.M.; Cortes, L.; Mota, M.; Munro, C.A.; Gow, N.A.; Goncalves, T. Modulation of Alternaria infectoria cell wall chitin and glucan synthesis by cell wall synthase inhibitors. Antimicrob. Agents Chemother. 2014, 58, 2894–2904. [Google Scholar] [CrossRef] [Green Version]
  170. Kovacs, R.; Nagy, F.; Toth, Z.; Bozo, A.; Balazs, B.; Majoros, L. Synergistic effect of nikkomycin Z with caspofungin and micafungin against Candida albicans and Candida parapsilosis biofilms. Lett. Appl. Microbiol. 2019, 69, 271–278. [Google Scholar] [CrossRef] [Green Version]
  171. Cheung, Y.Y.; Hui, M. Effects of Echinocandins in Combination with Nikkomycin Z against Invasive Candida albicans Bloodstream Isolates and the fks Mutants. Antimicrob. Agents Chemother. 2017, 61. [Google Scholar] [CrossRef] [Green Version]
Figure 1. Risk factors that contribute to clinical resistance. Information was collected from the following references: [5,13,47,54,76,95,100,103,122,126,141,142,143,144,145,146].
Figure 1. Risk factors that contribute to clinical resistance. Information was collected from the following references: [5,13,47,54,76,95,100,103,122,126,141,142,143,144,145,146].
Microorganisms 08 00154 g001
Table 1. Spectrum of activity and mechanisms of action and resistance of the major antifungal agents enrolled in the treatment of invasive candidiasis.
Table 1. Spectrum of activity and mechanisms of action and resistance of the major antifungal agents enrolled in the treatment of invasive candidiasis.
Antifungal ClassAntifungal DrugSpectrum of ActivityMechanism(s) of ActionMechanism(s) of Resistance
PolyenesAmphotericin BFungicidalPolyene molecules links to ergosterol in the fungal membrane by inserting into the lipid bilayers, creating pores that disrupt plasma membrane; oxidative damage.Mutations in the ERG3 gene affect ergosterol biosynthesis and content in the fungal membrane is responsible for a decrease access to the drug target; susceptibility to oxidative damage by increasing catalase activity.
Pyrimidine analogues5-FlucytosineFungicidalInhibition of cellular function and division by incorporating toxic fluorinated pyrimidine antimetabolites into DNA and RNA.Mutations in the enzyme uracil phosphoribosyltransferase (Fur1p), decreasing the formation of toxic antimetabolites.
AzolesFluconazole
Voriconazole
Posaconazole
FungistaticInhibition of the fungal cytochrome P450 14α-lanosterol demethylase and accumulation of toxic methylated intermediates, with resultant disruption of fungal cell membrane function and growth inhibition.Overexpression of cell membrane efflux pumps, decreasing drug concentration (upregulation or overexpression CDR and MDR genes); alteration of the target enzyme, decreasing affinity to the binding site (point mutation in ERG11 gene); upregulation of the target enzyme (overexpression of ERG11 gene).
EchinocandinsCaspofungin
Anidulafungin
Micafungin
FungicidalInhibition of β-(1,3) glucan synthase, decreasing the production of β-(1,3) glucan, which represents one of the major components of the fungal cell wall.Point mutations in FKS1 and FKS2 genes.

Share and Cite

MDPI and ACS Style

Costa-de-Oliveira, S.; Rodrigues, A.G. Candida albicans Antifungal Resistance and Tolerance in Bloodstream Infections: The Triad Yeast-Host-Antifungal. Microorganisms 2020, 8, 154. https://doi.org/10.3390/microorganisms8020154

AMA Style

Costa-de-Oliveira S, Rodrigues AG. Candida albicans Antifungal Resistance and Tolerance in Bloodstream Infections: The Triad Yeast-Host-Antifungal. Microorganisms. 2020; 8(2):154. https://doi.org/10.3390/microorganisms8020154

Chicago/Turabian Style

Costa-de-Oliveira, Sofia, and Acácio G. Rodrigues. 2020. "Candida albicans Antifungal Resistance and Tolerance in Bloodstream Infections: The Triad Yeast-Host-Antifungal" Microorganisms 8, no. 2: 154. https://doi.org/10.3390/microorganisms8020154

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