Next Article in Journal
Inventory of the Secondary Metabolite Biosynthetic Potential of Members within the Terminal Clade of the Fusarium solani Species Complex
Previous Article in Journal
The Extracellular Lipopeptides and Volatile Organic Compounds of Bacillus subtilis DHA41 Display Broad-Spectrum Antifungal Activity against Soil-Borne Phytopathogenic Fungi
Previous Article in Special Issue
Iron Starvation Induces Ferricrocin Production and the Reductive Iron Acquisition System in the Chromoblastomycosis Agent Cladophialophora carrionii
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Candida parapsilosis sensu stricto Antifungal Resistance Mechanisms and Associated Epidemiology

1
Department of Translational Research on New Technologies in Medicine and Surgery, University of Pisa, Via San Zeno, 37, 56127 Pisa, Italy
2
Department of Biology, University of Pisa, Via San Zeno, 37, 56127 Pisa, Italy
*
Author to whom correspondence should be addressed.
J. Fungi 2023, 9(8), 798; https://doi.org/10.3390/jof9080798
Submission received: 5 June 2023 / Revised: 18 July 2023 / Accepted: 25 July 2023 / Published: 28 July 2023
(This article belongs to the Special Issue Molecular Diagnosis, Genetics and Evolution of Human Pathogenic Fungi)

Abstract

:
Fungal diseases cause millions of deaths per year worldwide. Antifungal resistance has become a matter of great concern in public health. In recent years rates of non-albicans species have risen dramatically. Candida parapsilosis is now reported to be the second most frequent species causing candidemia in several countries in Europe, Latin America, South Africa and Asia. Rates of acquired azole resistance are reaching a worrisome threshold from multiple reports as in vitro susceptibility testing is now starting also to explore tolerance and heteroresistance to antifungal compounds. With this review, the authors seek to evaluate known antifungal resistance mechanisms and their worldwide distribution in Candida species infections with a specific focus on C. parapsilosis.

1. Introduction

Candida spp. infections have dramatically increased in the last twenty years [1]. Non-albicans species represent a rising concern in the hospital epidemiology of candidemia. Several reports define Candida parapsilosis sensu stricto as the second most frequent isolate from bloodstream infections, especially in Italy [2], Turkey [3,4] and Latin America—where in some reports, it was the first Candida species isolated from blood cultures [5,6]—Greece [7], South Africa [8] and Asia [9,10], also with worrisome azole-resistance rates.
Candida parapsilosis sensu stricto is a member of the commensal skin flora. Its role as a human opportunistic pathogen is seen mainly in immunocompromised subjects, low-weight at-birth newborns, onco-hematologic individuals and patients admitted to intensive care and burn units [11,12]. Invasive medical devices such as central lines and or other prostheses represent the main substrate of colonization and deep seeding due to its innate ability to form biofilm on organic or inorganic surfaces [13,14]. Plus, thanks to these intrinsic factors, C. parapsilosis sensu stricto is able to colonize inanimate materials and survive within the environment allowing for in-hospital spreading and patient-to-patient transmission via health workers’ hands as for multi-drug resistant bacteria [15].
Antimicrobial resistance is an issue in healthcare-associated invasive fungal infections [16,17] and despite the availability of new drugs [18] and the continuous research on alternative compounds [19,20,21,22], the increase in rates of antifungal resistance in most fungal infections and C. parapsilosis, in particular, is narrowing therapeutic options [23,24].
The aim of this review was to evaluate and depict a global picture of known antifungal resistance mechanisms in Candida species infections with a specific focus on C. parapsilosis sensu stricto and its worldwide epidemiology.

2. Antifungal Drugs and Associated Resistance Mechanisms in Candida Species in Comparison with Candida parapsilosis

2.1. Introduction to Antimicrobial Resistance in Medical Mycology

Antimicrobial resistance is the ability of a microorganism with no preexisting intrinsic resistance mechanism against anti-infective compounds to survive and even thrive in the presence of such antimicrobial drugs, delivered at recommended concentrations [16,17,25,26,27,28,29]. Antimicrobial resistance can be distinguished into intrinsic and acquired, independently from the pathogen, as it has been found in both fungi and bacteria. Intrinsic resistance is based on conformational aspects of the microorganism that are constitutionally present, such as molecular structures, enzymes and cellular components, targeted by selected drugs, that make it naturally non-susceptible to the selected agent. The European Committee on Antimicrobial Susceptibility Testing (EUCAST) defined it also as “Expected resistant phenotype” [30]. Intrinsic resistance is genetically determined in all the cells belonging to a single species and, therefore, present within the microorganism prior to exposure to the selected drug. In other terms, intrinsic resistance is inferred by: (i) the absence or (ii) the different conformation of the target of a specific drug, making it not feasible for the antimicrobial agent to exert its activity [31,32,33]. Acquired resistance, on the contrary, represents the development of a resistant phenotype of a specific microorganism, which was not known to harbor any intrinsic resistance to the selected drug, and it is associated with prolonged drug exposure [16,18,26,34,35].
Parallel to these definitions, the authors have also identified two distinct scenarios for antimicrobial resistance: (i) microbiologic resistance, known as the ability of the microorganism to grow when exposed to adequate concentrations of an anti-infective drug normally active on wild-type strains and (ii) clinical resistance where the antimicrobial compound is unable to eradicate the in vivo infection occurring in a patient despite the demonstrated in vitro susceptibility of the infectious strain to the selected and currently administered antimicrobial therapy [16,17,26,28,36,37].
Therefore, when referring to antimicrobial resistance as a whole, it can be described as the lack of microbial growth inhibition usually achieved by the effective antimicrobial compound administered at regular dosage reaching adequate in vivo concentrations. This lack of growth inhibition can be directly observed in in vitro testing with Minimum Inhibitory Concentrations (MICs) for the selected drug falling outside the susceptible interval range. It can also be drawn by clinical worsening of the patient with persistent isolation of the pathogen from the same clinical samples despite the administration of the anti-infective drug already proven to be effective at the in vitro susceptibility tests [16,17,26,36,37]. Finally, the two main driving factors for the development of both clinical resistance and persistent isolation of the pathogen from patients’ clinical samples are (i) the ability of Candida spp. to form biofilm and (ii) reduced drug concentrations achieved at the infected site.
When evaluating antimicrobial resistance in fungi, especially for yeast infections of the Candida genus, the Clinical and Laboratory Standards Institute (CLSI) and EUCAST provided, over the years, two broth microdilution standardized methods of in vitro susceptibility testing that has been accepted by the scientific community as the two reference techniques to evaluate antifungal susceptibility, allowing for detection of resistant phenotypes for several Candida species [26,37,38,39].
Various genetic and molecular mechanisms underlying antifungal resistance have been extensively studied in Candida albicans, as this is the most frequently isolated Candida species implied in human infections [17,40]. Gene editing and CRISPR-Cas9-based techniques have been the primary tool to confirm the role and the effects of newly found mutations in Candida species [41]. However, the complexity of antifungal resistance resides in the co-existence within a single resistant strain of several mechanisms, each of which independently contributes to the non-susceptible phenotype. Therefore, acknowledging the main driving mutation or resistance mechanisms might be hard to accomplish [16,17,36]. Despite advances in research, several antifungal resistance mechanisms in different species need to be further evaluated in consideration of the fact that in recent years non-albicans species have become a real matter of concern for public health [16,17,26,27].

2.2. Azole Resistance

Azole drugs are major antifungal compounds that have been extensively used in clinical practice. Their mechanism of action is to bind and inhibit the lanosterol 14-alpha-demethylase [42,43]. This enzyme is present within several fungi and its activity is associated with ergosterol synthesis, a fungal cell membrane component [44,45]. Azole compounds, especially fluconazole, used to be the drug of choice prior to the advent of echinocandins and the first report of suggestion of echinocandin as the first line of therapy in candidemia patients is stated in the ESCMID 2012 guidelines [46]. The same assumptions and recommendations were adopted by the Clinical Practice Guidelines for the Management of candidiasis by the Infectious Diseases Society of America in 2016 [47]. Generally speaking, Candida species have developed three distinct molecular mechanisms through which they exert azole resistance. The most frequently encountered in clinical practice is associated with the increased activity of efflux pumps [48,49,50], which are fungal membrane proteins that fall in the molecular domain of ATP-binding cassette (ABC) transporters and major facilitator superfamilies (MFS) that can be found also in bacteria, plants and animals [51,52]. Their molecular activity aims at removing the drug from within the microorganism. The binding of the efflux pump to the azole compounds leads to the excretion of the antifungal molecule and, therefore, plays a pivotal role in the development of drug resistance [43,44,50,53]. The molecular ways through which overexpression of the efflux pump is achieved are represented by specific mutations in the transcription factors genes as gain-of-function mutations [50,54,55]. The type of efflux pump overexpressed may vary upon the Candida species exposed to azole compounds as several different transporters have been reported to be overexpressed according to different Candida species, despite similar underlying genetic mechanisms. For example, Cdr1p and Cdr2p are two transmembrane transporters that belong to the ABC-Transporter superfamily that have been reported to be overexpressed in azole-resistant Candida albicans strains [27,53]. Within the same superfamily, CgCdr1p, CgPdh1p and CgSnq2p are efflux pumps present in Candida glabrata; CkAbc1p and CkAbc2p in Candida krusei and Cdr1p in Candida auris. Among efflux-pumps that belong to the other superfamily (MFS-Transporter), Mdr1p has been found to be overexpressed in Candida albicans while CgQdr2p and CgFlr1p in Candida glabrata [27,49,53,56,57,58].
In the case of C. albicans the three major efflux pumps mentioned above detected in azole-resistant clinical isolates are encoded by corresponding genes CaCDR1, CaCDR2 and CaMDR1. The transcription factor CaTac1p regulates the expression of the first two efflux pump genes, while CaMrr1p, another transcription factor, controls the expression of the last one mentioned. In their corresponding genes, CaTAC1 and MRR1 authors found several gain-of-function mutations leading to increased levels of expression of the three related efflux pumps [59,60,61,62].
The same molecular mechanisms with different transcriptional genes involved have been described in C. glabrata, for which, gain-of-functions mutations in the gene CgPDR1 encoding for the transcription factor of ABC-T pumps CgCDR1, CgSNQ22 and CgPDH1 correlates with increased expression of corresponding efflux pumps resulting in azole resistance [63].
The second and third resistance mechanisms described for azole resistance are related to mutations in the genes directly encoding enzymes correlated to ergosterol synthesis or their transcription factors. The second mechanism is known as “target-mutation”, while the third leads to ergosterol overexpression. To this point, the most frequently reported gene hosting in vivo mutations in azole-resistant Candida albicans and Candida glabrata strains is ERG11 [56,63,64]. ERG11 encodes for a cytochrome P450 known as Erg11p [65], which has a sterol-14α-demethylase activity converting lanosterol into 4,4-Dimethylcholesta-8,14,24-trienol [65]. Azole compounds, in particular fluconazole, can bind and disrupt this fungal metabolic pathway. They act specifically on this enzyme, leading to the accumulation of a toxic metabolite, 14 alpha-methyl-ergosta-8,24(28)-dien-3 beta,6 alpha-diol. This toxic metabolite causes yeast cell death through its intracellular accumulation [42,65]. As stated above, Candida spp. have developed two distinct ways to overcome this detrimental effect: (i) overexpression of the target gene and (ii) gene target mutations, altering the binding site where usually the effect of the antifungal drug is elicited [27,44,66,67,68,69,70,71]. To the first point, authors have highlighted the presence of gain-of-function mutations in the transcription factor genes implied in the regulation of ERG11 expression. Such genes are known as UCP2 and NTD80 and have been found in both C. albicans and C. glabrata azole-resistant strains [72,73,74]. Gain-of-function mutations associated with an increased expression of the UPC2 gene are A643V, G648D, G648S and Y642F [73].
On the second mechanism, several point mutations have been linked to actual in vivo and in vitro azole resistance due to target mutations. For example in C. albicans, Xiang et al. [70] reported five different point mutations (Single Nucleotide Polymorphisms, SNP) in ERG11 correlated with azole resistance and investigated their structural position on a 3D model of the target enzyme. Their work demonstrated that amino acid substitutions caused by such mutations were all located near the substrate channel of the target enzyme (A114S, Y132F, Y132H, K143Q and K143R) or the active binding site (G472R) [70]. Other point mutations with the same effect on azole resistance have emerged from the study conducted by Silva and colleagues regarding C. glabrata (C108G, C423T and A1581G) and C. krusei (Y166S, G524R) [75]. Other genes connected to the ergosterol synthesis which may play a role in the development of resistance are ERG2, ERG3 and ERG6. Such genes encode for enzymes that convert the intermediate product of the ergosterol synthesis after exposure to azole and inhibition of the Erg11 cytochrome P450 generating toxic metabolites that compromise cell growth and vitality in both C. albicans and C. glabrata [76,77,78,79,80]. Mutations targeting these genes, causing gene disruption, led to the acquisition of an azole-resistant phenotype at in vitro studies since the conversion of the intermediate metabolite into the toxic one was blocked and yeast pathogen could withstand the azole-induced inhibition of the Lanosterol-14α-demethylase [78,79,81].
Gain-of-function and point mutations are not the only genetic mechanisms underlying an azole-resistant phenotype. Aneuploidy, altered mismatch repair, loss of heterozygosity, increase in number of copies of target genes and trisomy of selected chromosomes that incorporate ABC-Transporters, MFS or ERG11 genes, have all been demonstrated to elicit a resistant phenotype in a previously susceptible one [26,27,50,82,83,84,85].
It is still a matter of debate whether alterations in the azole-intake pathway could play a role in inducing azole resistance or not. The contribution of azole import to resistance has yet to be elucidated since the actual protein implied in the transmembrane transportation carrying the drug into the yeast cell has not been described so far. Despite this, the kinetics of azole accumulation into Candida spp. do not reflect those of passive diffusion and, therefore, the role and the presence of a possible carrier-protein have been postulated [27,50].
Another major antifungal drug within the azole family is isavuconazole. Such compound is relatively new and has shown promising in vitro activity against the most frequently encountered Candida species in clinical practice, such as C. albicans, C. glabrata, C. parapsilosis, C tropicalis, C. krusei, C. kefyr and C. lusitaniae [86,87]. Moreover, isavuconazole is currently referred to as an alternative treatment of invasive aspergillosis, a therapeutic option for mucormycosis and a potential oral-step down therapy in the treatment of candidemia, according to the ACTIVE trial results [88,89]. However, despite its relatively short period of clinical use, isavuconazole resistance has been described [89,90]. Main mechanisms of isavuconazole resistance were found in azole-resistant Candida species overexpressing the CDR genes, same results were not observed in those Candida spp. with increased expression of MDR1 gene [91]. In addition, also ERG11 and/or ERG3 mutations were found to be associated with the development of isavuconazole resistance [90,91]. Azole resistance in Candida parapsilosis sensu stricto has become a clinically relevant issue in the last decade [92], with the World Health Organization introducing this opportunistic pathogen among the high-priority group of yeast and fungal infections [23]. Major mutations linked to clinically demonstrated acquired resistance to azole compounds and associated affected molecules are shown in Table 1.
C. parapsilosis sensu stricto had no intrinsic resistance to azole drugs, therefore, reports and studies on azole-resistant hospital outbreaks are related to acquired resistance [11]. In an early study conducted by Silva and colleagues, three C. parapsilosis azole-susceptible strains were exposed to fluconazole, voriconazole or posaconazole in order to induce resistance, then gene expression of ERG11 and efflux pumps were analyzed [93]. The results demonstrated that the resistance mechanism could be associated with G583R and K873N amino acid substitution mutations in the transcription gene of a MFS MDR1, known as MRR1 for the fluconazole- and voriconazole-resistant strains. In the same study upregulation of UPC2 and NDT80 genes encoding for transcriptional factors increased the expression of ERG11 resulting in posaconazole resistance [93]. To this point, Arastehfar et al. identified two amino acid substitutions P45H and Q371H in the UPC2, leading to its overexpression, in fluconazole-resistant and voriconazole-susceptible-to-intermediate strains of C. parapsilosis [4].
In another study still conducted by Silva and colleagues [109], induction of azole resistance was obtained after exposing the yeast pathogens to several antifungals at different gradients. Their results pointed out that among azole compounds, fluconazole exposure took 15 days to induce resistance in previously susceptible isolates whether along the same period, no change in the susceptibility patterns of posaconazole was observed [109]. Plus, authors reported that induced fluconazole resistance would also affect susceptibility to voriconazole and vice versa; however, these two compounds showed no induced cross-resistance to posaconazole [109]. Surprisingly, after elimination of the azole pressure and subsequent cultures without further exposure to the previously mentioned antifungal compounds no substantial change was observed in the resistant susceptibility profiles. Plus, the same strains underwent treatment with known efflux pump inhibitors. Incredibly, these isolates did not revert the acquired resistant phenotype presenting high MIC values for fluconazole and voriconazole. This brought to the conclusion that the eventual underlying resistant mechanism could not be referred to efflux pumps since both the removal of the drug and the treatment with efflux pump inhibitors had no impact on the acquired resistance mechanism [109].
The residual susceptibility to posaconazole in fluconazole- and voriconazole-resistant C. parapsilosis strains can be explained partially because of the number of domains in the target site of the lanosterol-14α-demethylase that are bound by the different azoles. For instance, both fluconazole and voriconazole present only one binding site while posaconazolee has two of them. This is also the reason why overexpression of ERG11 is the resistance mechanism for posaconazole [93]. To this point, it is important to mention that also upregulation of MDR1 does not affect posaconazole susceptibility since this compound is a poor substrate of the previously mentioned efflux pump [53,110].
One more aspect of azole resistance should be further elucidated, as the two transcription factor genes UPC2 and NTD80 implied in the expression of enzymes correlated with ergosterol synthesis (ERG11 ERG2 ERG3 ERG4 ERG6 ERG25) reported in C. albicans were also found to be overexpressed in fluconazole-, voriconazole- and posaconazole-resistant C. parapsilosis [98]. In fact, their deletion restored complete susceptibility to all these antifungal compounds, however, in the same study, Branco et al. found that disruption of UPC2 had a more incisive reduction in MIC values of azole drugs than NTD80 [98].
Interesting findings on the underlying azole-resistance mechanism resulted from an experiment conducted by Souza and colleagues [95] on nine strains of fluconazole-resistant C. parapsilosis. Mutations in the ERG11 and in the efflux pump genes were explored. The results showed that all resistant strains harbored a missense mutation in the ERG11 gene generating the following amino acid substitution Y132F. This kind of mutation changed the protein structure leading to loss of binding activity with fluconazole [95], still other resistance mechanisms were found as overexpression of ERG11, CDR1 and less frequently MDR1 [95].
In a study on a Brazilian ICU cohort of COVID-19 patients with candidemia due to fluconazole-resistant C. parapsilosis, Daneshnia et al. found that only 35.1% of isolates showed the K143R mutation in the ERG11 gene. Interestingly, all fluconazole resistant isolates presented the L518F mutation in the TAC1 gene, which is a transcription factor of CDR1, which was demonstrated to be a causative mutation of acquired fluconazole- and voriconazole-resistance in the same study [100]. Berkow and colleagues identified two more mutations in the TAC1 transcription factor gene of C. parapsilosis (G650E and L978W) correlated with overexpression and upregulation of the target efflux pump Cdr1p with acquired fluconazole- and voriconazole-resistant phenotypes [97].
There is a well-established relationship between the type of efflux pump overexpressed and the associated resistance spectrum for Candida species other than C. parapsilosis. Overexpression of CDR efflux pump class, but not MDR, shows cross-resistance to all antifungal azole drugs, while the second class only affects mainly fluconazole [36]. However, regarding C. parapsilosis, Branco et al. [103] reported a case of cross-resistance between fluconazole and voriconazole directly correlated with a specific mutation G604R that induced overexpression of the MRR1 transcription factor gene resulting with the overexpression of the Mdr1 efflux pump [103].
Finally, a study conducted by Grossman et al. provided a great effort in elucidating the most frequent resistance mechanism for azole resistance in C. parapsilosis [111]. In their study, these authors evaluated and sequenced the genome of 30 fluconazole-resistant isolates obtained from blood-stream infections, demonstrating that 57% presented a SNP in the ERG11 gene resulting with the Y132F amino acid substitution previously mentioned [111]. Anyhow, also overexpression of MDR1 was registered, however, its frequency was less observed than the previous mutation. These authors reported that SNP correlated with MRR1 were more difficult to investigate and further research would have been required [111]. To this point, Branco and colleagues later on described two missense mutations into the MRR1 gene with amino acid substitution G583R and K873N imputable of determining the fluconazole- and voriconazole-resistant phenotype [94]. Still, data from a recent world-wide surveillance study conducted by Castanheira et al. analyzing multiple strains from different countries clearly determined that azole-resistance in C. parapsilosis was mainly driven by Y132F substitution in the ERG11 gene with a smaller role played by efflux pumps [112].
Last, it is important to mention that in a study from Arasthefar et al. [4] conducted during a clonal outbreak of candidemia due to azole-resistant C. parapsilosis, in addition to the Y132F substitution in ERG11, also the substitution K143R was described. Such amino acid change had been previously found also in C. albicans. Still, another important gene should be mentioned, as the ERG3 gene, which is also implied in the ergosterol synthesis, has been found to be target of point mutations with consequent development of azole resistance in C. parapsilosis. To this point, Branco et al. found a specific missense mutation (R135I) that led to loss of function of the enzyme in a posaconazole-resistant isolate [98].
Data regarding species-specific resistance mechanisms to isavuconazole in C. parapsilosis are scarce, however the previously mentioned mechanisms, described for other species, as overexpression of CDR1 gene and ERG11 target mutations, could be also found in C. parapsilosis. Anyhow, reports highlight that only a relatively small proportion of C. parapsilosis sensu stricto are non-wild-type and/or resistant to isavuconazole, as reported by Desnos-Ollivier et al. [87] and Marcos-Zambrano et al. [86], being respectively 0.8 and 1.1%.
Within the psilosis complex, in the context of azole-resistance, authors have reported a specific mutation in C. orthopsilosis known as A395T mutation in the CoERG11 gene. Such mutation is associated with a non-synonymous amino acid substitution Y132F and was proven to induce azole-resistance in previously susceptible C. orthopsilosis isolates [113]. Data regarding azole-resistance mechanisms in the two members of the psilosis group other than C. parapsilosis are anyhow lacking, however, it might be helpful to highlight that reports from different countries at in vitro tests showed very low rates of non-wild type MIC phenotypes for both C. orthopsilosis and C. metapsilosis for azole compounds [114,115].
An overview of major azole-resistance mechanisms for C. parapsilosis sensu stricto is depicted in Figure 1.
Five major mechanisms connected to azole-resistance in C. parapsilosis. SNPs in the TAC1 and MRR1 gene are associated with their overexpression and consequent upregulation of their targets, Cdr1p and Mdr1p respectively. SNPs in the UPC2 and NTD80 genes are associated with overexpression of ERG11, SNPs in the ERG11 gene alter the yeast target enzyme of azole compounds. SNPs in the ERG3 gene inducing loss of function mutations reduce the conversion of intermediate azole compounds in toxic metabolites with increase of Ergosta-7-enol (yellow oval) that could replace ergosterol (blue oval) in the fungal cell membrane without altering its molecular structure stability. (Image Created with BioRender.com).

2.3. Echinocandin Resistance

Echinocandins are a group of antifungal drugs that target specifically the β-(1,3) D-glucan synthase, which is encoded by two genes FKS1 and FKS2 that to a certain extent are redundant [116]. Glucans are polysaccharide components of the fungal cell wall, and their synthesis inhibition by echinocandins leads to cell death [116]. Precisely, the non-competitive molecular bond is established by the drug and a specific subunit of the fungal enzyme, known as Fks1p [117]. Their spectrum is broader than fluconazole and rates of fungal eradication were reported to be higher than fluconazole [118]; therefore, echinocandins are the recommended treatment in case of candidemia and invasive fungal infections due to Candida spp. as first-line empiric therapy [47].
Resistance to echinocandins is reported to be below 1% in C. albicans [40] clinical isolates and less than 10% in C. glabrata [119] and it has been linked to point mutations in the genes that encode the β-(1,3) D-glucan synthase causing an amino acidic substitution in the active-binding site of the target enzyme. A 645 serine to proline (S645P), phenylalanine (S645F) and tyrosine (S645Y) substitution in the Fks1p subunit triggers the development of echinocandin-resistance in C. albicans [28,120,121,122,123]. S645P substitution was reported to be the most prevalent among C. albicans [123]. Similar mutations have emerged in C. glabrata and C. krusei [124]. Caspofungin, micafungin and anidulafungin are all affected by these mutations, both in hetero or homozygosis since they are dominant and always associated with elevated MIC values [104]. In C. glabrata specifically, mutations in the FKS2 gene have been associated with a major impact on resistance than those present in the FKS1 gene [125]. Echinocandin-resistant phenotypes usually correlate with non-susceptibility to all antifungal drugs of the class with the exception of a single mutation in the FKS2 gene (Fks2p-S663F). This mutation was found in a C. glabrata strain where authors described a loss of drug activity for anidulafungin and caspofungin but not for micafungin [126]. Anyhow mutations obtained from clinical isolates in the FKS1-2 hot spot regions are known to affect the entire class of antifungal drugs [33,126].
Mutations in the hot spot region of FKS1 and FKS2 are not the only resistance mechanism described in echinocandin-resistant C. albicans and C. glabrata [127]. To this point, response to stress conditions may play a pivotal role, especially when fungal pathogens are exposed to echinocandins with alteration of the cell wall. In fact, when the integrity of the cell wall is disrupted, due to the β-(1,3) D-glucan synthase inhibition, studies have demonstrated that Rho1, a GTP-ase protein that represents the second subunit of the β-(1,3) D-glucan synthase gets activated. In the Fks1p subunit, which is the other subunit of the β-(1,3) D-glucan synthase, as previously mentioned, resides the catalytic activity of the enzyme, the actual site where the β-(1-3) D glucan is synthesized, that is targeted by echinocandins, while in the other subunit, Rho1, resides the regulatory activity of the enzyme itself. Authors speculate that activation of Rho1 may induce overexpression of the β-(1,3) D-glucan synthase while also triggering intracellular signaling of the protein kinase C (PKC) enabling fungal cell to activate a series of stress-responses, to compensate and restore the integrity of the cell wall through increase in chitin synthesis [128,129,130]. Also Ca2+/calcineurin, an intracellular stress-response pathway, contributes to the increase in chitin synthesis [131]. Last, but not least, Rho1 does not only activate PKC intracellular signaling pathway but it also provides upregulation of the FKS genes [127]. Despite being extensively studied in vitro, the clinical relevance of the above-mentioned molecular mechanisms and their role in treatment failure and antimicrobial resistance have yet to be demonstrated.
As for C. parapsilosis sensu stricto echinocandin-resistance has been less frequently reported than azole-resistance. Echinocandins exert a fungistatic effect on C. parapsilosis differently from the fungicidal activity displayed on other Candida species. This is due to a constitutional amino acid change in one of the hot spot regions of the Fks1p found to be naturally present in this kind of fungal species [104]. All mutations and relative effects on echinocandin susceptibility profile in C. parapsilosis are reported in Table 1. This constitutional substitution reported for C. parapsilosis accounted for its intrinsic reduced susceptibility to echinocandins associated with MIC values higher than other Candida species. Such intrinsic mutation was found in the hot spot region 1 of the subunit Fks1p and it was a proline to alanine substitution (P660A) [104]. This naturally occurring polymorphism has been also detected within the other species of the psilosis group like C. orthopsilosis and C. metapsilosis [104], still non-wild type phenotypes for such species for echinocandins are rarely seen in clinical practice [132,133,134], but the rarity of the isolation of such species does not allow to draw firm conclusions as more studies are needed to evaluate prevalence of non-wild type phenotypes for echinocandins. However other mutations previously described in echinocandin-resistant C. albicans or C. glabrata strains that were found within the hot spot region of the FKS1 and FKS2 genes were not present in echinocandin-resistant C. parapsilosis in the study of Martì-Carrizosa [105]. Indeed, they found that both mutations V595I and F1386S detected in echinocandin-resistant C. parapsilosis isolates were placed outside the hot spot regions [105]. These mutations were previously reported by Johnson and colleagues to be associated with acquired resistance to echinocandin in Saccharomyces cerevisiae [135]. Similar findings were reported by one study from a Brazilian outbreak of fluconazole-resistant and echinocandin-tolerant C. parapsilosis causing candidemia among COVID-19 patients. In this study, the authors linked the previously cited mutations to echinocandin-tolerant phenotype, and highlighted the presence of another specific mutation E1393G in the FKS1 gene, which was also linked to echinocandin-tolerance [100]. Further investigations correlated the presence of such mutation along with the above mentioned V595I, S745L and F1386S with the in vitro development of echinocandin-resistance [107]. Still, other studies reported also hot spot regions of the Fks1p to be the target of specific mutations affecting negatively echinocandin-susceptibility. For example, a recent study found that R658G mutation in the hot-spot region 1 of Fks1p was associated with a micafungin-resistant phenotype. This substitution was discovered in four micafungin-resistant C. parapsilosis strains isolated from blood cultures in 2020 [106]. A report from a Chinese study of a C. parapsilosis pan-echinocandin-resistant strain isolated from blood cultures revealed the presence of another mutation known as S656P still in the hot spot region 1 of Fks1p [108].
Surprisingly, another mechanism of echinocandin resistance was reported by Ryback et al. observing that a mutation G111R in ERG3 correlated with an increase in all echinocandins MIC. This was the first study to ever correlate an ERG3 loss of function with an acquired resistant phenotype to echinocandins, even though identification of the actual resistance mechanism is still matter of research [99].
Worthy of mention, especially in the case of C. parapsilosis senso strictu is rezafungin, which is a second generation echinocandin. Its pharmacokinetic/pharmacodynamic properties allow for a reduction in liver toxicity with a prolonged half-life, exerting the same inhibition observed for all other echinocandins on the β-1,3-D-glucan synthase [136]. By looking at the distribution of MIC reported within the psilosis group, C. parapsilosis sensu stricto showed higher MIC values (4 μg/mL) for this molecule, while C. metapsilosis was 0.5 μg/mL and C. orthopsilosis was 1 μg/mL [137]. As expected, all three psilosis species demonstrated MIC values higher than those reported for all other Candida species, due to their previously mentioned natural polymorphism [138]. Up to 2021, no resistance to rezafungin in C. parapsilosis sensu stricto was documented [139]; however, in 2022 Siopi et al. reported a case of pan-echinocandin C. parapsilosis sensu stricto—including also rezafungin—with an isolate harboring a new mutation in the HS region of the FKS1 gene (F652S) [140].

2.4. Polyene Resistance

Polyenes are a class of drugs that comprise Amphotericin B (AMB), Nystatin and Amphotericin A, with the first recognized as a major systemic antifungal drug and one of the first to be used in clinical practice [141]. AMB mechanism of action resides in the ability of the molecule to bind the ergosterol in the fungal membrane resulting in pore formation and loss of intracellular electrolytes causing lastly cell death [142]. AMB showed a broad spectrum of activity exerting a fungicidal effect on several Candida spp. and filamentous fungi [143,144]. Despite the long time since its introduction in clinical practice, rates of acquired resistance to AMB remained low and only rare cases have been reported [33,145]. Among yeast pathogens C. glabrata, C. krusei, Candida haemulonii, C. lusitaniae, C. auris and C. guillermondii are species that have been most frequently associated with AMB-resistance [126,145]. C. parapsilosis was listed among the AMB-susceptible fungal isolates [146].
Authors suggest that a reduction in the ergosterol composition of the fungal membrane associated with ERG2, ERG3 and ERG6 loss of function mutations might represent the underlying AMB-resistance mechanisms [33,68,126,145,147,148], since they reduce the amount of ergosterol present in the fungal membrane and, therefore, the target of AMB itself [149,150]. In addition, also up-regulation of ERG5, ERG6 and ERG25 has been associated with acquired AMB-resistance, since this modification led to the synthesis of a different sterol than ergosterol. Once inside the fungal cell membrane this new sterol intermediate displays a reduced binding activity to the antifungal drug ensuring anyhow structural stability to the yeast [78,145]. Still, upregulation of ERG genes and loss of function mutations are not the only mechanisms responsible for AMB-resistance. As observed for echinocandins, the stress-response may play an important role in the survival of the fungal cell. AMB-induced membrane alteration and consequent oxidative stress induces the acquisition of a resistant phenotype by increasing composition in chitin content of cell wall and by reducing fluidity of the membrane [151]. To this point, authors have highlighted another fungal resistance mechanism activated after exposure to AMB, which is adaptation and response to drug-induced oxidative stress. Fungal pathogens under AMB drug pressure might develop an increase in levels of oxidative stress-response proteins, such as catalase and heat shock protein 90 (HSP90), countering the negative effects of reactive oxygens species [130]. A latter mechanism of resistance reported by Healey et al. [152] in 2016, that is not exclusively related to AMB, focused on genes related to mismatch repair in C. glabrata. Disruption of the MSH2 gene increased mutation rates among other genes normally involved in resistance to azoles, echinocandins and also AMB leading to a multi-drug resistant phenotype [152].
Despite these findings, AMB-resistance mechanisms in C. parapsilosis sensu stricto need to be further investigated and thankfully it is still a rare phenomenon with reports showing an extremely low rate of resistance among different countries and across different regions. [153].

2.5. Flucytosine Resistance

Protein and DNA synthesis are the metabolic target pathways of flucytosine [154]. After administration, flucytosine gets transported into the yeast cell thanks to a cytosine permease and converted to 5-Fluorouracile (5-FU) by a cytosine deaminase present within the yeast pathogens. Next 5-FU gets converted to 5-fluorouridine triphosphate and 5-fluorouridine monophosphate, the first compound alters protein synthesis interfering directly with amino-acylation of tRNA once it has been integrated in the RNA molecule [155,156]. On the other hand, thymidylate synthase is the target of the second active metabolite of 5-FU, resulting in inhibition of DNA synthesis [157]. The resistance mechanisms described for this drug were loss of function mutations correlated with the genes encoding proteins implied in the import of the drug as cytosine permease (FCY2), or involved in its intracellular metabolism as cytosine deaminase (FCY1) and uridine monophosphate phosphorylase (FUR1) [158,159,160]. Such mutations lead to a reduction of these target enzymes reducing both the uptake and the metabolism of flucytosine. Another described mechanism for flucytosine resistance in some resistant Candida spp. strains appears to be overexpression of the substrate increasing pyrimidine synthesis [161,162]. It is also important to mention that Candida species develop flucytosine-resistance rapidly after treatment exposure, even within 48 h after initiation, therefore, international guidelines do not recommend monotherapy with flucytosine supporting instead combination therapy with AMB or azole compounds in selected cases [47].
Data on flucytosine-resistance in C. parapsilosis sensu stricto are scarce, since its use in clinical practice as a single drug agent in monotherapy is commonly avoided. The first report of development of flucytosine-resistance during therapy in C. parapsilosis was described by Hoeperich et al. in 1974 [163], where authors found a reduced cytosine deaminase activity in the resistant strain. No further molecular investigations could be performed at the time to evaluate underlying resistant mutations. However, along with the previously mentioned resistance mechanisms, Sun et al. suggested an adjunctive genetic adaptation in C. parapsilosis regarding yeast response to flucytosine [164]. C. parapsilosis showed to have ortholog genes encoding for the same enzymes implied in the metabolism of flucytosine. In addition, this species was found to have chromosome aneuploidy, in particular trisomy of the chromosome 5 as a potential response and adaptation to the drug [164]. However, such genetic modification does not fully evolve in clinical resistance, anyhow Sun et al. reported that it was correlated with increased antifungal tolerance [164]. To this point, it is worthy to mention that primary objective of their study was to investigate the effect of such genetic modification on caspofungin-resistance and/or tolerance. Later, the authors found a cross-adaptation to flucytosine as they observed an increased tolerance [164]. The biological explanation and interpretation given by the authors refers to a particular gene that is normally found on chromosome 5 in C. parapsilosis that encodes for chitin known as CHS7. Such gene ends up to be overexpressed under trisomy conditions like in this case [164,165].

2.6. Antifungal Tolerance

The concept of drug tolerance was first introduced when observing bacterial isolates able to survive in the presence of antibiotics at concentrations above the MIC without any known underlying resistance mechanism [166,167]. This phenomenon was reflected in vitro by a slow growth of a small proportion of cells within a single colony of the microorganism, showing tolerance to the specific antimicrobial drug [16,29,166]. Such microorganisms, however, did not harbor any known resistance mechanism and once tested again for the selected antimicrobial molecule only a small sub-proportion of them still grew under MIC concentrations, suggesting that tolerance should be referred to a peculiar physiological and/or epigenetic state of the microorganism instead of genetic acquisition of a resistant phenotype [166]. Therefore, at in vitro antifungal susceptibility tests tolerant isolates are included in the susceptible category and cannot be distinguished by non-tolerant ones due to their slow growth [26]. In vitro demonstration of this phenomenon has been defined as “trailing growth” at broth microdilution methods, where wells in which antimicrobial drugs were present at an inhibitory concentration hosted a slow growth of the pathogen [166]. The same concept can be translated in yeast pathogens especially in Candida species [16,168,169]. Authors impute tolerance to the presence of persister cells among the microorganism population tested for antifungal resistance [170,171], as others suggest that aneuploidy might also be involved [172]. As stated by Berman and colleagues in regards of fungal microorganisms, tolerance is the ability of yeasts to slowly grow above MIC values [26]. Such growth would not be detectable before 48 h of incubation [26]. The same authors proposed and hypothesized that different cellular stress-responses among fungi of the same isogenic population might be an explanatory factor contributing to drug tolerance, but data confirming such assumption have yet to be provided [26]. In C. albicans studies pointed out that a different composition in the sphingolipid profile in cell membrane might be involved in the development of azole-tolerance [173]. Also increased chitin synthesis—especially with echinocandin molecules—may play a significant role in defining antifungal tolerance favoring survival of the yeast cells and slow growth rate after 24 h incubation [131,174]. Clinical consequences of drug tolerance represent a fervid field of research, with several authors reporting a correlation with the development of antimicrobial resistance [107,127,175,176] and with mortality and therapeutic failure even in fungal diseases [176,177]. Antifungal drug tolerance varies from one class to the other as it has been more frequently observed with azole compounds rather than echinocandins, since the first class mentioned is known to have a fungistatic effect [16,92]. Indeed, the proportions of C. parapsilosis sensu stricto and C. glabrata cells found within the in vitro trailing growth, that are able to grow slowly under drug concentrations higher than the MIC values, are different between azole and echinocandins [92,127], with more than 1% of total fungal population tolerant to azole drugs and less than 1% to echinocandins, although this last class of drugs has only a fungistatic effect on C. parapsilosis [92]. Antifungal tolerance is difficult to assess via routinely available in vitro tests, therefore, authors proposed specific tests to achieve valuable and interpretable results, such as for echinocandin tolerance in the study from Daneshnia [100]. In this study, C. parapsilosis cells were incubated in RPMI1640 liquid medium added with the intermediate breakpoint micafungin value (4 μg/mL) according to CLSI, and plating was performed at several time intervals comparing Colony Forming Units (CFU) with untreated controls [100]. In this study, the authors reported several mutations implied with echinocandin-tolerance in C. parapsilosis in addition to those previously reported as S745L (found outside the Hot Spot region 1 of the FKS1 gene) and A1422G and M1328I (found both outside the Hot Spot region 2 in the FKS1 gene) [107]. Another in vitro test proposed by Berman and colleagues is the “fraction of growth” [26]. This test compares fungal growth within the MIC inhibition zone on solid medium of tolerant colonies after prolonged incubation time (48 h) and the fungal growth observed beyond the same area. Such measurement allows also for an estimation of the degree of tolerance. [26]. In order to do so, the authors also rely on the use of automated software to estimate such distance quantifying the grade of antifungal tolerance [26]. Even liquid medium tests have been proposed to evaluate such microbiological phenomenon, falling under the name of “Supra MIC growth” [26].
To better elucidate the relevance of antifungal tolerance and its implications on the development of echinocandin resistance, it is mandatory to mention a study from Daneshina et al. [107] where the in vitro selection of echinocandin-resistant C. parapsilosis isolates happened only in echinocandin tolerant cells after being exposed and plated on agar solid medium supplemented with echinocandin highlighting an inducible resistant phenotype from tolerant yeast strains [107].

2.7. Heteroresistance

Heteroresistance was firstly described in bacterial microorganisms as Staphylococcus spp., Acinetobacter spp., Myocobacterium tuberculosis and then in a fungal opportunistic pathogen, Cryptococcus neoformans [178,179,180,181]. As for tolerance, it is a microbiological phenomenon that takes place in a very reduced subset of microorganisms within a bacterial or fungal population differing from the previous one as it happens in one cell in 105–106 CFU of susceptible colonies. Although rarer than tolerance, it correlates with a detectable resistant phenotype at in vitro [167]. Fungal pathogens showing heteroresistance may reach up to eight-fold the MIC values registered in common in vitro susceptibility tests; however, genetic resistance and heteroresitance remain two distinct microbiological phenomena [26]. For example, in two yeast pathogens, C. glabrata and C. neoformans, heteroresistance to fluconazole was observed in less than 1% of fungal population, but it was anyhow linked to selection of the resistant strain and subsequent treatment failure [182,183]. As for tolerance, fungal isolates may show different grades of heteroresistance as hypothesized through a mouse model of C. glabrata kidney infection, where highly heteroresistant isolates correlated with higher percentages of persistent infections [182]. Higher levels of heteroresistance could be associated with clinically relevant consequences in humans. The proposed underlying genetically based resistance mechanism for these two fungal pathogens is target drug/efflux pump gene aneuploidy, but still no consensus among researchers has been reached as aneuploidy could only partially explain the resistant phenotype [183,184]. Another clinically relevant issue related to heteroresistance is that it cannot be detected at standard antimicrobial susceptibility tests, this is caused by the reduced number of microorganisms constituting the heteroresistant population [185,186].
In the case of C. parapsilosis sensu stricto, heteroresistance to echinocandins was appointed by Zhai and colleagues to be correlated with prophylaxis failure, thus enhancing the risk of breakthrough infections [185]. Rates of C. parapsilosis echinocandin heteroresistance ranged between 0.1% and 0.01%, within an otherwise fully susceptible colony [92,185].

3. Epidemiological Landscape of Candida parapsilosis sensu stricto Resistance

Among all antifungal drugs, azoles are the most studied in terms of antimicrobial resistance for C. parapsilosis with rates higher than those reported for all other drugs and continuously increasing in the last twenty years [17,40,92]. According to the 2006–2016 SENTRY surveillance study, 3.9% of C. parapsilosis isolates analyzed were resistant to fluconazole [40], and differences were observed according to the geographical area, with 4.6% of C. parapsilosis strains isolated from European countries and 4.3% from Latin America. In a meta-analysis from Yamin and colleagues, pooled prevalence of fluconazole resistance was 15.2% up to 2022 [153]. Voriconazole resistance rates from the same investigation reported a pooled prevalence of 4.7% in the meta-analysis from Yamin [153] and high cross-resistance rates with fluconazole (32.7% of fluconazole-resistant isolates susceptible to voriconazole) into the SENTRY report [40]. However, further data from monocentric studies revealed higher rates in fluconazole-resistance than the average reported, especially from Europe (10–20% Spain and Greece; 20–30% Italy; 30–40% Turkey) Latin America (10–20%) and South Africa (40–60%) [2,3,6,7,8,92,187,188,189,190,191,192]. Among all, Govender et al. in 2016 described an astonishing rate of fluconazole-resistance in the South African province of Guateng, with only 37% of fluconazole-susceptible C. parapsilosis isolated from bloodstream infections [8].
Data regarding mutations found in azole-resistant C. parapsilosis isolates pointed out that the most frequent alterations leading to the acquisition of a resistant phenotype were the Y132F substitution in ERG11 along with the upregulation of MDR1 especially in European surveys and reports [92,112,187,188,193]. Considering only amino acid substitutions in the ERG11 gene, Ceballos-Garzon and colleagues reported that Y132F was the single point mutation related to azole-resistance in Italy, South Africa, Brazil, Mexico and France [187]. Association between Y132F and K143R was observed in USA, India, Colombia, Spain and Turkey; in the last two states, in addition to Y132F and K143R also the substitution G458S was reported to be present in the same strain [187,193,194].
It has been observed that the spreading of azole-resistant C. parapsilosis happens through hospital outbreaks of invasive infections, especially in the case of strains harboring the Y132F substitution in the ERG11 gene [92,188,192,195]. Along with this, azole-resistant C. parapsilosis is able to persist in the hospital environment causing infections even in patients without a previous history of azole exposure [92]. In-hospital transmission is carried out through contamination of health care environment, medical devices and health-care operators’ hands [196,197]. Noteworthy, most bacterial isolates harboring resistance genes, like in the case of plasmid-based carbapenemases, are selected through antibiotic pressure. In fact, its removal would restore colonization of the susceptible strain within the microbial niche. However, such a behavior is not observed in azole-resistant C. parapsilosis as the majority of patients affected by invasive infections during a hospital outbreak is drug naïve. Therefore, some authors suggested that fitness cost in azole-resistant C. parapsilosis could be equal to the susceptible strain. In addition, in-host survival time is longer for azole-resistant than -susceptible strains, thus highlighting once again that the resistant yeast pathogen is able to better adapt to host’s conditions than susceptible counterparts [4,92,188].
Among other antifungal drugs, echinocandins have been extensively used in the past few years to treat invasive infections caused by azole-resistant C. parapsilosis and are now considered the drug of choice in such clinical scenario [198]. However, as stated previously, this class of molecules displays fungistatic effect on C. parapsilosis as MIC values for echinocandin drugs are higher than for other species due the P660A polymorphism in Fks1p [104]. Despite echinocandin resistance being a seldom clinical phenomenon and rarely reported, some authors described an increased tolerance and acquired resistance [107,108], as described in studies from China [108], Turkey [106], Spain [105], Greece [140] and Brazil [100]. One study from Meletiadis et al. reported a prevalence of echinocandin-resistance in C. parapsilosis of 3.2% [199]. To this point, it is important to mention another study from a multicenter investigation in Spain conducted by Cantón and colleagues reporting a very low prevalence in echinocandin resistance of 0.6% in C. parapsilosis isolates recovered from blood cultures [200].
Data on the use of other antifungal drugs in the context of azole-resistant C. parapsilosis invasive infections are scarce and not often reported [201]. Despite the availability of liposomal formulation of AMB that reduced rates of adverse events, this compound is a therapeutic option reserved for only selected cases. In a clinical survey of more than 2000 isolates recovered from blood cultures, rates of AMB resistance was set up to 3% [202], according to a meta-analysis from Yamin and colleagues pooled prevalence of AMB resistance was 1.3%, with few discrepancies between different geographical regions [153]. However, considering its reduced clinical use in the context of azole-resistant C. parapsilosis, further data and surveys are required.
Last, flucytosine-resistance has been rarely investigated as this molecule retains a narrow clinical niche in which its use might be recommended. However, Ostrosky-Zeichner et al. and Quindos et al. reported rates of flucytosine-resistant C. parapsilosis in between 2–6.4%, respectively [202,203].

4. Conclusions

Azole-resistant C. parapsilosis is a major threat in public health. All three major strategies to develop resistance found in C. albicans have been elucidated in this species, however, the most frequently reported in clinical practice is an association of target mutations due to Y132F substitution in ERG11 along with upregulation of MDR1 conferring fluconazole and voriconazole cross-resistance. Data on development of resistance to other molecules, especially echinocandins, are emerging at a worrisome rate. However, the future focus of research should aim at investigating predisposing conditions and risk factors for the development of acquired resistance before the manifestation of the resistant phenotype itself. Future routinely performed microbiologic in vitro diagnostic tests should, therefore, be able to explore and report different levels of antifungal tolerance and heteroresistance in order to identify patients infected and or colonized with strains at risk of developing resistance.

Author Contributions

Conceptualization, I.F., C.R. and A.L.; methodology, I.F. and N.P.; writing—original draft preparation, I.F.; writing—review and editing, A.L.; visualization, I.F.; supervision, A.T. and A.L. All authors have read and agreed to the published version of the manuscript.

Funding

This study has been supported by The European Union—NextGenerationEU, PNRR «THE» (Tuscany health ecosystem) Spoke 7 Innovating translational medicine-Sub-project 5. Project Code: ECS00000017, CUPI53C22000780001.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

No new data were created.

Conflicts of Interest

The authors declare no conflict of interest. The funders had no role in the design of the study; in the collection, analyses, or interpretation of data; in the writing of the manuscript; or in the decision to publish the results.

References

  1. Seagle, E.E.; Williams, S.L.; Chiller, T.M. Recent Trends in the Epidemiology of Fungal Infections. Infect. Dis. Clin. N. Am. 2021, 35, 237–260. [Google Scholar] [CrossRef] [PubMed]
  2. Martini, C.; Torelli, R.; de Groot, T.; De Carolis, E.; Morandotti, G.A.; De Angelis, G.; Posteraro, B.; Meis, J.F.; Sanguinetti, M. Prevalence and Clonal Distribution of Azole-Resistant Candida parapsilosis Isolates Causing Bloodstream Infections in a Large Italian Hospital. Front. Cell. Infect. Microbiol. 2020, 10, 232. [Google Scholar] [CrossRef] [PubMed]
  3. Ulu Kilic, A.; Alp, E.; Cevahir, F.; Ture, Z.; Yozgat, N. Epidemiology and Cost Implications of Candidemia, a 6-Year Analysis from a Developing Country. Mycoses 2017, 60, 198–203. [Google Scholar] [CrossRef] [PubMed]
  4. Arastehfar, A.; Daneshnia, F.; Hilmioğlu-Polat, S.; Fang, W.; Yaşar, M.; Polat, F.; Metin, D.Y.; Rigole, P.; Coenye, T.; Ilkit, M.; et al. First Report of Candidemia Clonal Outbreak Caused by Emerging Fluconazole-Resistant Candida Parapsilosis Isolates Harboring Y132F and/or Y132F+K143R in Turkey. Antimicrob. Agents Chemother. 2020, 64, e01001-20. [Google Scholar] [CrossRef] [PubMed]
  5. Rodriguez, L.; Bustamante, B.; Huaroto, L.; Agurto, C.; Illescas, R.; Ramirez, R.; Diaz, A.; Hidalgo, J. A Multi-Centric Study of Candida Bloodstream Infection in Lima-Callao, Peru: Species Distribution, Antifungal Resistance and Clinical Outcomes. PLoS ONE 2017, 12, e0175172. [Google Scholar] [CrossRef] [Green Version]
  6. Thomaz, D.Y.; de Almeida, J.N.; Sejas, O.N.E.; Del Negro, G.M.B.; Carvalho, G.O.M.H.; Gimenes, V.M.F.; de Souza, M.E.B.; Arastehfar, A.; Camargo, C.H.; Motta, A.L.; et al. Environmental Clonal Spread of Azole-Resistant Candida parapsilosis with Erg11-Y132F Mutation Causing a Large Candidemia Outbreak in a Brazilian Cancer Referral Center. J. Fungi 2021, 7, 259. [Google Scholar] [CrossRef]
  7. Siopi, M.; Tarpatzi, A.; Kalogeropoulou, E.; Damianidou, S.; Vasilakopoulou, A.; Vourli, S.; Pournaras, S.; Meletiadis, J. Epidemiological Trends of Fungemia in Greece with a Focus on Candidemia during the Recent Financial Crisis: A 10-Year Survey in a Tertiary Care Academic Hospital and Review of Literature. Antimicrob. Agents Chemother. 2020, 64, e01516-19. [Google Scholar] [CrossRef]
  8. Govender, N.P.; Patel, J.; Magobo, R.E.; Naicker, S.; Wadula, J.; Whitelaw, A.; Coovadia, Y.; Kularatne, R.; Govind, C.; Lockhart, S.R.; et al. Emergence of Azole-Resistant Candida parapsilosis Causing Bloodstream Infection: Results from Laboratory-Based Sentinel Surveillance in South Africa. J. Antimicrob. Chemother. 2016, 71, 1994–2004. [Google Scholar] [CrossRef] [Green Version]
  9. Xiao, M.; Sun, Z.-Y.; Kang, M.; Guo, D.-W.; Liao, K.; Chen, S.C.-A.; Kong, F.; Fan, X.; Cheng, J.-W.; Hou, X.; et al. Five-Year National Surveillance of Invasive Candidiasis: Species Distribution and Azole Susceptibility from the China Hospital Invasive Fungal Surveillance Net (CHIF-NET) Study. J. Clin. Microbiol. 2018, 56, e00577-18. [Google Scholar] [CrossRef] [Green Version]
  10. Kakeya, H.; Yamada, K.; Kaneko, Y.; Yanagihara, K.; Tateda, K.; Maesaki, S.; Takesue, Y.; Tomono, K.; Kadota, J.; Kaku, M.; et al. National Trends in the Distribution of Candida Species Causing Candidemia in Japan from 2003 to 2014. Med. Mycol. J. 2018, 59, E19–E22. [Google Scholar] [CrossRef] [Green Version]
  11. Tóth, R.; Nosek, J.; Mora-Montes, H.M.; Gabaldon, T.; Bliss, J.M.; Nosanchuk, J.D.; Turner, S.A.; Butler, G.; Vágvölgyi, C.; Gácser, A. Candida parapsilosis: From Genes to the Bedside. Clin. Microbiol. Rev. 2019, 32, e00111-18. [Google Scholar] [CrossRef] [Green Version]
  12. Lupetti, A.; Tavanti, A.; Davini, P.; Ghelardi, E.; Corsini, V.; Merusi, I.; Boldrini, A.; Campa, M.; Senesi, S. Horizontal Transmission of Candida parapsilosis Candidemia in a Neonatal Intensive Care Unit. J. Clin. Microbiol. 2002, 40, 2363–2369. [Google Scholar] [CrossRef] [Green Version]
  13. Zoppo, M.; Fiorentini, F.; Rizzato, C.; Di Luca, M.; Lupetti, A.; Bottai, D.; Colone, M.; Stringaro, A.; De Bernardis, F.; Tavanti, A. Role of CpALS4790 and CpALS0660 in Candida parapsilosis Virulence: Evidence from a Murine Model of Vaginal Candidiasis. J. Fungi 2020, 6, 86. [Google Scholar] [CrossRef]
  14. Branco, J.; Miranda, I.M.; Rodrigues, A.G. Candida parapsilosis Virulence and Antifungal Resistance Mechanisms: A Comprehensive Review of Key Determinants. J. Fungi 2023, 9, 80. [Google Scholar] [CrossRef]
  15. Trofa, D.; Gácser, A.; Nosanchuk, J.D. Candida parapsilosis, an Emerging Fungal Pathogen. Clin. Microbiol. Rev. 2008, 21, 606–625. [Google Scholar] [CrossRef] [Green Version]
  16. Gow, N.A.R.; Johnson, C.; Berman, J.; Coste, A.T.; Cuomo, C.A.; Perlin, D.S.; Bicanic, T.; Harrison, T.S.; Wiederhold, N.; Bromley, M.; et al. The Importance of Antimicrobial Resistance in Medical Mycology. Nat. Commun. 2022, 13, 5352. [Google Scholar] [CrossRef]
  17. 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]
  18. Wiederhold, N.P. Pharmacodynamics, Mechanisms of Action and Resistance, and Spectrum of Activity of New Antifungal Agents. J. Fungi 2022, 8, 857. [Google Scholar] [CrossRef]
  19. Lupetti, A.; Paulusma-Annema, A.; Welling, M.M.; Senesi, S.; van Dissel, J.T.; Nibbering, P.H. Candidacidal Activities of Human Lactoferrin Peptides Derived from the N Terminus. Antimicrob. Agents Chemother. 2000, 44, 3257–3263. [Google Scholar] [CrossRef] [Green Version]
  20. Lupetti, A.; Danesi, R.; van ’t Wout, J.W.; van Dissel, J.T.; Senesi, S.; Nibbering, P.H. Antimicrobial Peptides: Therapeutic Potential for the Treatment of Candida Infections. Expert Opin. Investig. Drugs 2002, 11, 309–318. [Google Scholar] [CrossRef]
  21. Fais, R.; Rizzato, C.; Franconi, I.; Tavanti, A.; Lupetti, A. Synergistic Activity of the Human Lactoferricin-Derived Peptide HLF1-11 in Combination with Caspofungin against Candida Species. Microbiol. Spectr. 2022, 10, e0124022. [Google Scholar] [CrossRef] [PubMed]
  22. Fais, R.; Di Luca, M.; Rizzato, C.; Morici, P.; Bottai, D.; Tavanti, A.; Lupetti, A. The N-Terminus of Human Lactoferrin Displays Anti-Biofilm Activity on Candida parapsilosis in Lumen Catheters. Front. Microbiol. 2017, 8, 2218. [Google Scholar] [CrossRef] [PubMed]
  23. Fisher, M.C.; Denning, D.W. The WHO Fungal Priority Pathogens List as a Game-Changer. Nat. Rev. Microbiol. 2023, 21, 211–212. [Google Scholar] [CrossRef] [PubMed]
  24. WHO Fungal Priority Pathogens List to Guide Research, Development and Public Health Action. Available online: https://www.who.int/publications/i/item/9789240060241 (accessed on 15 February 2023).
  25. Turnidge, J.; Paterson, D.L. Setting and Revising Antibacterial Susceptibility Breakpoints. Clin. Microbiol. Rev. 2007, 20, 391–408. [Google Scholar] [CrossRef] [Green Version]
  26. Berman, J.; Krysan, D.J. Drug Resistance and Tolerance in Fungi. Nat. Rev. Microbiol. 2020, 18, 319–331. [Google Scholar] [CrossRef]
  27. Bhattacharya, S.; Sae-Tia, S.; Fries, B.C. Candidiasis and Mechanisms of Antifungal Resistance. Antibiotics 2020, 9, 312. [Google Scholar] [CrossRef]
  28. Pristov, K.E.; Ghannoum, M.A. Resistance of Candida to Azoles and Echinocandins Worldwide. Clin. Microbiol. Infect. 2019, 25, 792–798. [Google Scholar] [CrossRef]
  29. 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] [Green Version]
  30. Eucast: Expected Phenotypes. Available online: https://www.eucast.org/expert_rules_and_expected_phenotypes/expected_phenotypes (accessed on 14 May 2023).
  31. Dudiuk, C.; Macedo, D.; Leonardelli, F.; Theill, L.; Cabeza, M.S.; Gamarra, S.; Garcia-Effron, G. Molecular Confirmation of the Relationship between Candida guilliermondii Fks1p Naturally Occurring Amino Acid Substitutions and Its Intrinsic Reduced Echinocandin Susceptibility. Antimicrob. Agents Chemother. 2017, 61, e02644-16. [Google Scholar] [CrossRef] [Green Version]
  32. Buil, J.B.; Oliver, J.D.; Law, D.; Baltussen, T.; Zoll, J.; Hokken, M.W.J.; Tehupeiory-Kooreman, M.; Melchers, W.J.G.; Birch, M.; Verweij, P.E. Resistance Profiling of Aspergillus Fumigatus to Olorofim Indicates Absence of Intrinsic Resistance and Unveils the Molecular Mechanisms of Acquired Olorofim Resistance. Emerg. Microbes Infect. 2022, 11, 703–714. [Google Scholar] [CrossRef]
  33. Arendrup, M.C.; Patterson, T.F. Multidrug-Resistant Candida: Epidemiology, Molecular Mechanisms, and Treatment. J. Infect. Dis. 2017, 216, S445–S451. [Google Scholar] [CrossRef] [Green Version]
  34. Wiederhold, N.P. Antifungal Susceptibility Testing: A Primer for Clinicians. Open Forum Infect. Dis. 2021, 8, ofab444. [Google Scholar] [CrossRef]
  35. Bienvenu, A.L.; Argaud, L.; Aubrun, F.; Fellahi, J.L.; Guerin, C.; Javouhey, E.; Piriou, V.; Rimmele, T.; Chidiac, C.; Leboucher, G. A Systematic Review of Interventions and Performance Measures for Antifungal Stewardship Programmes. J. Antimicrob. Chemother. 2018, 73, 297–305. [Google Scholar] [CrossRef] [Green Version]
  36. Pfaller, M.A. Antifungal Drug Resistance: Mechanisms, Epidemiology, and Consequences for Treatment. Am. J. Med. 2012, 125, S3–S13. [Google Scholar] [CrossRef]
  37. Cowen, L.E.; Sanglard, D.; Howard, S.J.; Rogers, P.D.; Perlin, D.S. Mechanisms of Antifungal Drug Resistance. Cold Spring Harb. Perspect. Med. 2015, 5, a019752. [Google Scholar] [CrossRef]
  38. Clinical and Laboratory Standards Institute (CLSI). Reference Method for Broth Dilution Antifungal Susceptibility Testing of Filamentous Fungi, 3rd ed.; M38-A2; CLSI: Wayne, PA, USA, 2017. [Google Scholar]
  39. Eucast: Breakpoints for Antifungals. Available online: https://www.eucast.org/astoffungi/clinicalbreakpointsforantifungals (accessed on 5 December 2022).
  40. 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]
  41. Zoppo, M.; Poma, N.; Di Luca, M.; Bottai, D.; Tavanti, A. Genetic Manipulation as a Tool to Unravel Candida Parapsilosis Species Complex Virulence and Drug Resistance: State of the Art. J. Fungi 2021, 7, 459. [Google Scholar] [CrossRef]
  42. Kelly, S.L.; Lamb, D.C.; Corran, A.J.; Baldwin, B.C.; Kelly, D.E. Mode of Action and Resistance to Azole Antifungals Associated with the Formation of 14 Alpha-Methylergosta-8,24(28)-Dien-3 Beta,6 Alpha-Diol. Biochem. Biophys. Res. Commun. 1995, 207, 910–915. [Google Scholar] [CrossRef]
  43. Lupetti, A.; Danesi, R.; Campa, M.; Del Tacca, M.; Kelly, S. Molecular Basis of Resistance to Azole Antifungals. Trends Mol. Med. 2002, 8, 76–81. [Google Scholar] [CrossRef]
  44. Spampinato, C.; Leonardi, D. Candida Infections, Causes, Targets, and Resistance Mechanisms: Traditional and Alternative Antifungal Agents. Biomed. Res. Int. 2013, 2013, 204237. [Google Scholar] [CrossRef] [Green Version]
  45. Delattin, N.; Cammue, B.P.A.; Thevissen, K. Reactive Oxygen Species-Inducing Antifungal Agents and Their Activity against Fungal Biofilms. Future Med. Chem. 2014, 6, 77–90. [Google Scholar] [CrossRef] [PubMed]
  46. Cornely, O.A.; Bassetti, M.; Calandra, T.; Garbino, J.; Kullberg, B.J.; Lortholary, O.; Meersseman, W.; Akova, M.; Arendrup, M.C.; Arikan-Akdagli, S.; et al. ESCMID* Guideline for the Diagnosis and Management of Candida Diseases 2012: Non-Neutropenic Adult Patients. Clin. Microbiol. Infect. 2012, 18 (Suppl. S7), 19–37. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  47. 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]
  48. Perea, S.; López-Ribot, J.L.; Kirkpatrick, W.R.; McAtee, R.K.; Santillán, R.A.; Martínez, 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] [Green Version]
  49. White, T.C. Increased MRNA Levels of ERG16, CDR, and MDR1 Correlate with Increases in Azole Resistance in Candida albicans Isolates from a Patient Infected with Human Immunodeficiency Virus. Antimicrob. Agents Chemother. 1997, 41, 1482–1487. [Google Scholar] [CrossRef] [Green Version]
  50. Nishimoto, A.T.; Sharma, C.; Rogers, P.D. Molecular and Genetic Basis of Azole Antifungal Resistance in the Opportunistic Pathogenic Fungus Candida albicans. J. Antimicrob. Chemother. 2020, 75, 257–270. [Google Scholar] [CrossRef]
  51. Paulsen, I.T.; Brown, M.H.; Skurray, R.A. Proton-Dependent Multidrug Efflux Systems. Microbiol. Rev. 1996, 60, 575–608. [Google Scholar] [CrossRef]
  52. Gbelska, Y.; Krijger, J.-J.; Breunig, K.D. Evolution of Gene Families: The Multidrug Resistance Transporter Genes in Five Related Yeast Species. FEMS Yeast Res. 2006, 6, 345–355. [Google Scholar] [CrossRef] [Green Version]
  53. 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] [Green Version]
  54. Sasse, C.; Dunkel, N.; Schäfer, T.; Schneider, S.; Dierolf, F.; Ohlsen, K.; Morschhäuser, J. The Stepwise Acquisition of Fluconazole Resistance Mutations Causes a Gradual Loss of Fitness in Candida albicans. Mol. Microbiol. 2012, 86, 539–556. [Google Scholar] [CrossRef]
  55. Prasad, R.; Rawal, M.K.; Shah, A.H. Candida Efflux ATPases and Antiporters in Clinical Drug Resistance. Adv. Exp. Med. Biol. 2016, 892, 351–376. [Google Scholar] [CrossRef]
  56. 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]
  57. Tsao, S.; Rahkhoodaee, F.; Raymond, M. Relative Contributions of the Candida albicans ABC Transporters Cdr1p and Cdr2p to Clinical Azole Resistance. Antimicrob. Agents Chemother. 2009, 53, 1344–1352. [Google Scholar] [CrossRef] [Green Version]
  58. Calabrese, D.; Bille, J.; Sanglard, D. A Novel Multidrug Efflux Transporter Gene of the Major Facilitator Superfamily from Candida albicans (FLU1) Conferring Resistance to Fluconazole. Microbiology 2000, 146 Pt 11, 2743–2754. [Google Scholar] [CrossRef] [Green Version]
  59. Coste, A.; Turner, V.; Ischer, F.; Morschhäuser, 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] [Green Version]
  60. Dunkel, N.; Blass, J.; Rogers, P.D.; Morschhäuser, 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] [Green Version]
  61. Chen, C.-G.; Yang, Y.-L.; Shih, H.-I.; Su, C.-L.; Lo, H.-J. CaNdt80 Is Involved in Drug Resistance in Candida albicans by Regulating CDR1. Antimicrob. Agents Chemother. 2004, 48, 4505–4512. [Google Scholar] [CrossRef] [Green Version]
  62. Mogavero, S.; Tavanti, A.; Senesi, S.; Rogers, P.D.; Morschhäuser, J. Differential Requirement of the Transcription Factor Mcm1 for Activation of the Candida albicans Multidrug Efflux Pump MDR1 by Its Regulators Mrr1 and Cap1. Antimicrob. Agents Chemother. 2011, 55, 2061–2066. [Google Scholar] [CrossRef] [Green Version]
  63. Ferrari, S.; Ischer, F.; Calabrese, D.; Posteraro, B.; Sanguinetti, M.; Fadda, G.; Rohde, B.; Bauser, C.; Bader, O.; Sanglard, D. Gain of Function Mutations in CgPDR1 of Candida glabrata Not Only Mediate Antifungal Resistance but Also Enhance Virulence. PLoS Pathog. 2009, 5, e1000268. [Google Scholar] [CrossRef] [Green Version]
  64. Arthington-Skaggs, B.A.; Jradi, H.; Desai, T.; Morrison, C.J. Quantitation of Ergosterol Content: Novel Method for Determination of Fluconazole Susceptibility of Candida albicans. J. Clin. Microbiol. 1999, 37, 3332–3337. [Google Scholar] [CrossRef]
  65. Veen, M.; Stahl, U.; Lang, C. Combined Overexpression of Genes of the Ergosterol Biosynthetic Pathway Leads to Accumulation of Sterols in Saccharomyces Cerevisiae. FEMS Yeast Res. 2003, 4, 87–95. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  66. 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]
  67. Cuenca-Estrella, M. Antifungal Drug Resistance Mechanisms in Pathogenic Fungi: From Bench to Bedside. Clin. Microbiol. Infect. 2014, 20, 54–59. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  68. Anderson, J.B. Evolution of Antifungal-Drug Resistance: Mechanisms and Pathogen Fitness. Nat. Rev. Microbiol. 2005, 3, 547–556. [Google Scholar] [CrossRef] [PubMed]
  69. Pam, V.K.; Akpan, J.U.; Oduyebo, O.O.; Nwaokorie, F.O.; Fowora, M.A.; Oladele, R.O.; Ogunsola, F.T.; Smith, S.I. Fluconazole Susceptibility and ERG11 Gene Expression in Vaginal Candida Species Isolated from Lagos Nigeria. Int. J. Mol. Epidemiol. Genet. 2012, 3, 84–90. [Google Scholar]
  70. Xiang, M.-J.; Liu, J.-Y.; Ni, P.-H.; Wang, S.; Shi, C.; Wei, B.; Ni, Y.-X.; Ge, H.-L. Erg11 Mutations Associated with Azole Resistance in Clinical Isolates of Candida albicans. FEMS Yeast Res. 2013, 13, 386–393. [Google Scholar] [CrossRef] [Green Version]
  71. Zhang, L.; Yang, H.-F.; Liu, Y.-Y.; Xu, X.-H.; Ye, Y.; Li, J.-B. Reduced Susceptibility of Candida albicans Clinical Isolates to Azoles and Detection of Mutations in the ERG11 Gene. Diagn. Microbiol. Infect. Dis. 2013, 77, 327–329. [Google Scholar] [CrossRef]
  72. Yang, H.; Tong, J.; Lee, C.W.; Ha, S.; Eom, S.H.; Im, Y.J. Structural Mechanism of Ergosterol Regulation by Fungal Sterol Transcription Factor Upc2. Nat. Commun. 2015, 6, 6129. [Google Scholar] [CrossRef] [Green Version]
  73. 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]
  74. Nagi, M.; Nakayama, H.; Tanabe, K.; Bard, M.; Aoyama, T.; Okano, M.; Higashi, S.; Ueno, K.; Chibana, H.; Niimi, M.; et al. Transcription Factors CgUPC2A and CgUPC2B Regulate Ergosterol Biosynthetic Genes in Candida glabrata. Genes Cells 2011, 16, 80–89. [Google Scholar] [CrossRef]
  75. Dos Silva, D.B.; Rodrigues, L.M.C.; de Almeida, A.A.; de Oliveira, K.M.P.; Grisolia, A.B. Novel Point Mutations in the ERG11 Gene in Clinical Isolates of Azole Resistant Candida Species. Mem. Inst. Oswaldo Cruz 2016, 111, 192–199. [Google Scholar] [CrossRef] [Green Version]
  76. Hull, C.M.; Parker, J.E.; Bader, O.; Weig, M.; Gross, U.; Warrilow, A.G.S.; Kelly, D.E.; Kelly, S.L. Facultative Sterol Uptake in an Ergosterol-Deficient Clinical Isolate of Candida glabrata Harboring a Missense Mutation in ERG11 and Exhibiting Cross-Resistance to Azoles and Amphotericin B. Antimicrob. Agents Chemother. 2012, 56, 4223–4232. [Google Scholar] [CrossRef] [Green Version]
  77. Chau, A.S.; Gurnani, M.; Hawkinson, R.; Laverdiere, M.; Cacciapuoti, A.; McNicholas, P.M. Inactivation of Sterol Δ5,6-Desaturase Attenuates Virulence in Candida albicans. Antimicrob. Agents Chemother. 2005, 49, 3646–3651. [Google Scholar] [CrossRef] [Green Version]
  78. 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] [Green Version]
  79. Xu, D.; Jiang, B.; Ketela, T.; Lemieux, S.; Veillette, K.; Martel, N.; Davison, J.; Sillaots, S.; Trosok, S.; Bachewich, C.; et al. Genome-Wide Fitness Test and Mechanism-of-Action Studies of Inhibitory Compounds in Candida albicans. PLoS Pathog. 2007, 3, e92. [Google Scholar] [CrossRef] [Green Version]
  80. Arthington, B.A.; Bennett, L.G.; Skatrud, P.L.; Guynn, C.J.; Barbuch, R.J.; Ulbright, C.E.; Bard, M. Cloning, Disruption and Sequence of the Gene Encoding Yeast C-5 Sterol Desaturase. Gene 1991, 102, 39–44. [Google Scholar] [CrossRef]
  81. Morio, F.; Pagniez, F.; Lacroix, C.; Miegeville, M.; Le Pape, P. Amino Acid Substitutions in the Candida albicans Sterol Δ5,6-Desaturase (Erg3p) Confer Azole Resistance: Characterization of Two Novel Mutants with Impaired Virulence. J. Antimicrob. Chemother. 2012, 67, 2131–2138. [Google Scholar] [CrossRef] [Green Version]
  82. Ford, C.B.; Funt, J.M.; Abbey, D.; Issi, L.; Guiducci, C.; Martinez, D.A.; Delorey, T.; Li, B.Y.; White, T.C.; Cuomo, C.; et al. The Evolution of Drug Resistance in Clinical Isolates of Candida albicans. Elife 2015, 4, e00662. [Google Scholar] [CrossRef] [Green Version]
  83. Selmecki, A.; Forche, A.; Berman, J. Aneuploidy and Isochromosome Formation in Drug-Resistant Candida albicans. Science 2006, 313, 367–370. [Google Scholar] [CrossRef] [Green Version]
  84. Marichal, P.; Vanden Bossche, H.; Odds, F.C.; Nobels, G.; Warnock, D.W.; Timmerman, V.; Van Broeckhoven, C.; Fay, S.; Mose-Larsen, P. Molecular Biological Characterization of an Azole-Resistant Candida glabrata Isolate. Antimicrob. Agents Chemother. 1997, 41, 2229–2237. [Google Scholar] [CrossRef] [Green Version]
  85. Perepnikhatka, V.; Fischer, F.J.; Niimi, M.; Baker, R.A.; Cannon, R.D.; Wang, Y.K.; Sherman, F.; Rustchenko, E. Specific Chromosome Alterations in Fluconazole-Resistant Mutants of Candida albicans. J. Bacteriol. 1999, 181, 4041–4049. [Google Scholar] [CrossRef] [PubMed]
  86. Marcos-Zambrano, L.J.; Gómez, A.; Sánchez-Carrillo, C.; Bouza, E.; Muñoz, P.; Escribano, P.; Guinea, J. Isavuconazole Is Highly Active in Vitro against Candida Species Isolates but Shows Trailing Effect. Clin. Microbiol. Infect. 2018, 24, 1343.e1–1343.e4. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  87. Desnos-Ollivier, M.; Bretagne, S.; Boullié, A.; Gautier, C.; Dromer, F.; Lortholary, O. French Mycoses Study Group Isavuconazole MIC Distribution of 29 Yeast Species Responsible for Invasive Infections (2015–2017). Clin. Microbiol. Infect. 2019, 25, 634.e1–634.e4. [Google Scholar] [CrossRef] [PubMed]
  88. Kullberg, B.J.; Viscoli, C.; Pappas, P.G.; Vazquez, J.; Ostrosky-Zeichner, L.; Rotstein, C.; Sobel, J.D.; Herbrecht, R.; Rahav, G.; Jaruratanasirikul, S.; et al. Isavuconazole Versus Caspofungin in the Treatment of Candidemia and Other Invasive Candida Infections: The ACTIVE Trial. Clin. Infect. Dis. 2019, 68, 1981–1989. [Google Scholar] [CrossRef] [Green Version]
  89. Logan, A.; Wolfe, A.; Williamson, J.C. Antifungal Resistance and the Role of New Therapeutic Agents. Curr. Infect. Dis. Rep. 2022, 24, 105–116. [Google Scholar] [CrossRef]
  90. Ellsworth, M.; Ostrosky-Zeichner, L. Isavuconazole: Mechanism of Action, Clinical Efficacy, and Resistance. J. Fungi 2020, 6, 324. [Google Scholar] [CrossRef]
  91. Sanglard, D.; Coste, A.T. Activity of Isavuconazole and Other Azoles against Candida Clinical Isolates and Yeast Model Systems with Known Azole Resistance Mechanisms. Antimicrob. Agents Chemother. 2016, 60, 229–238. [Google Scholar] [CrossRef] [Green Version]
  92. Daneshnia, F.; de Almeida Júnior, J.N.; Ilkit, M.; Lombardi, L.; Perry, A.M.; Gao, M.; Nobile, C.J.; Egger, M.; Perlin, D.S.; Zhai, B.; et al. Worldwide Emergence of Fluconazole-Resistant Candida parapsilosis: Current Framework and Future Research Roadmap. Lancet Microbe 2023, 4, e470–e480. [Google Scholar] [CrossRef]
  93. Silva, A.P.; Miranda, I.M.; Guida, A.; Synnott, J.; Rocha, R.; Silva, R.; Amorim, A.; Pina-Vaz, C.; Butler, G.; Rodrigues, A.G. Transcriptional Profiling of Azole-Resistant Candida parapsilosis Strains. Antimicrob. Agents Chemother. 2011, 55, 3546–3556. [Google Scholar] [CrossRef] [Green Version]
  94. Branco, J.; Silva, A.P.; Silva, R.M.; Silva-Dias, A.; Pina-Vaz, C.; Butler, G.; Rodrigues, A.G.; Miranda, I.M. Fluconazole and Voriconazole Resistance in Candida parapsilosis Is Conferred by Gain-of-Function Mutations in MRR1 Transcription Factor Gene. Antimicrob. Agents Chemother. 2015, 59, 6629–6633. [Google Scholar] [CrossRef] [Green Version]
  95. Souza, A.C.R.; Fuchs, B.B.; Pinhati, H.M.S.; Siqueira, R.A.; Hagen, F.; Meis, J.F.; Mylonakis, E.; Colombo, A.L. Candida parapsilosis Resistance to Fluconazole: Molecular Mechanisms and In Vivo Impact in Infected Galleria Mellonella Larvae. Antimicrob. Agents Chemother. 2015, 59, 6581–6587. [Google Scholar] [CrossRef] [Green Version]
  96. Zhang, L.; Xiao, M.; Watts, M.R.; Wang, H.; Fan, X.; Kong, F.; Xu, Y.-C. Development of Fluconazole Resistance in a Series of Candida parapsilosis Isolates from a Persistent Candidemia Patient with Prolonged Antifungal Therapy. BMC Infect. Dis. 2015, 15, 340. [Google Scholar] [CrossRef] [Green Version]
  97. Berkow, E.L.; Manigaba, K.; Parker, J.E.; Barker, K.S.; Kelly, S.L.; Rogers, P.D. Multidrug Transporters and Alterations in Sterol Biosynthesis Contribute to Azole Antifungal Resistance in Candida parapsilosis. Antimicrob. Agents Chemother. 2015, 59, 5942–5950. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  98. Branco, J.; Ola, M.; Silva, R.M.; Fonseca, E.; Gomes, N.C.; Martins-Cruz, C.; Silva, A.P.; Silva-Dias, A.; Pina-Vaz, C.; Erraught, C.; et al. Impact of ERG3 Mutations and Expression of Ergosterol Genes Controlled by UPC2 and NDT80 in Candida parapsilosis Azole Resistance. Clin. Microbiol. Infect. 2017, 23, 575.e1–575.e8. [Google Scholar] [CrossRef]
  99. Rybak, J.M.; Dickens, C.M.; Parker, J.E.; Caudle, K.E.; Manigaba, K.; Whaley, S.G.; Nishimoto, A.T.; Luna-Tapia, A.; Roy, S.; Zhang, Q.; et al. Loss of C-5 Sterol Desaturase Activity Results in Increased Resistance to Azole and Echinocandin Antifungals in a Clinical Isolate of Candida parapsilosis. Antimicrob. Agents Chemother. 2017, 61, e00651-17. [Google Scholar] [CrossRef] [Green Version]
  100. Daneshnia, F.; de Almeida Júnior, J.N.; Arastehfar, A.; Lombardi, L.; Shor, E.; Moreno, L.; Verena Mendes, A.; Goreth Barberino, M.; Thomaz Yamamoto, D.; Butler, G.; et al. Determinants of Fluconazole Resistance and Echinocandin Tolerance in C. parapsilosis Isolates Causing a Large Clonal Candidemia Outbreak among COVID-19 Patients in a Brazilian ICU. Emerg. Microbes Infect. 2022, 11, 2264–2274. [Google Scholar] [CrossRef]
  101. Díaz-García, J.; Gómez, A.; Alcalá, L.; Reigadas, E.; Sánchez-Carrillo, C.; Pérez-Ayala, A.; Gómez-García de la Pedrosa, E.; González-Romo, F.; Merino-Amador, P.; Cuétara, M.S.; et al. Evidence of Fluconazole-Resistant Candida parapsilosis Genotypes Spreading across Hospitals Located in Madrid, Spain and Harboring the Y132F ERG11p Substitution. Antimicrob. Agents Chemother. 2022, 66, e0071022. [Google Scholar] [CrossRef]
  102. Doorley, L.A.; Rybak, J.M.; Berkow, E.L.; Zhang, Q.; Morschhäuser, J.; Rogers, P.D. Candida parapsilosis Mdr1B and Cdr1B Are Drivers of Mrr1-Mediated Clinical Fluconazole Resistance. Antimicrob. Agents Chemother. 2022, 66, e00289-22. [Google Scholar] [CrossRef]
  103. Branco, J.; Ryan, A.P.; Pinto E Silva, A.; Butler, G.; Miranda, I.M.; Rodrigues, A.G. Clinical Azole Cross-Resistance in Candida Parapsilosis Is Related to a Novel MRR1 Gain-of-Function Mutation. Clin. Microbiol. Infect. 2022, 28, 1655.e5–1655.e8. [Google Scholar] [CrossRef]
  104. Garcia-Effron, G.; Katiyar, S.K.; Park, S.; Edlind, T.D.; Perlin, D.S. A Naturally Occurring Proline-to-Alanine Amino Acid Change in Fks1p in Candida parapsilosis, Candida orthopsilosis, and Candida metapsilosis Accounts for Reduced Echinocandin Susceptibility. Antimicrob. Agents Chemother. 2008, 52, 2305–2312. [Google Scholar] [CrossRef] [Green Version]
  105. Arastehfar, A.; Daneshnia, F.; Hilmioglu-Polat, S.; Ilkit, M.; Yasar, M.; Polat, F.; Metin, D.Y.; Dokumcu, Ü.Z.; Pan, W.; Hagen, F.; et al. Genetically Related Micafungin-Resistant Candida parapsilosis Blood Isolates Harbouring Novel Mutation R658G in Hotspot 1 of Fks1p: A New Challenge? J. Antimicrob. Chemother. 2021, 76, 418–422. [Google Scholar] [CrossRef] [PubMed]
  106. Martí-Carrizosa, M.; Sánchez-Reus, F.; March, F.; Cantón, E.; Coll, P. Implication of Candida parapsilosis FKS1 and FKS2 Mutations in Reduced Echinocandin Susceptibility. Antimicrob. Agents Chemother. 2015, 59, 3570–3573. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  107. Daneshnia, F.; Arastehfar, A.; Lombardi, L.; Binder, U.; Scheler, J.; Vahedi, R.; Hagen, F.; Lass-Flörl, C.; Mansour, M.K.; Butler, G.; et al. Candida parapsilosis Isolates Carrying Mutations Outside of FKS1 Hotspot Regions Confer High Echinocandin Tolerance and Facilitate the Development of Echinocandin Resistance. Int. J. Antimicrob. Agents 2023, 62, 106831. [Google Scholar] [CrossRef] [PubMed]
  108. Ning, Y.; Xiao, M.; Perlin, D.S.; Zhao, Y.; Lu, M.; Li, Y.; Luo, Z.; Dai, R.; Li, S.; Xu, J.; et al. Decreased Echinocandin Susceptibility in Candida parapsilosis Causing Candidemia and Emergence of a Pan-Echinocandin Resistant Case in China. Emerg. Microbes Infect. 2023, 12, 2153086. [Google Scholar] [CrossRef] [PubMed]
  109. Pinto e Silva, A.T.; Costa-de-Oliveira, S.; Silva-Dias, A.; Pina-Vaz, C.; Rodrigues, A.G. Dynamics of in Vitro Acquisition of Resistance by Candida Parapsilosis to Different Azoles. FEMS Yeast Res. 2009, 9, 626–633. [Google Scholar] [CrossRef] [Green Version]
  110. Li, X.; Brown, N.; Chau, A.S.; López-Ribot, J.L.; Ruesga, M.T.; Quindos, G.; Mendrick, C.A.; Hare, R.S.; Loebenberg, D.; DiDomenico, B.; et al. Changes in Susceptibility to Posaconazole in Clinical Isolates of Candida albicans. J. Antimicrob. Chemother. 2004, 53, 74–80. [Google Scholar] [CrossRef]
  111. Grossman, N.T.; Pham, C.D.; Cleveland, A.A.; Lockhart, S.R. Molecular Mechanisms of Fluconazole Resistance in Candida Parapsilosis Isolates from a U.S. Surveillance System. Antimicrob. Agents Chemother. 2015, 59, 1030–1037. [Google Scholar] [CrossRef] [Green Version]
  112. Castanheira, M.; Deshpande, L.M.; Messer, S.A.; Rhomberg, P.R.; Pfaller, M.A. Analysis of Global Antifungal Surveillance Results Reveals Predominance of Erg11 Y132F Alteration among Azole-Resistant Candida parapsilosis and Candida tropicalis and Country-Specific Isolate Dissemination. Int. J. Antimicrob. Agents 2020, 55, 105799. [Google Scholar] [CrossRef]
  113. Rizzato, C.; Poma, N.; Zoppo, M.; Posteraro, B.; Mello, E.; Bottai, D.; Lupetti, A.; Sanguinetti, M.; Tavanti, A. CoERG11 A395T Mutation Confers Azole Resistance in Candida orthopsilosis Clinical Isolates. J. Antimicrob. Chemother. 2018, 73, 1815–1822. [Google Scholar] [CrossRef]
  114. Arastehfar, A.; Khodavaisy, S.; Daneshnia, F.; Najafzadeh, M.-J.; Mahmoudi, S.; Charsizadeh, A.; Salehi, M.-R.; Zarrinfar, H.; Raeisabadi, A.; Dolatabadi, S.; et al. Molecular Identification, Genotypic Diversity, Antifungal Susceptibility, and Clinical Outcomes of Infections Caused by Clinically Underrated Yeasts, Candida orthopsilosis, and Candida metapsilosis: An Iranian Multicenter Study (2014–2019). Front. Cell. Infect. Microbiol. 2019, 9, 264. [Google Scholar] [CrossRef] [Green Version]
  115. Zhang, W.; Zhan, M.; Wang, N.; Fan, J.; Han, X.; Li, C.; Liu, J.; Li, J.; Hou, Y.; Wang, X.; et al. In Vitro Susceptibility Profiles of Candida parapsilosis Species Complex Subtypes from Deep Infections to Nine Antifungal Drugs. J. Med. Microbiol. 2023, 72, 001640. [Google Scholar] [CrossRef]
  116. Douglas, C.M.; D’Ippolito, J.A.; Shei, G.J.; Meinz, M.; Onishi, J.; Marrinan, J.A.; Li, W.; Abruzzo, G.K.; Flattery, A.; Bartizal, K.; et al. Identification of the FKS1 Gene of Candida albicans as the Essential Target of 1,3-Beta-D-Glucan Synthase Inhibitors. Antimicrob. Agents Chemother. 1997, 41, 2471–2479. [Google Scholar] [CrossRef] [Green Version]
  117. Grover, N.D. Echinocandins: A Ray of Hope in Antifungal Drug Therapy. Indian J. Pharmacol. 2010, 42, 9–11. [Google Scholar] [CrossRef] [Green Version]
  118. Reboli, A.C.; Rotstein, C.; Pappas, P.G.; Chapman, S.W.; Kett, D.H.; Kumar, D.; Betts, R.; Wible, M.; Goldstein, B.P.; Schranz, J.; et al. Anidulafungin versus Fluconazole for Invasive Candidiasis. N. Engl. J. Med. 2007, 356, 2472–2482. [Google Scholar] [CrossRef] [Green Version]
  119. Farmakiotis, D.; Tarrand, J.J.; Kontoyiannis, D.P. Drug-Resistant Candida glabrata Infection in Cancer Patients. Emerg. Infect. Dis. 2014, 20, 1833–1840. [Google Scholar] [CrossRef] [Green Version]
  120. Medici, N.P.; Del Poeta, M. New Insights on the Development of Fungal Vaccines: From Immunity to Recent Challenges. Mem. Inst. Oswaldo Cruz 2015, 110, 966–973. [Google Scholar] [CrossRef] [Green Version]
  121. Walker, L.A.; Gow, N.A.R.; Munro, C.A. Fungal Echinocandin Resistance. Fungal Genet. Biol. 2010, 47, 117–126. [Google Scholar] [CrossRef] [Green Version]
  122. 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]
  123. Balashov, S.V.; Park, S.; Perlin, D.S. Assessing Resistance to the Echinocandin Antifungal Drug Caspofungin in Candida albicans by Profiling Mutations in FKS1. Antimicrob. Agents Chemother. 2006, 50, 2058–2063. [Google Scholar] [CrossRef] [Green Version]
  124. Katiyar, S.K.; Alastruey-Izquierdo, A.; Healey, K.R.; Johnson, M.E.; Perlin, D.S.; Edlind, T.D. Fks1 and Fks2 Are Functionally Redundant but Differentially Regulated in Candida glabrata: Implications for Echinocandin Resistance. Antimicrob. Agents Chemother. 2012, 56, 6304–6309. [Google Scholar] [CrossRef] [Green Version]
  125. Shields, R.K.; Kline, E.G.; Healey, K.R.; Kordalewska, M.; Perlin, D.S.; Nguyen, M.H.; Clancy, C.J. Spontaneous Mutational Frequency and FKS Mutation Rates Vary by Echinocandin Agent against Candida glabrata. Antimicrob. Agents Chemother. 2019, 63, e01692-18. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  126. Arendrup, M.C.; Perlin, D.S.; Jensen, R.H.; Howard, S.J.; Goodwin, J.; Hope, W. Differential In Vivo Activities of Anidulafungin, Caspofungin, and Micafungin against Candida glabrata Isolates with and without FKS Resistance Mutations. Antimicrob. Agents Chemother. 2012, 56, 2435–2442. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  127. Healey, K.R.; Perlin, D.S. Fungal Resistance to Echinocandins and the MDR Phenomenon in Candida glabrata. J. Fungi 2018, 4, 105. [Google Scholar] [CrossRef] [PubMed]
  128. Lesage, G.; Bussey, H. Cell Wall Assembly in Saccharomyces Cerevisiae. Microbiol. Mol. Biol. Rev. 2006, 70, 317–343. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  129. Levin, D.E. Cell Wall Integrity Signaling in Saccharomyces Cerevisiae. Microbiol. Mol. Biol. Rev. 2005, 69, 262–291. [Google Scholar] [CrossRef] [Green Version]
  130. 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] [Green Version]
  131. Lee, K.K.; MacCallum, D.M.; Jacobsen, M.D.; Walker, L.A.; Odds, F.C.; Gow, N.A.R.; 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]
  132. Wu, J.; Gan, C.; Li, J.; Liu, Y.; Chen, Z.; Zhang, Y.; Yi, G.; Sui, J.; Xu, J. Species Diversity and Antifungal Susceptibilities of Oral Yeasts from Patients with Head and Neck Cancer. Infect. Drug Resist. 2021, 14, 2279–2288. [Google Scholar] [CrossRef]
  133. De Alegría Puig, C.R.; del Sol García Merino, M.; De Malet Pintos-Fonseca, A.; Agüero Balbín, J. Characterization, Antifungal Susceptibility and Virulence of Candida parapsilosis Complex Isolates in a Tertiary Hospital in Cantabria, Northern Spain. Enfermedades Infecc. Y Microbiol. Clínica 2023, 41, 99–102. [Google Scholar] [CrossRef]
  134. Guo, J.; Zhang, M.; Qiao, D.; Shen, H.; Wang, L.; Wang, D.; Li, L.; Liu, Y.; Lu, H.; Wang, C.; et al. Prevalence and Antifungal Susceptibility of Candida parapsilosis Species Complex in Eastern China: A 15-Year Retrospective Study by ECIFIG. Front. Microbiol. 2021, 12, 644000. [Google Scholar] [CrossRef]
  135. Johnson, M.E.; Katiyar, S.K.; Edlind, T.D. New Fks Hot Spot for Acquired Echinocandin Resistance in Saccharomyces Cerevisiae and Its Contribution to Intrinsic Resistance of Scedosporium Species. Antimicrob. Agents Chemother. 2011, 55, 3774–3781. [Google Scholar] [CrossRef] [Green Version]
  136. Ong, V.; Hough, G.; Schlosser, M.; Bartizal, K.; Balkovec, J.M.; James, K.D.; Krishnan, B.R. Preclinical Evaluation of the Stability, Safety, and Efficacy of CD101, a Novel Echinocandin. Antimicrob. Agents Chemother. 2016, 60, 6872–6879. [Google Scholar] [CrossRef] [Green Version]
  137. Tóth, Z.; Forgács, L.; Locke, J.B.; Kardos, G.; Nagy, F.; Kovács, R.; Szekely, A.; Borman, A.M.; Majoros, L. In Vitro Activity of Rezafungin against Common and Rare Candida Species and Saccharomyces cerevisiae. J Antimicrob. Chemother. 2019, 74, 3505–3510. [Google Scholar] [CrossRef] [Green Version]
  138. Arendrup, M.C.; Meletiadis, J.; Zaragoza, O.; Jørgensen, K.M.; Marcos-Zambrano, L.J.; Kanioura, L.; Cuenca-Estrella, M.; Mouton, J.W.; Guinea, J. Multicentre Determination of Rezafungin (CD101) Susceptibility of Candida Species by the EUCAST Method. Clin. Microbiol. Infect. 2018, 24, 1200–1204. [Google Scholar] [CrossRef] [Green Version]
  139. Hoenigl, M.; Sprute, R.; Egger, M.; Arastehfar, A.; Cornely, O.A.; Krause, R.; Lass-Flörl, C.; Prattes, J.; Spec, A.; Thompson, G.R.; et al. The Antifungal Pipeline: Fosmanogepix, Ibrexafungerp, Olorofim, Opelconazole, and Rezafungin. Drugs 2021, 81, 1703–1729. [Google Scholar] [CrossRef]
  140. Siopi, M.; Papadopoulos, A.; Spiliopoulou, A.; Paliogianni, F.; Abou-Chakra, N.; Arendrup, M.C.; Damoulari, C.; Tsioulos, G.; Giannitsioti, E.; Frantzeskaki, F.; et al. Pan-Echinocandin Resistant C. parapsilosis Harboring an F652S Fks1 Alteration in a Patient with Prolonged Echinocandin Therapy. J. Fungi 2022, 8, 931. [Google Scholar] [CrossRef]
  141. Dismukes, W.E. Introduction to Antifungal Drugs. Clin. Infect. Dis. 2000, 30, 653–657. [Google Scholar] [CrossRef] [Green Version]
  142. Baginski, M.; Sternal, K.; Czub, J.; Borowski, E. Molecular Modelling of Membrane Activity of Amphotericin B, a Polyene Macrolide Antifungal Antibiotic. Acta Biochim. Pol. 2005, 52, 655–658. [Google Scholar] [CrossRef]
  143. Thompson, G.R.; Le, T.; Chindamporn, A.; Kauffman, C.A.; Alastruey-Izquierdo, A.; Ampel, N.M.; Andes, D.R.; Armstrong-James, D.; Ayanlowo, O.; Baddley, J.W.; et al. Global Guideline for the Diagnosis and Management of the Endemic Mycoses: An Initiative of the European Confederation of Medical Mycology in Cooperation with the International Society for Human and Animal Mycology. Lancet Infect. Dis. 2021, 21, e364–e374. [Google Scholar] [CrossRef]
  144. Chen, S.C.-A.; Perfect, J.; Colombo, A.L.; Cornely, O.A.; Groll, A.H.; Seidel, D.; Albus, K.; de Almedia, J.N.; Garcia-Effron, G.; Gilroy, N.; et al. Global Guideline for the Diagnosis and Management of Rare Yeast Infections: An Initiative of the ECMM in Cooperation with ISHAM and ASM. Lancet Infect. Dis. 2021, 21, e375–e386. [Google Scholar] [CrossRef]
  145. Cavassin, F.B.; Baú-Carneiro, J.L.; Vilas-Boas, R.R.; Queiroz-Telles, F. Sixty Years of Amphotericin B: An Overview of the Main Antifungal Agent Used to Treat Invasive Fungal Infections. Infect. Dis. Ther. 2021, 10, 115–147. [Google Scholar] [CrossRef] [PubMed]
  146. Ellis, D. Amphotericin B: Spectrum and Resistance. J. Antimicrob. Chemother. 2002, 49, 7–10. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  147. Stone, N.R.H.; Bicanic, T.; Salim, R.; Hope, W. Liposomal Amphotericin B (AmBisome®): A Review of the Pharmacokinetics, Pharmacodynamics, Clinical Experience and Future Directions. Drugs 2016, 76, 485–500. [Google Scholar] [CrossRef] [Green Version]
  148. White, T.C.; Marr, K.A.; Bowden, R.A. Clinical, Cellular, and Molecular Factors That Contribute to Antifungal Drug Resistance. Clin. Microbiol. Rev. 1998, 11, 382–402. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  149. Haynes, M.P.; Chong, P.L.; Buckley, H.R.; Pieringer, R.A. Fluorescence Studies on the Molecular Action of Amphotericin B on Susceptible and Resistant Fungal Cells. Biochemistry 1996, 35, 7983–7992. [Google Scholar] [CrossRef]
  150. Kelly, S.L.; Lamb, D.C.; Kelly, D.E.; Loeffler, J.; Einsele, H. Resistance to Fluconazole and Amphotericin in Candida albicans from AIDS Patients. Lancet 1996, 348, 1523–1524. [Google Scholar] [CrossRef]
  151. Bahmed, K.; Bonaly, R.; Coulon, J. Relation between Cell Wall Chitin Content and Susceptibility to Amphotericin B in Kluyveromyces, Candida and Schizosaccharomyces Species. Res. Microbiol. 2003, 154, 215–222. [Google Scholar] [CrossRef]
  152. Healey, K.R.; Zhao, Y.; Perez, W.B.; Lockhart, S.R.; Sobel, J.D.; Farmakiotis, D.; Kontoyiannis, D.P.; Sanglard, D.; Taj-Aldeen, S.J.; Alexander, B.D.; et al. Prevalent Mutator Genotype Identified in Fungal Pathogen Candida glabrata Promotes Multi-Drug Resistance. Nat. Commun. 2016, 7, 11128. [Google Scholar] [CrossRef] [Green Version]
  153. Yamin, D.; Akanmu, M.H.; Al Mutair, A.; Alhumaid, S.; Rabaan, A.A.; Hajissa, K. Global Prevalence of Antifungal-Resistant Candida parapsilosis: A Systematic Review and Meta-Analysis. Trop. Med. Infect. Dis. 2022, 7, 188. [Google Scholar] [CrossRef]
  154. Vermes, A.; Guchelaar, H.J.; Dankert, J. Flucytosine: A Review of Its Pharmacology, Clinical Indications, Pharmacokinetics, Toxicity and Drug Interactions. J. Antimicrob. Chemother. 2000, 46, 171–179. [Google Scholar] [CrossRef]
  155. Pfaller, M.A.; Messer, S.A.; Boyken, L.; Huynh, H.; Hollis, R.J.; Diekema, D.J. In Vitro Activities of 5-Fluorocytosine against 8,803 Clinical Isolates of Candida Spp.: Global Assessment of Primary Resistance Using National Committee for Clinical Laboratory Standards Susceptibility Testing Methods. Antimicrob. Agents Chemother. 2002, 46, 3518–3521. [Google Scholar] [CrossRef] [Green Version]
  156. Waldorf, A.R.; Polak, A. Mechanisms of Action of 5-Fluorocytosine. Antimicrob. Agents Chemother. 1983, 23, 79–85. [Google Scholar] [CrossRef] [Green Version]
  157. Diasio, R.B.; Bennett, J.E.; Myers, C.E. Mode of Action of 5-Fluorocytosine. Biochem. Pharmacol. 1978, 27, 703–707. [Google Scholar] [CrossRef]
  158. Papon, N.; Noël, T.; Florent, M.; Gibot-Leclerc, S.; Jean, D.; Chastin, C.; Villard, J.; Chapeland-Leclerc, F. Molecular Mechanism of Flucytosine Resistance in Candida lusitaniae: Contribution of the FCY2, FCY1, and FUR1 Genes to 5-Fluorouracil and Fluconazole Cross-Resistance. Antimicrob. Agents Chemother. 2007, 51, 369–371. [Google Scholar] [CrossRef] [Green Version]
  159. Fasoli, M.; Kerridge, D. Isolation and Characterization of Fluoropyrimidine-Resistant Mutants in Two Candida Species. Ann. N. Y. Acad. Sci. 1988, 544, 260–263. [Google Scholar] [CrossRef]
  160. Delma, F.Z.; Al-Hatmi, A.M.S.; Brüggemann, R.J.M.; Melchers, W.J.G.; de Hoog, S.; Verweij, P.E.; Buil, J.B. Molecular Mechanisms of 5-Fluorocytosine Resistance in Yeasts and Filamentous Fungi. J. Fungi 2021, 7, 909. [Google Scholar] [CrossRef]
  161. Francis, P.; Walsh, T.J. Evolving Role of Flucytosine in Immunocompromised Patients: New Insights into Safety, Pharmacokinetics, and Antifungal Therapy. Clin. Infect. Dis. 1992, 15, 1003–1018. [Google Scholar] [CrossRef] [Green Version]
  162. Polak, A. 5-Fluorocytosine--Current Status with Special References to Mode of Action and Drug Resistance. Contrib. Microbiol. Immunol. 1977, 4, 158–167. [Google Scholar]
  163. Hoeprich, P.D.; Ingraham, J.L.; Kleker, E.; Winship, M.J. Development of Resistance to 5-Fluorocytosine in Candida parapsilosis during Therapy. J. Infect. Dis. 1974, 130, 112–118. [Google Scholar] [CrossRef]
  164. Sun, L.-L.; Li, H.; Yan, T.-H.; Cao, Y.-B.; Jiang, Y.-Y.; Yang, F. Aneuploidy Enables Cross-Tolerance to Unrelated Antifungal Drugs in Candida parapsilosis. Front. Microbiol. 2023, 14, 1137083. [Google Scholar] [CrossRef]
  165. Pavelka, N.; Rancati, G.; Zhu, J.; Bradford, W.D.; Saraf, A.; Florens, L.; Sanderson, B.W.; Hattem, G.L.; Li, R. Aneuploidy Confers Quantitative Proteome Changes and Phenotypic Variation in Budding Yeast. Nature 2010, 468, 321–325. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  166. Fridman, O.; Goldberg, A.; Ronin, I.; Shoresh, N.; Balaban, N.Q. Optimization of Lag Time Underlies Antibiotic Tolerance in Evolved Bacterial Populations. Nature 2014, 513, 418–421. [Google Scholar] [CrossRef] [PubMed]
  167. Falagas, M.E.; Makris, G.C.; Dimopoulos, G.; Matthaiou, D.K. Heteroresistance: A Concern of Increasing Clinical Significance? Clin. Microbiol. Infect. 2008, 14, 101–104. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  168. Marr, K.A.; Rustad, T.R.; Rex, J.H.; White, T.C. The Trailing End Point Phenotype in Antifungal Susceptibility Testing Is PH Dependent. Antimicrob. Agents Chemother. 1999, 43, 1383–1386. [Google Scholar] [CrossRef] [Green Version]
  169. Arastehfar, A.; Lass-Flörl, C.; Garcia-Rubio, R.; Daneshnia, F.; Ilkit, M.; Boekhout, T.; Gabaldon, T.; Perlin, D.S. The Quiet and Underappreciated Rise of Drug-Resistant Invasive Fungal Pathogens. J. Fungi 2020, 6, 138. [Google Scholar] [CrossRef]
  170. Lewis, K. Multidrug Tolerance of Biofilms and Persister Cells. Curr. Top. Microbiol. Immunol. 2008, 322, 107–131. [Google Scholar] [CrossRef]
  171. Lewis, K. Persister Cells, Dormancy and Infectious Disease. Nat. Rev. Microbiol. 2007, 5, 48–56. [Google Scholar] [CrossRef]
  172. Yang, F.; Lu, H.; Wu, H.; Fang, T.; Berman, J.; Jiang, Y.-Y. Aneuploidy Underlies Tolerance and Cross-Tolerance to Drugs in Candida parapsilosis. Microbiol. Spectr. 2021, 9, e0050821. [Google Scholar] [CrossRef]
  173. Hickman, M.A.; Zeng, G.; Forche, A.; Hirakawa, M.P.; Abbey, D.; Harrison, B.D.; Wang, Y.-M.; Su, C.; Bennett, R.J.; Wang, Y.; et al. The ‘Obligate Diploid’ Candida albicans Forms Mating-Competent Haploids. Nature 2013, 494, 55–59. [Google Scholar] [CrossRef] [Green Version]
  174. 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, e00071-17. [Google Scholar] [CrossRef] [Green Version]
  175. Windels, E.M.; Michiels, J.E.; Fauvart, M.; Wenseleers, T.; Van den Bergh, B.; Michiels, J. Bacterial Persistence Promotes the Evolution of Antibiotic Resistance by Increasing Survival and Mutation Rates. ISME J. 2019, 13, 1239–1251. [Google Scholar] [CrossRef]
  176. Levinson, T.; Dahan, A.; Novikov, A.; Paran, Y.; Berman, J.; Ben-Ami, R. Impact of Tolerance to Fluconazole on Treatment Response in Candida Albicans Bloodstream Infection. Mycoses 2021, 64, 78–85. [Google Scholar] [CrossRef]
  177. Astvad, K.M.T.; Sanglard, D.; Delarze, E.; Hare, R.K.; Arendrup, M.C. Implications of the EUCAST Trailing Phenomenon in Candida Tropicalis for the In Vivo Susceptibility in Invertebrate and Murine Models. Antimicrob. Agents Chemother. 2018, 62, e01624-18. [Google Scholar] [CrossRef] [Green Version]
  178. Nunes, A.P.F.; Teixeira, L.M.; Iorio, N.L.P.; Bastos, C.C.R.; de Sousa Fonseca, L.; Souto-Padrón, T.; dos Santos, K.R.N. Heterogeneous Resistance to Vancomycin in Staphylococcus Epidermidis, Staphylococcus Haemolyticus and Staphylococcus Warneri Clinical Strains: Characterisation of Glycopeptide Susceptibility Profiles and Cell Wall Thickening. Int. J. Antimicrob. Agents 2006, 27, 307–315. [Google Scholar] [CrossRef]
  179. Pournaras, S.; Ikonomidis, A.; Markogiannakis, A.; Maniatis, A.N.; Tsakris, A. Heteroresistance to Carbapenems in Acinetobacter Baumannii. J. Antimicrob. Chemother. 2005, 55, 1055–1056. [Google Scholar] [CrossRef] [Green Version]
  180. Rinder, H.; Mieskes, K.T.; Löscher, T. Heteroresistance in Mycobacterium Tuberculosis. Int. J. Tuberc. Lung Dis. 2001, 5, 339–345. [Google Scholar]
  181. Yamazumi, T.; Pfaller, M.A.; Messer, S.A.; Houston, A.K.; Boyken, L.; Hollis, R.J.; Furuta, I.; Jones, R.N. Characterization of Heteroresistance to Fluconazole among Clinical Isolates of Cryptococcus Neoformans. J. Clin. Microbiol. 2003, 41, 267–272. [Google Scholar] [CrossRef] [Green Version]
  182. Ben-Ami, R.; Zimmerman, O.; Finn, T.; Amit, S.; Novikov, A.; Wertheimer, N.; Lurie-Weinberger, M.; Berman, J. Heteroresistance to Fluconazole Is a Continuously Distributed Phenotype among Candida glabrata Clinical Strains Associated with In Vivo Persistence. mBio 2016, 7, e00655-16. [Google Scholar] [CrossRef] [Green Version]
  183. Sionov, E.; Chang, Y.C.; Garraffo, H.M.; Kwon-Chung, K.J. Heteroresistance to Fluconazole in Cryptococcus Neoformans Is Intrinsic and Associated with Virulence. Antimicrob. Agents Chemother. 2009, 53, 2804–2815. [Google Scholar] [CrossRef] [Green Version]
  184. Stone, N.R.; Rhodes, J.; Fisher, M.C.; Mfinanga, S.; Kivuyo, S.; Rugemalila, J.; Segal, E.S.; Needleman, L.; Molloy, S.F.; Kwon-Chung, J.; et al. Dynamic Ploidy Changes Drive Fluconazole Resistance in Human Cryptococcal Meningitis. J. Clin. Investig. 2019, 129, 999–1014. [Google Scholar] [CrossRef] [Green Version]
  185. Zhai, B.; Liao, C.; Jaggavarapu, S.; Rolling, T.; Bergin, S.A.; Gjonbalaj, M.; Miranda, E.; Babady, N.E.; Butler, G.; Taur, Y.; et al. Echinocandin Heteroresistance Causes Prophylaxis Failure and Facilitates Breakthrough Candida parapsilosis Infection. medRxiv 2022. [Google Scholar] [CrossRef]
  186. Wootton, M.; Howe, R.A.; Hillman, R.; Walsh, T.R.; Bennett, P.M.; MacGowan, A.P. A Modified Population Analysis Profile (PAP) Method to Detect Hetero-Resistance to Vancomycin in Staphylococcus Aureus in a UK Hospital. J. Antimicrob. Chemother. 2001, 47, 399–403. [Google Scholar] [CrossRef] [PubMed]
  187. Ceballos-Garzon, A.; Peñuela, A.; Valderrama-Beltrán, S.; Vargas-Casanova, Y.; Ariza, B.; Parra-Giraldo, C.M. Emergence and Circulation of Azole-Resistant C. albicans, C. auris and C. parapsilosis Bloodstream Isolates Carrying Y132F, K143R or T220L Erg11p Substitutions in Colombia. Front. Cell. Infect. Microbiol. 2023, 13, 1136217. [Google Scholar] [CrossRef] [PubMed]
  188. Arastehfar, A.; Hilmioğlu-Polat, S.; Daneshnia, F.; Pan, W.; Hafez, A.; Fang, W.; Liao, W.; Şahbudak-Bal, Z.; Metin, D.Y.; Júnior, J.N.d.A.; et al. Clonal Candidemia Outbreak by Candida parapsilosis Carrying Y132F in Turkey: Evolution of a Persisting Challenge. Front. Cell. Infect. Microbiol. 2021, 11, 676177. [Google Scholar] [CrossRef] [PubMed]
  189. Thomaz, D.Y.; Del Negro, G.M.B.; Ribeiro, L.B.; da Silva, M.; Carvalho, G.O.M.H.; Camargo, C.H.; de Almeida, J.N.; Motta, A.L.; Siciliano, R.F.; Sejas, O.N.E.; et al. A Brazilian Inter-Hospital Candidemia Outbreak Caused by Fluconazole-Resistant Candida parapsilosis in the COVID-19 Era. J. Fungi 2022, 8, 100. [Google Scholar] [CrossRef]
  190. Thomaz, D.Y.; de Almeida, J.N., Jr.; Lima, G.M.E.; de Oliveira Nunes, M.; Camargo, C.H.; de Carvalho Grenfell, R.; Benard, G.; Del Negro, G.M.B. An Azole-Resistant Candida parapsilosis Outbreak: Clonal Persistence in the Intensive Care Unit of a Brazilian Teaching Hospital. Front. Microbiol. 2018, 9, 2997. [Google Scholar] [CrossRef]
  191. Mamali, V.; Siopi, M.; Charpantidis, S.; Samonis, G.; Tsakris, A.; Vrioni, G.; on Behalf of the Candi-Candi Network. Increasing Incidence and Shifting Epidemiology of Candidemia in Greece: Results from the First Nationwide 10-Year Survey. J. Fungi 2022, 8, 116. [Google Scholar] [CrossRef]
  192. Presente, S.; Bonnal, C.; Normand, A.-C.; Gaudonnet, Y.; Fekkar, A.; Timsit, J.-F.; Kernéis, S. Hospital Clonal Outbreak of Fluconazole-Resistant Candida parapsilosis Harboring the Y132F ERG11p Substitution in a French Intensive Care Unit. Antimicrob. Agents Chemother. 2023, 67, e0113022. [Google Scholar] [CrossRef]
  193. Díaz-García, J.; Gómez, A.; Machado, M.; Alcalá, L.; Reigadas, E.; Sánchez-Carrillo, C.; Pérez-Ayala, A.; Gómez-García De La Pedrosa, E.; González-Romo, F.; Cuétara, M.S.; et al. Blood and Intra-Abdominal Candida spp. from a Multicentre Study Conducted in Madrid Using EUCAST: Emergence of Fluconazole Resistance in Candida parapsilosis, Low Echinocandin Resistance and Absence of Candida auris. J. Antimicrob. Chemother. 2022, 77, 3102–3109. [Google Scholar] [CrossRef]
  194. Alcoceba, E.; Gómez, A.; Lara-Esbrí, P.; Oliver, A.; Beltrán, A.F.; Ayestarán, I.; Muñoz, P.; Escribano, P.; Guinea, J. Fluconazole-Resistant Candida parapsilosis Clonally Related Genotypes: First Report Proving the Presence of Endemic Isolates Harbouring the Y132F ERG11 Gene Substitution in Spain. Clin. Microbiol. Infect. 2022, 28, 1113–1119. [Google Scholar] [CrossRef]
  195. Fekkar, A.; Blaize, M.; Bouglé, A.; Normand, A.-C.; Raoelina, A.; Kornblum, D.; Kamus, L.; Piarroux, R.; Imbert, S. Hospital Outbreak of Fluconazole-Resistant Candida parapsilosis: Arguments for Clonal Transmission and Long-Term Persistence. Antimicrob. Agents Chemother. 2021, 65, e02036-20. [Google Scholar] [CrossRef]
  196. Hernández-Castro, R.; Arroyo-Escalante, S.; Carrillo-Casas, E.M.; Moncada-Barrón, D.; Alvarez-Verona, E.; Hernández-Delgado, L.; Torres-Narváez, P.; Lavalle-Villalobos, A. Outbreak of Candida parapsilosis in a Neonatal Intensive Care Unit: A Health Care Workers Source. Eur. J. Pediatr. 2010, 169, 783–787. [Google Scholar] [CrossRef]
  197. Plouffe, J.F.; Brown, D.G.; Silva, J.; Eck, T.; Stricof, R.L.; Fekety, F.R. Nosocomial Outbreak of Candida parapsilosis Fungemia Related to Intravenous Infusions. Arch. Intern. Med. 1977, 137, 1686–1689. [Google Scholar] [CrossRef]
  198. Chiotos, K.; Vendetti, N.; Zaoutis, T.E.; Baddley, J.; Ostrosky-Zeichner, L.; Pappas, P.; Fisher, B.T. Comparative Effectiveness of Echinocandins versus Fluconazole Therapy for the Treatment of Adult Candidaemia Due to Candida parapsilosis: A Retrospective Observational Cohort Study of the Mycoses Study Group (MSG-12). J. Antimicrob. Chemother. 2016, 71, 3536–3539. [Google Scholar] [CrossRef] [Green Version]
  199. Meletiadis, J.; Curfs-Breuker, I.; Meis, J.F.; Mouton, J.W. In Vitro Antifungal Susceptibility Testing of Candida Isolates with the EUCAST Methodology, a New Method for ECOFF Determination. Antimicrob. Agents Chemother. 2017, 61, 10–1128. [Google Scholar] [CrossRef] [Green Version]
  200. Cantón, E.; Pemán, J.; Quindós, G.; Eraso, E.; Miranda-Zapico, I.; Álvarez, M.; Merino, P.; Campos-Herrero, I.; Marco, F.; de la Pedrosa, E.G.G.; et al. Prospective Multicenter Study of the Epidemiology, Molecular Identification, and Antifungal Susceptibility of Candida parapsilosis, Candida orthopsilosis, and Candida metapsilosis Isolated from Patients with Candidemia. Antimicrob. Agents Chemother. 2011, 55, 5590–5596. [Google Scholar] [CrossRef] [Green Version]
  201. Binder, U.; Arastehfar, A.; Schnegg, L.; Hörtnagl, C.; Hilmioğlu-Polat, S.; Perlin, D.S.; Lass-Flörl, C. Efficacy of LAMB against Emerging Azole- and Multidrug-Resistant Candida parapsilosis Isolates in the Galleria Mellonella Model. J. Fungi 2020, 6, 377. [Google Scholar] [CrossRef]
  202. Ostrosky-Zeichner, L.; Rex, J.H.; Pappas, P.G.; Hamill, R.J.; Larsen, R.A.; Horowitz, H.W.; Powderly, W.G.; Hyslop, N.; Kauffman, C.A.; Cleary, J.; et al. Antifungal Susceptibility Survey of 2000 Bloodstream Candida Isolates in the United States. Antimicrob. Agents Chemother. 2003, 47, 3149–3154. [Google Scholar] [CrossRef] [Green Version]
  203. Quindós, G.; Ruesga, M.T.; Martín-Mazuelos, E.; Salesa, R.; Alonso-Vargas, R.; Carrillo-Muñoz, A.J.; Brena, S.; San Millán, R.; Pontón, J. In-Vitro Activity of 5-Fluorocytosine against 1,021 Spanish Clinical Isolates of Candida and Other Medically Important Yeasts. Rev. Iberoam. Micol. 2004, 21, 63–69. [Google Scholar] [PubMed]
Figure 1. Overview of azole-resistance mechanisms in C. parapsilosis sensu stricto.
Figure 1. Overview of azole-resistance mechanisms in C. parapsilosis sensu stricto.
Jof 09 00798 g001
Table 1. Azole and echinocandin tolerance and resistance mechanisms were chronologically reported.
Table 1. Azole and echinocandin tolerance and resistance mechanisms were chronologically reported.
Azole-Resistant C. parapsilosis sensu stricto
Mechanism of ResistanceAmino Acid ChangeGeneEffect on Antifungal DrugsReference
Gain-of-function mutationG583RMRR1FLU-R, VOR-R[93,94]
Gain-of-function mutationK873NMRR1FLU-R, VOR-R[93,94]
Upregulation-UPC2FLU-R, VOR-R, POS-R[93] ¥
Upregulation-NTD80FLU-R, VOR-R, POS-R[93] ¥
Target changeY132FERG11FLU-R, VOR-R[95]
UpregulationL986PMRR1FLU-R, VOR-S/I[96]
UpregulationG650ETAC1FLU-R, VOR-R[97]
UpregulationL978WTAC1FLU-R, VOR-R[97]
Loss of functionR135IERG3FLU-R, VOR-R, POS-R[98]
Loss of functionG111RERG3FLU-R, VOR-R, POS-R[99]
UpregulationP45HUPC2FLU-R, VOR-S/I[4]
UpregulationQ371HUPC2FLU-R, VOR-I[4]
Target changeK143RERG11FLU-R[100]
UpregulationL518FTAC1FLU-R, VOR-R[100]
Target changeG458SERG11FLU-R, VOR-R[101]
UpregulationA854VMRR1FLU-R[102]
UpregulationR479KMRR1FLU-R[102]
UpregulationI283RMRR1FLU-R[102]
Gain-of-function mutationG604RMRR1FLU-R, VOR-R[103]
Echinocandin-Tolerant/Resistant C. parapsilosis sensu stricto
Mechanism of ResistanceAminoacid ChangeGeneEffect on Antifungal DrugsReference
Target change *P660AHS1-FKS1ANF, CS, MYC reduced susceptibility[104]
Target changeV595Inon-HS1-FKS1 °ECT §, CS-I[105]
Target changeF1386Snon-HS2-FKS1 #ECT §, ANF-R, MYC-I[105]
Loss of functionG111RERG3ANF-I, MYC-I/R[99]
Target changeR658GHS1-FKS1MYC-R[106]
Target changeE1393Gnon-HS2-FKS1 #ECT §[100]
Target changeA1422Gnon-HS2-FKS1 #ECT §[107]
Target changeM1328Inon-HS2-FKS1 #ECT §[107]
Target changeS745Lnon-HS1-FKS1 °ECT §[107]
Target changeS656PHS1-FKS1ANF-R, MYC-R, CS-R[108]
¥ in this study only gene expression levels were evaluated, no mutation was reported. * constitutively present; § ECT = echinocandin tolerance; ° outside the Hot-spot region 1 of FKS1; # outside the Hot-spot region 2 of FKS1; FLU = fluconazole; VOR = voriconazole; POS = posaconazole; MYC = micafungin; ANF = anidulafungin; CS = caspofungin.
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Franconi, I.; Rizzato, C.; Poma, N.; Tavanti, A.; Lupetti, A. Candida parapsilosis sensu stricto Antifungal Resistance Mechanisms and Associated Epidemiology. J. Fungi 2023, 9, 798. https://doi.org/10.3390/jof9080798

AMA Style

Franconi I, Rizzato C, Poma N, Tavanti A, Lupetti A. Candida parapsilosis sensu stricto Antifungal Resistance Mechanisms and Associated Epidemiology. Journal of Fungi. 2023; 9(8):798. https://doi.org/10.3390/jof9080798

Chicago/Turabian Style

Franconi, Iacopo, Cosmeri Rizzato, Noemi Poma, Arianna Tavanti, and Antonella Lupetti. 2023. "Candida parapsilosis sensu stricto Antifungal Resistance Mechanisms and Associated Epidemiology" Journal of Fungi 9, no. 8: 798. https://doi.org/10.3390/jof9080798

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