Download as pdf or txt
Download as pdf or txt
You are on page 1of 13

Fungal Genetics and Biology 132 (2019) 103254

Contents lists available at ScienceDirect

Fungal Genetics and Biology


journal homepage: www.elsevier.com/locate/yfgbi

Review

Facilitators of adaptation and antifungal resistance mechanisms in clinically T


relevant fungi
Margriet W.J. Hokkena,b, , B.J. Zwaanc, W.J.G. Melchersa,b, P.E. Verweija,b

a
Department of Medical Microbiology, Radboud University Medical Center, Radboud Institute for Molecular Life Sciences, Geert Grooteplein Zuid 10, 6525GA Nijmegen,
the Netherlands
b
Center of Expertise in Mycology Radboudumc/CWZ, Radboud University Medical Center, Radboud Institute for Molecular Life Sciences, Geert Grooteplein Zuid 10,
6525GA Nijmegen, the Netherlands
c
Department of Plant Sciences, Laboratory of Genetics, Wageningen University, Droevendaalsesteeg 1, 6708PB Wageningen, the Netherlands

ARTICLE INFO ABSTRACT

Keywords: Opportunistic fungal pathogens can cause a diverse range of diseases in humans. The increasing rate of fungal
Aspergillus spp. infections caused by strains that are resistant to commonly used antifungals results in difficulty to treat diseases,
Candida spp. with accompanying high mortality rates. Existing and newly emerging molecular resistance mechanisms rapidly
Cryptococcus spp. spread in fungal populations and need to be monitored. Fungi exhibit a diversity of mechanisms to maintain
Antifungal resistance mechanisms
physiological resilience and create genetic variation; processes which eventually lead to the selection and spread
Antifungal compounds
Adaptation
of resistant fungal pathogens. To prevent and anticipate this dispersion, the role of evolutionary factors that
Mutation rate drive fungal adaptation should be investigated.
Reproduction In this review, we provide an overview of resistance mechanisms against commonly used antifungal com-
pounds in the clinic and for which fungal resistance has been reported. Furthermore, we aim to summarize and
elucidate potent generators of genetic variability across the fungal kingdom that aid adaptation to stressful
environments. This knowledge can lead to recognizing potential niches that facilitate fast resistance develop-
ment and can provide leads for new management strategies to battle the emerging resistant populations in the
clinic and the environment.

1. Introduction Additionally, organ/tissue transplant patients are at high-risk for


fungal infections (Khan et al., 2015; Shoham and Marr, 2014). Fur-
1.1. The rise of fungal diseases and resistance as a clinical challenge thermore, the causative agents of endemic mycoses cause infections in
both healthy and immunocompromised humans and occur in specific
Fungi are widespread and may cause a wide spectrum of diseases in ecological niches (Malcolm and Chin-Hong, 2013; Salzer et al., 2018).
humans (Chotirmall et al., 2014; Latgé, 2001; Schmiedel and Zimmerli, Factors such as prolonged lung colonization and frequent exposure
2016), with most infections occurring by the opportunistic pathogens have proved to greatly influence the severity of these fungal infections.
Candida albicans, Aspergillus fumigatus, and Cryptococcus neoformans. Since the introduction of the first antifungal compounds in the late
Typically, in healthy humans, fungal infections are efficiently eradi- 1960s, resistant fungal isolates have been reported in many countries
cated by the immune system. Therefore, fungal infections are more worldwide (Buil et al., 2019; Hagiwara, 2018; Snelders et al., 2008;
common in patients with immunodeficiencies, or patients who have Sorrell and Chen, 2007; Webb et al., 2018). As a result, the incidence of
underlying chronic illnesses such as asthma, leukemia, or cystic fibrosis invasive infections caused by resistant fungi has increased over the past
(Bhatt et al., 2011; Denning et al., 2014). Frequent exposure to airborne decades (Lestrade et al., 2019; Wiederhold, 2017) and antifungal re-
fungi can additionally lead to inflammatory responses rather than in- sistance increasingly becomes a concern for clinicians.
vasive infections, such as asthma caused by Cladosporium spp. and Al- Although treating invasive fungal infections with antifungals is es-
ternaria spp., or inflamed alveoli caused by the pathogen Pneumocystis sential to increase survival, treatment options remain limited. This is
jiroveci (Cordonnier et al., 2016; Denning et al., 2014; Hoving and Kolls, because most compounds that are effective against fungi are toxic to
2017). mammalian cells due to their shared cellular eukaryotic structures. The


Corresponding author.
E-mail address: margriet.hokken@radboudumc.nl (M.W.J. Hokken).

https://doi.org/10.1016/j.fgb.2019.103254
Received 12 June 2019; Received in revised form 16 July 2019; Accepted 16 July 2019
Available online 19 July 2019
1087-1845/ © 2019 The Authors. Published by Elsevier Inc. This is an open access article under the CC BY-NC-ND license
(http://creativecommons.org/licenses/BY-NC-ND/4.0/).
M.W.J. Hokken, et al. Fungal Genetics and Biology 132 (2019) 103254

Table 1
The major antifungal classes and their cellular targets.
Chemical class Compounds Target pathway Molecular target

Azoles Fluconazole, itraconazole, voriconazole, posaconazole, isavuconazole, Ergosterol biosynthesis 14α-sterol demethylase
ketoconazole
Echinocandins Anidulafungin, caspofungin, micafungin 1,3 β-glucan biosynthesis 1,3 β-glucan synthase complex catalytic
subunit
Polyenes Amphotericin B, nyastatin, natamycin Ergosterol biosynthesis Ergosterol in fungal membranes
Allylamines Terbinafine, naftifine Ergosterol biosynthesis Squalene epoxidase
Pyrimidine analogs Flucytosine Pyrimidine salvage pathway DNA/RNA molecules

introduction of the triazoles and the echinocandins presented more maintenance therapy for cryptococcal meningitis in HIV patients and
treatment options in the 1990s, but these advances were readily fol- non-HIV patients (Zhao et al., 2018). Second-generation triazoles in-
lowed by the emergence of triazole and echinocandin resistant fungal clude voriconazole and posaconazole, which are both used for the
pathogens (Denning et al., 1997; Ksiezopolska and Gabaldón, 2018; treatment and prophylaxis of invasive aspergillosis (IA) respectively
Moudgal et al., 2004; Perfect and Cox, 1999; Verweij et al., 1998). (Lat and Thompson, 2011; Sandherr and Maschmeyer, 2011). Both
Present-day antifungal treatment constitutes five main classes of anti- drugs are currently used for prophylaxis, although posaconazole treat-
fungals, which comprise the polyenes, imidazoles and trizoles, allyla- ment resulted in less side effects than voriconazole according to a re-
mines, echinocandins and the compound 5-flucytosine (Chen and cent study (Tang et al., 2018). In 2015, isavuconazole was approved
Sorrell, 2007; Ghannoum and Rice, 1999). For an overview of common and showed efficacy similar to voriconazole against IA and has been
antifungal compounds used in the clinic and their cellular targets, see licensed for treatment of mucormycosis (Jenks et al., 2018).
Table 1. Azole resistance was first reported in the clinic in 1986 in C. albicans
(Smith et al., 1986), and has now been found in many pathogenic
1.2. The adaptive potential of fungi species including A. fumigatus, Histoplasma capsulatum and C. neofor-
mans (Denning et al., 1997, Lamb et al., 1999).
Fungi can quickly adapt to new and challenging environments in
nature, equipping them with intrinsic defenses to overcome cellular 2.1.2. Echinocandins
stress induced by antifungal compounds (; Berkow and Lockhart, 2018; The echinocandins micafungin, caspofungin, and anidulafungin
Ksiezopolska and Gabaldón, 2018; Pré et al., 2018). Adaptation in this target fungal cell wall synthesis by inhibiting FKS1, a glucan synthase
sense is described as overcoming environmental challenges by gaining which is essential in the biosynthesis of 1,3 ß-glucans in the fungal cell
beneficial mutations or adjusting cellular physiology. How quickly wall (Walker et al., 2010). They prove effective against invasive can-
these adaptations arise and spread after natural selection, depends on didiasis, exert moderate activity against Aspergillus spp., but are in-
forces that drive evolution itself such as (a)sexual reproduction, ploidy effective in infections caused by Fusarium spp., Cryptococcus spp. or
changes, and genetic stability. These processes greatly influence the mucorales (Eschenauer et al., 2007; Maligie and Selitrennikoff, 2005).
ability to develop and pass on resistance mechanisms, allowing re- Echinocandin resistance has been reported extensively in many Candida
sistance to spread through the population, regionally and globally. spp., (Jensen et al., 2013; Park et al., 2005; Perlin, 2007), and has also
The variety of molecular mechanisms that were discovered over the been found in A. fumigatus (Jiménez-Ortigosa et al., 2017). Prevalence
last decades have not only given us insight in the many ways in which of echinocandin resistant isolates was observed to be < 3% in 2010 in
fungal pathogens can exhibit intrinsic or acquired resistance (Sanglard, various Candida isolates (Castanheira et al., 2010), but a more recent
2019), but has given us insights in the origin of drug-resistant fungal study observed a prevalence of > 12% in C. glabrata clinical isolates
isolates as well. (Alexander et al., 2013).
In this review, we aim to give an overview of the resistance me-
chanisms currently discovered in fungal pathogens. In addition, the 2.1.3. Polyenes
factors that drive evolution and spread of resistant fungal isolates are The polyene compounds, which include amphotericin B and nyas-
discussed. We hope this contributes to a greater understanding towards tatin, bind to ergosterol specifically, which causes disruption of the
resistant fungal communities, and ultimately to the prevention and membrane integrity with subsequent cellular leakage as a consequence.
improved treatment of severe fungal infections. Furthermore, amphotericin B induces oxidative damage (Mesa-Arango
et al., 2012). Amphotericin B is a fungicidal against many different
2. Antifungal compounds fungi, but due to renal failure as a side-effect it is a second-line treat-
ment (Sawaya et al., 1995). Resistance to amphotericin B is not com-
2.1. Chemical classes and their mode of action monly reported, although a recent study observed increasing resistance
rates in a hospital in Brazil, where 27% of the A. fumigatus isolates had
The antifungal compounds that are currently used in clinical elevated MICs (> 2 mg/L) for amphotericin B (Luzia et al., 2018).
therapy inhibit cell wall biosynthesis, interfere with sterol metabolism, Amphotericin B resistance may possibly arise when genes in the er-
or interfere with pyrimidine metabolism (Perfect, 2018). They inhibit gosterol biosynthesis pathway carry mutations. This eventually results
mechanisms uniquely present in fungi or target proteins that share little in strains with little to no ergosterol in their membranes, and sub-
homology with their mammalian counterparts. sequent amphotericin B resistance as the target molecule is absent from
the cells. This mechanism has been described for C. tropicalis, C. lusi-
2.1.1. Azoles taniae, C. albicans, A. terreus and C. neoformans (Forastiero et al., 2013;
The azole class compounds disrupt ergosterol biosynthesis by in- Kelly et al., 1997, 1995; Walsh et al., 2003; Young et al., 2003).
hibiting the sterol 14α-demethylase, which is a cytochrome p450 en-
zyme (Bossche et al., 1986; Liu et al., 2016). The first-generation tria- 2.1.4. Allylamines
zoles fluconazole and itraconazole are commonly used to treat Candida The allylamines inhibit the squalene epoxidase enzyme, which is
spp. infections and nail infections respectively (Martin, 1999; Nelson essential in the ergosterol biosynthesis pathway. They are commonly
et al., 1994; Richardson, 1997). Fluconazole is often used as a used to treat skin and nail infections (Mukherjee et al., 2003). A

2
M.W.J. Hokken, et al. Fungal Genetics and Biology 132 (2019) 103254

1. Biofilm formation 2. Loss of mitochondrial function 3. Stress signaling Increased


Mdm31
caspofungin/fluconazole tolerance
Azole resistance Tom70
Calcineurin
drug resistance increased activation
IMM OMM

Complex I Hsp90 Regulates HOG signaling


Hyphal development

IMM
Cell wall integrity

hsp90
transcription

4. Mutations in cell wall biosynthesis genes 5. Inhibited pyrimidine salvage pathway


1

2 3 Fcy1 Fur1
5FC

4
7+8 5FU
Fks1 Fks2 5 5FC resistance
Rho1 fcyB
6 Inhibition of protein synthesis
regulates disturbs
Inhibition of DNA synthesis
β-(1,3)-GS complex

6. Increased drug efflux 7. Upregulation and mutation of target genes 8. Lipid metabolism bypass
in ergosterol biosynthesis Azole resistance
Bind to Polyene resistance
Cdr1b

inactivation

incorporation
Cyp51A Erg1

Membrane
Cyp51A Erg3
AtrR
cdr1b drug resistance TR TR cyp51A Azole resistance 14α-methyl sterols Prevention of
Transcription Transcription Allylamine resistance toxic sterol production

Mutation 5FC Allylamines Azoles Echinocandins Polyenes Ergosterol

Fig. 1. Antifungal resistance mechanisms found in fungi. 1: Biofilm formation reduces penetration of antifungal compounds into the population. 2: Loss of mi-
tochondrial complex I was shown to result in azole resistant isolates. 3: Enhanced production of Hsp90 induces stress-signaling pathways and leads to enhanced
tolerance for azoles and echinocandins. 4: Mutations in 1,3 β-glucan synthase complex catalytic subunits prevent interference in cell wall biosynthesis by echino-
candins. 5: Mutations in the cytosine permease FcyB prevent flucytosine from entering the cell. Mutations in the cytosine deaminase Fcy1 or the uracil phos-
phoribosyltransferase Fur1 prevent conversion of flucytosine to the toxic 5-fluorouracil. 6: Increased activity of drug-efflux transporters result in less intracellular
drug accumulation. 7: Mutations in the squalene expoxidase erg1 or the 14α-sterol demethylase Cyp51A prevent activity of allylamines and azoles, respectively.
Upregulation of the 14α-sterol demethylase has evolved together with target site alterations and ensures enzyme availability. 8: Absence of the sterol desaturase Erg3
prevents the production of toxic sterols if Cyp51A is inhibited by azole compounds. Non-toxic sterol intermediates are incorporated in the membrane instead.

relatively high resistance rate of 32% was observed in 63 Trichophyton 2.2. In vitro resistance development to low dosages of antifungal compounds
interdigitale isolates in India (Singh et al., 2018), but is as yet unclear
whether resistance is rising elsewhere. Resistant isolates were shown to Resistance development is shown to result more likely from selec-
have single amino acid changes in the target gene, erg1. Another study tion at sublethal concentrations of the drug, then from exposure to
suggested that terbinafine resistance in A. nidulans resulted from above-MIC concentrations in A. fumigatus (Escribano et al., 2012; Zhang
overexpression of salA, a salicylate 1-monooxygenase which facilitates et al., 2017b). Sublethal concentrations supposedly permit the fungus to
napthalene degradation, which may use terbinafine as a substrate as it still grow and reproduce, among others through (adaptive) phenotypic
contains a napthalene ring (Graminha et al., 2004). plasticity, to allow natural selection to act. Various studies have de-
scribed antifungal resistance induction in a laboratory setting at various
drug concentrations, to investigate how rapid resistance can be induced
2.1.5. Pyrimidine analogs and which mechanisms will arise in the exposed isolates (Cowen et al.,
5-Flucytosine (5FC) is a pyrimidine analog used in combination 2000; Huang et al., 2011).
therapy, but studies have shown that many fungal pathogens are able to De novo induction of voriconazole resistance has been shown to be
rapidly develop 5FC resistance (Charlier et al., 2016). In the cell, 5FC is possible already after five weeks culturing which started with sublethal
converted to toxic 5-fluorouracil, which interferes with DNA and RNA concentrations at 0.125 mg L−1, although no cyp51A/B mutations were
structures (Perfect, 2018). Resistant isolates that were found include C. found in the resistant A. fumigatus isolates (Händel et al., 2015; Verweij
glabrata, A. fumigatus, and C. gatti (Edlind and Katiyar, 2010; Gsaller et al., 2009). In another study, cross-resistant A. fumigatus isolates were
et al., 2018; Vu et al., 2018). obtained already after three weeks of culturing in the presence of the
fungicide difenoconazole at a concentration of 1 mg L−1 (Zhang et al.,
2017b).

3
M.W.J. Hokken, et al. Fungal Genetics and Biology 132 (2019) 103254

In T. rubrum, azole resistance was induced in vitro after one year of 3.2. Structural target site alterations
passaging isolates at sub MIC concentrations (Hryncewicz-Gwóźdź
et al., 2013). In a different study, no itraconazole or efinaconazole re- The acquisition of mutations in genes encoding antifungal targets
sistance was observed in T. rubrum after 24 weeks of culturing (Iwata has resulted in resistance in many pathogenic fungi. Non-synonymous
et al., 2014). SNPs lead to amino-acid alterations, an altered structure and reduced
Another study included four pairs of C. glabrata isolates which binding affinity of the antifungal to the target (Balkis et al., 2002), thus
consisted of one susceptible and one resistant isolate, retrieved from the enzyme can remain functional. This mechanism has evolved in
four patients which received cumulative doses of 6, 10, 12, and 12.2 g isolates that were exposed to echinocandins, allylamines, and is the
of fluconazole. Interestingly, sampling times between the susceptible resistance mechanism found in pan-azole resistant A. fumigatus isolates
and the resistant isolates were 15, 35, 40 and 60 days, respectively, that have spread across all continents (Hagiwara et al., 2016b; Snelders
implying that lower treatment doses lead to a faster resistance devel- et al., 2008). Frequently reported amino acid substitutions that confer
opment (Sanguinetti et al., 2005). In vitro micafungin resistance was azole resistance in A. fumigatus include positions G54, G138, F216,
observed after exposure to increasing concentrations of the drug M220, and G448 (Camps et al., 2012a,b; Chowdhary et al., 2017;
starting at 0.015 μg/ml (Bordallo-Cardona et al., 2017). Resistance Vermeulen et al., 2013). Furthermore, specific point mutations are as-
mutations developed when drug concentrations were close to the MIC, sociated with a tandem repeat in the promoter region (Snelders et al.,
on average after 3 days. 2008; Van Der Linden et al., 2013). This repeat confers upregulation of
These findings support the statement that initial exposure to lower cyp51A expression, as a duplication of this region provides an addi-
drug concentrations lead to a faster resistance development, although tional binding site for the transcriptional regulator SrbA, which acti-
resistant isolates have been recovered from patients that have received vates expression of cyp51A (Hortschansky et al., 2016). Additionally,
prophylaxis with itraconazole for a longer period of time, even if serum azole resistance through increased transcription of cyp51A is also ob-
levels were higher than the MIC according to the EUCAST guidelines served in isolates harboring a P88L mutation in the HapE subunit of the
(Pilmis et al., 2018). Multi-resistant A. fumigatus is often recovered from CBC complex (Camps et al., 2012a,b). The TR34/L98H and the TR46/
the environment, for instance compost heaps, which contain a very low Y121F/T289A containing isolates were shown to be highly resistant to
concentration of agricultural azoles used for crop protection. Due to itraconazole and voriconazole, respectively (Snelders et al., 2008; Van
cross-resistance, these isolates are also resistant for medical azoles Der Linden et al., 2013). Various studies have highlighted that specific
(Snelders et al., 2012; Zhang et al., 2017b). The TR34/L98H and TR46/ point mutations cause the Cyp51A enzyme to be more unstable (Paul
Y121F/T289A cyp51A resistance conferring genotypes that likely have et al., 2017; Song et al., 2018), which could explain why a tandem
originated in the environment, have spread and are the cause of repeat is essential for a fully resistant phenotype, as unstable enzymes
treatment failure in patients with invasive aspergillosis (Fisher et al., can cause a decrease in overall fitness.
2018; Garcia-Rubio et al., 2017; Lestrade et al., 2019; Zhang et al., In Candida spp., there is much more variation in ERG11 point mu-
2017a). tations, but never accompanied by tandem repeat region variation (Fan
et al., 2018; Flowers et al., 2015). The most occurring SNPs that were
proved to confer azole resistance include Y132, K143, E266, and G464
3. Mechanisms of antifungal resistance
(Flowers et al., 2015). In C. neoformans, the single amino acid sub-
stitution G484S was shown to confer resistance against fluconazole
Fungal pathogens have developed a variety of molecular mechan-
(Bosco-Borgeat et al., 2016; Rodero et al., 2003).
isms that enable them to exhibit intrinsic or acquired resistance to all
In isolates resistant for the allylamine terbinafine, amino acid sub-
antifungal compounds currently used. For an overview of the me-
stitutions are often found in the squalene epoxidase, the target of ter-
chanisms highlighted in this review, see Fig. 1.
binafine. In A. fumigatus, a F389L substitution in the ergA gene was
found to confer resistance to terbinafine (Rocha et al., 2006). Terbi-
3.1. Biofilm formation nafine is commonly used to treat dermatophyte infections, but a recent
study observed a high prevalence of terbinafine resistant T. interdigitale
Many pathogenic yeasts and molds have the ability to form a bio- in India (Singh et al., 2018). They described the known F397L sub-
film, which works as a shield against host defense and antifungal stitution, and the L393F substitution in the squalene epoxidase gene
compounds and increases adherence to the host surface (Fanning and erg1. This last substitution has already been confirmed to result in
Mitchell, 2012). A biofilm consists of a network of cells that are asso- terbinafine resistance, as this substitution was introduced in the squa-
ciated to each other or a surface, which produce extracellular matrix lene epoxidase of C. albicans, and heterologously expressed in S. cere-
(ECM) that provides protection against a stressful environment visiae, which resulted in a resistant phenotype (Osborne et al., 2005).
(Ramage et al., 2012). The ECM formed is thought to prevent the dif- Echinocandin resistance through non-synonymous SNPs has been
fusion of antifungals to the cells. Furthermore, a high amount of ex- found in the FKS1 gene, which encodes the 1,3 β-glucan synthase, one
tracellular DNA was found in the EMC in C. albicans biofilms, which is of the main components of the fungal cell wall. In resistant isolates of C.
thought to contribute to the structure and formation of the biofilm albicans, the most common amino acid substitutions found is a serine
(Martins et al., 2010). Biofilms formed by C. albicans exhibit very high substitution in codon 645 or an arginine substitution in codon 1361
MICs against multiple antifungals after only 72 h of growth (Chandra (Balashov et al., 2006; Laverdière et al., 2006; Park et al., 2005). These
et al., 2001). Biofilms also conferred high levels of resistance against mutations are also found in other Candida spp. (Perlin et al., 2015).
antifungals of various classes in Trichosporon asahii, A. fumigatus and C. Although echinocandin resistance in A. fumigatus is rare, in a clinical
neoformans (Bonaventura et al., 2006; Kaur and Singh, 2015; Martinez isolate that was retrieved after micafungin treatment failure, a F675S
and Casadevall, 2015; Müller et al., 2011). A recent study showed that substitution was found in the FKS1 gene (Jiménez-Ortigosa et al.,
in high- biofilm forming isolates of C. albicans, various amino acid 2017). It is apparent that point mutations that confer echinocandin
metabolism pathways were significantly up-regulated compared to low- resistance in the FKS1 gene are present in regions 641–649 and
biofilm forming isolates. The AAT1 gene encoding an aspartate ami- 1345–1365 (Walker et al., 2010). Echinocandin resistance in C. glabrata
notransferase was found to play a central role in these metabolic is mediated by point mutations in FKS1 or in its paralog FKS2, although
pathways, and its significant up-regulation in high- biofilm forming only FKS1 point mutations were reported to result in resistant isolates
isolates suggests it could be a potential target for antifungal compound in C. albicans and C. dubliniensis (Katiyar et al., 2012; Prigent et al.,
development and as a biomarker for these high- biofilm forming isolates 2017). The latter study showed that in three patients, corresponding to
(Rajendran et al., 2016). 8% of all patients treated with echinocandins, resistance arose within

4
M.W.J. Hokken, et al. Fungal Genetics and Biology 132 (2019) 103254

one month of therapy. regulatory proteins can play in for instance azole resistance (Hagiwara
As described above, the antifungal 5-flucytosine (5FC) is currently et al., 2017).
only used in combination therapy, as it quickly induces resistance An overview of the regulatory networks determining efflux-pump
(Vermes et al., 2000). This resistance often results when the Fcy2 per- expression can be found in a review by Paul and Moye-Rowley (2014).
mease is mutated which prohibits 5FC to enter the cell, or when the Resistance to terbinafine via increased drug-efflux has not been re-
Fur1 or Fcy enzymes are mutated, so that the conversion of 5FC to the ported extensively. A transcriptome study in C. albicans highlighted
toxic 5-fluorouracil is prevented. Although low-level resistance may several up-regulated transporter genes among which MFS1 and CDR1,
occur due to minor amino acid substitution in the Fcy2 permease, full after exposure to 4 mg/L terbinafine (Yue-bin et al., 2007). Ad-
resistance is only achieved with mutations that result in a null-pheno- ditionally, terbinafine resistance through efflux-pump upregulation has
type for Fcy2 or Fur1 in C. albicans (Edlind and Katiyar, 2010). not been reported in Trichophyton spp. (Yamada et al., 2017).
In C. glabrata, the expression of membrane transporters CgFLR1
3.3. Metabolic bypass CgFLR2 resulted in a 5FC resistant phenotype. Deletion of these
transporters led to accumulation of intracellular 5FC, indicating that
A lesser observed mechanism to gain antifungal resistance, en- these transporters play a direct role in 5FC efflux (Pais et al., 2016).
compasses the prevention of accumulating toxic metabolites by the Aside from drug-efflux machineries, membrane transporters have
inactivation of another enzyme in the pathway. In C. albicans, this been identified that confer antifungal resistance via their decreased
mechanism has proven to confer resistance against azole compounds drug-influx capacity. Recently, it has been found in A. fumigatus that the
(Kelly et al., 1997). When the sterol Δ5,6-desaturase ERG3 is nonfunc- main importer of 5FC, the FcyB transporter, is upregulated at pH5
tional, 14α-methylated sterol intermediates are not converted to toxic which explains the limited activity of this drug at pH7 (Gsaller et al.,
sterol intermediates, preventing disturbance of the cellular membranes. 2018).
Furthermore, this inactivation does not always lead to a decrease in
virulence (Vale-Silva et al., 2012). Similar mechanisms were found in
the ERG2 and ERG6 genes of C. albicans (Vincent et al., 2013). Many 3.5. Mitochondrial alterations possibly facilitate resistance
non-cyp51A azole resistant isolates are not analyzed for mutations in
their ERG3 gene, so this resistance mechanism could be more wide- Various studies have elucidated potential roles for the mitochondria
spread than currently believed. However, deletion mutants of erg3A and to be involved in antifungal drug resistance (Brun et al., 2004; CHAB-
erg3B were created in A. fumigatus and did not lead to any differences in ASSE et al., 2009; Sanglard et al., 2001). A C. glabrata isolate with loss
antifungal drug susceptibility, although significant differences in er- of mitochondrial function showed increased fluconazole resistance up
gosterol content were found in all strains (Alcazar-fuoli et al., 2006). to 50 μg/ml, an effect that was appointed to the differential expression
This could be explained by the differences in ergosterol biosynthesis of azole-resistance related genes due to loss of mitochondrial function
found between fungal taxa (Alcazar-Fuoli and Mellado, 2012). (Sanglard et al., 2001).
A study on A. fumigatus strains that harbor an E180D mutation in
3.4. Overexpression of efflux-pumps the 29.9KD subunit of the mitochondrial complex I showed that these
strains were azole resistant, although also less virulent. The strain had
The ability of fungal pathogens to excrete antifungal compounds an MIC of 2–4 mg/l itraconazole, whereas the parental strain had an
works through overexpression of specific efflux pumps and is a re- MIC of 0.25 mg/l. This finding is supported by the fact that inactivation
sistance mechanism found in many fungal species. Although this me- of complex I through chemical inhibitors also confers azole resistance.
chanism is widespread, it is not found in echinocandin or polyene re- The authors suggest that this comes from a rebalancing of the hypoxic
sistant fungal isolates, as these are not substrates of efflux pumps and response. Upon addition of azoles, the hypoxic response is ‘unin-
exercise fungicidal activity on the outside of the cell (Cannon et al., tentionally’ activated by dysfunctional oxygen sensing because of sterol
2009). Increased gene expression feeds the synthesis of more efflux biosynthesis inhibition, although oxygen levels are normal (Bromley
pump protein complexes results in increased antifungals transportation et al., 2016). As the parental strain displays a reduced activity of
to outside the cell. These transporters usually belong to the ABC or MFS complex I when hypoxia is induced, the strain with the mutated com-
transporter family (Perlin et al., 2014) and blocking these transporters plex I is thought to restore its growth through this mechanism when
reduces the MICs significantly (Holmes et al., 2016). azoles are present.
Azole resistance through up-regulated efflux pump expression was A recent study on difenaconazole resistant A. fumigatus isolates
reported for the MDR1 (Multi-Drug Resistant), CDR1, and CDR2 conducted by Zhang et al., described mitochondrial DNA mutations in a
(Candida-Drug Resistant) genes in many Candida spp. (Choi et al., 2016; superoxide dismutase, the inner membrane protein Mdm31 and the
Moran et al., 1998; Sanglard et al., 2001, 1997). Homologs of these outer membrane protein Tom70 (Zhang et al., 2017b). These studies
transporters are also present in A. fumigatus and have shown to confer suggest that there might be a direct relation between mitochondrial
azole resistance (White, 1997). Furthermore, many unknown putative dysfunction and azole resistance.
drug-efflux transporters were shown to be up-regulated in A. fumigatus
after exposure to itraconazole (Hokken et al., 2019). These transporters
should be investigated to elucidate their potential role in developing 3.6. Activation of stress pathways
azole resistance. The C. neoformans and C. gatti genome also contain
MDR1 homologs which were shown to facilitate azole efflux (Basso Cellular stress signaling is essential to survive stressful conditions in
et al., 2015), however, the role of other known efflux pumps was never the environment, such as high salt concentrations or high temperatures
fully elucidated in these species. (Hagiwara et al., 2016a). The mechanisms that help cells cope with
Gain-of-function mutations in the Mrr1, Upc2, and Cap1 transcrip- stress, can also provide protection against drug-induced stress condi-
tion factors that regulate MDR1 expression, induced fluconazole re- tions (Cowen and Steinbach, 2008). One of the conserved key reg-
sistance in C. albicans (Schubert et al., 2011). Furthermore, the Tac1 ulators in stress signaling is the Hsp90 chaperone, which is involved in
transcription factor (TF) was shown to regulate the expression of the protein folding and stabilizing many proteins involved in signal trans-
CDR1 and CDR2 efflux-pumps, and the development of homozygosity duction (Cowen, 2015). In A. fumigatus and A. terreus, Hsp90 inhibitors
for this TF results in strong up-regulation of these pumps (Coste et al., lead to increased sensitivity to caspofungin. Furthermore, reduced
2006, 2004). In A. fumigatus, transcription factor AtrR was found to functioning of Hsp90 was shown to increase azole susceptibility in C.
regulate cyp51A and cdr1B expression, emphasizing the important role albicans (Cowen, 2009).

5
M.W.J. Hokken, et al. Fungal Genetics and Biology 132 (2019) 103254

3.7. Adjustment of membrane homeostasis environment is not ideal, this process is referred to as adaptive phe-
notypic plasticity (APP) (Stern et al., 2007). Before any mutations arise
Eukaryotic membranes are dynamic structures that contain many in the genome, the cells need to maintain growth and reproduction to
lipid species that contribute to membrane integrity and maintenance. give themselves a chance to develop hard-wired (i.e. genetic) resistance
Azoles and amphotericin B have an impact on membrane home- mutations. One way by which insight into these processes can be
ostasis by disturbing cellular ergosterol content, which leads to mem- gained, is through measuring the differences in gene expression, as
brane destabilization and subsequent lysis of the cell these are controlled by regulatory genes that are often mediated
(Georgopapadakou et al., 1987). Therefore, it is suggested that the cell through environmental impulses (Aubin-Horth and Renn, 2009;
adjust its membrane homeostasis to make it more rigid, or more fluid. Schlichting and Piglucci, 1993; Schlichting and Smith, 2002). Differ-
This can for instance be achieved by changing the ratio of phospholipid ential gene expression after drug exposure can lead to increased sur-
species (PLs), as all these PLs have different chemico-physical proper- vival chances through the activation of compensatory mechanisms. This
ties which contribute to the nature of the membrane (Renne and de gives the fungus a chance to develop resistance mutations by main-
Kroon, 2018). Additionally, sterols and sphingolipids also contribute to taining growth and reproduction. Various studies highlight differences
membrane integrity and are thought to function together in complexes in gene expression after exposure to antifungal compounds.
(Guan et al., 2009). In C. albicans, depletion of either ergosterol or A recent study using A. fumigatus clinical isolates exposed to itra-
sphingolipids was shown to increase drug-susceptibilities, an effect that conazole demonstrated the up-regulation of specific efflux pumps, dif-
was appointed to the increased permeability of the membrane. Fur- ferential expression patterns in ergosterol and phospholipid biosynth-
thermore, a decreased surface localization of the drug-efflux transporter esis routes, and in MAPK cascade proteins (Hokken et al., 2019).
Cdr1p was found in S. cerevisiae cells with disturbed sterol-sphingolipid Another study investigated the gene expression of C. albicans after
interactions. caspofungin exposure, and found that chitin production was increased
In another study, it was demonstrated that null-mutations in the (Walker et al., 2015). Compensatory mechanisms further include acti-
FEN1 or FEN12 genes confer azole resistance in C. albicans (Gao et al., vation of the calcineurin pathway, which activates several chitin syn-
2018). These enzymes synthesize sphingolipid precursors, and in- thases upon caspofungin exposure (Munro et al., 2007; Walker et al.,
activation resulted in a significant increase in various ceramide pre- 2015).
cursors, and an overall increase of sphingolipid species. In a recent transcriptomic analysis in C. glabrata, isogenic azole
susceptible and resistant isolates were assessed. It was found that ex-
pression of the adhesin-encoding EPA3 gene was increased by 4-fold in
4. Origins of antifungal resistance development
the resistant isolate, and that overexpression of this gene promoted
biofilm formation and decreased accumulation of radiolabeled clo-
The emergence of antifungal resistance mechanisms is the result of
trimazole (Cavalheiro et al., 2018). Finally, a particular example
natural selection after the origination of beneficial genetic of physio-
highlighted that these transcriptomic alterations are not only tem-
logic adaptations to antifungal compounds. The arise of these adapta-
porary; after culturing C. glabrata isolates in rich medium containing
tions is dependent on processes that influence genetic and physiological
50 μg/ml fluconazole, both the efflux transporter encoding genes CDR1
changes, such as genetic stability and sexual reproduction, and will be
and CDR2 were up-regulated, and remained up-regulated for up to 200
discussed below. For an overview of the processes that contribute to
generations after switching to drug-free medium (Sanglard et al., 2001).
adaptations to antifungal compounds, see Fig. 2.
These changes in gene expression help facilitate fungal resilience
against antifungal agents.
4.1. Increased antifungal drug resilience through phenotypic plasticity

Cells have the ability to change their metabolism as a response to 4.2. Mutation frequencies and mutation rates
changing conditions, as they are not able to move away from the un-
favorable environment, containing for instance antifungal compounds. The rate in which spontaneous mutations arise in an organism
When these responses allow cells to grow and divide when the greatly determines the genetic variability, and thus the chance that

Mutator strains Increased


mutation frequency Selection of resistant fungi
Normal strain
AGTCGGCATTCGAA HGT UV
In patients and environment
Wildtype TTGACAGAACCG ROS
Time ?
AGTCCGCATTCGAA
Aneuploidy
Mutator strain ATGACCGAAGCG
AGGCCTGATCCGAA Heterokaryon

Msh2

Generators of genomic variance


Population dynamics
G1 G2 G3
Parasexual
Phenotypic Maintainance of Sexual reproduction reproduction
plasticity multiple genotypes a
- FCZ CDR1B α Ascospore
Hyphal + nuclear
No regulation formation
EFH1 fusion
Up-regulation Increased Genetic
Up-regulation colonization reshuffling Haploidization
+ FCZ CDR1B 2n 1n

Fig. 2. Generation of genetic and physiologic variability in the population facilitates selection to antifungal agents.

6
M.W.J. Hokken, et al. Fungal Genetics and Biology 132 (2019) 103254

resistance conferring mutations are present in the population. Various also lead to a decreased fitness, no changes in virulence were found
studies have suggested a possible influence of the environment on (Boyce et al., 2017). Furthermore, more than half of the isolates of
mutational rates (Galhardo et al., 2012; Hoffman and Hercus, 2000). clinical C. glabrata isolates contained a mutation in the MSH2 gene, and
Stressful conditions could increase the mutational rate because main- msh2Δ strains were shown to develop resistance to caspofungin sig-
tenance and repair pathways function at a lower level, which results in nificantly faster than parental strain containing a wildtype MSH2 gene
an increased generation of mutations over time. Most of these muta- (Healey et al., 2016). Generation of a msh2Δ/msh2Δ C. albicans strain in
tions will be detrimental, but there is also a higher chance to develop the laboratory also resulted in elevated mutational frequencies on flu-
mutations which confer resistance to antifungal compounds. conazole-containing medium. Furthermore, rad50Δ/rad50Δ and
As exposure to antifungal compounds in patients or in the en- pms1Δ/pms1Δ strains were also shown to exhibit a higher appearance of
vironment creates a stressful environment for pathogenic fungi, these resistant colonies, emphasizing the important role of double-strand
mutational processes should be looked into. Processes that can be in- break repair and the DNA damage repair machineries (Legrand et al.,
vestigated to gain insight into the capabilities of a species to generate 2007).
mutations are the mutation frequency and the mutational rate of which However, in another study, clinical S. cerevisiae isolates retrieved
it is derived. from patients harboring a genotype which was previously indicated to
The mutation frequency is determined by calculating the ratio of exhibit a mutator phenotype in vitro, were only observed to have slight
viable colonies growing on drug-containing plates over the starting increase in mutational rate (Demogines et al., 2008; Skelly et al., 2017).
inoculum (Shields et al., 2019), whereas the mutation rate is the fre- Up to date, no mutator phenotypes of Aspergillus spp. have been found
quency of new mutations in general, in a single gene or cell over a nor emerged in the clinic, although laboratory strains with defective
certain time. For instance, the mutation rate of S. cerevisiae was de- DNA repair machinery by deletion of atrA or atmA were shown to ac-
termined to be 0.0027 per generation; the chance of mutation per base quire voriconazole resistance more frequently (dos Reis et al., 2017).
pair was set at 2.2 × 10−10, multiplied by the total genome size which Mutator phenotypes are more likely to be found in fungi with a
counts 1.2 × 107 basepairs (Drake et al., 1998; Gou et al., 2019). Al- haploid genome, as the phenotype will be determined by only one set of
though mutation frequency is much easier to determine and does not chromosomes. If haploid mutator cells with acquired resistance muta-
require whole genome sequencing, the latter method is able to distin- tions would undergo (para)sexual reproduction, rearrangement of ge-
guish between the mutational rates of different genes. This led to the netic material could lead to offspring harboring only the beneficial
observation that the mutational rate can vary across the genome (Lang resistance mutation, without suffering possible fitness costs of an in-
and Murray, 2008). stable genome.
The mutation frequency is thus dependent on the intrinsic muta-
tional rate of a species, but external factors which confer stressful 4.4. Chromosomal aneuploidy
growth conditions were also shown to influence this number.
In C. albicans, the spontaneous mutation frequency appeared to Antifungal resistance development is greatly driven by the ability to
differ between exposure to the various echinocandins, as treatment generate genomic variance in the gene pool. Polyploid cells often ra-
with caspofungin seemed to result into resistant isolates more often pidly generate genetic diversity within a population due to their in-
than the other echinocandins (Shields et al., 2019), although they share stability (Gerstein et al., 2015). Many yeasts have the ability to lose or
their cellular target. gain certain segments of chromosomes, or even complete chromosomes
Another particular example comes from a study conducted by Lamb (Gordon et al., 2011). Chromosomal aneuploidy facilitated antifungal
et al., where various wild strains of Penicillium lanosum and A. niger resistance in various pathogenic fungi, and in some cases even their
were isolated from two different sides of the same mountain (Lamb virulence (Kwon-chung and Chang, 2012).
et al., 2008). The south-facing side of the mountain received up to Pathogenic species such as C. albicans display this genome flexibility
800% more damaging solar radiation than the north-side of the which can lead to antifungal resistance. The trisomy of chromosome 5
mountain, which resulted in two very different environments. The was seven times as prevalent in fluconazole-resistant C. albicans isolates
isolates that were taken from the south-side appeared to have a as in fluconazole-sensitive isolates, and these chromosomes also dis-
2.44–8.86. fold higher mutation frequency in controlled laboratory played a high segmental aneuploidy (Selmecki et al., 2006). The an-
conditions, than the isolates that originated from the much milder, euploid chromosomes found were the trisomy of chromosome 7, and
north-side. These results possibly indicate that stressful environments the isochromosome 5 which consists of two copies of the left arm of the
generate more genomic diversity in filamentous fungi. original chromosome (Selmecki et al., 2009). Caspofungin tolerance
Mutation rates are currently only determined for S. cerevisae and was also found in aneuploid C. albicans isolates, which either lost one
Schizosaccharomyces pombe (Drake et al., 1998; Farlow et al., 2015). For copy of chromosome 5, or showed combined monosomy of the left arm
filamentous fungi, mutation rates are more difficult to determine be- and trisomy of the right arm of chromosome 5 (Yang et al., 2017). In-
cause of their typical hyphal growth (Baracho and Baracho, 2003), al- terestingly, stress due to chemicals was shown to increase the rate at
though knowledge about mutational rates for each species could give which spontaneous aneuploids occurred by 100-fold in S. cerevisiae,
valuable insights into the chances of developing antifungal resistance. although inhibition of HSP90 resulted in the highest chromosome in-
stability and subsequent aneuploidy (Chen et al., 2012). Resistance of
4.3. The development of mutator strains the haploid C. neoformans to fluconazole was appointed to the doubling
of chromosome 1, 4, 10 or 12 (Sionov et al., 2010). Recently, this an-
The occurrence of mutator strains is a phenomenon that is observed euploidy was discovered to be the result of the inability to degrade the
in various yeasts (Healey et al., 2016; Nicolas et al., 2014). Most often septum after cell division, resulting in cells with increased DNA content
by defects in DNA repair mechanisms, these strains gain mutations (Altamirano et al., 2017). These large cells have been described as
quickly which increases their chances to develop resistance mutations. ‘Titan cells’ and have been found to have higher chance of surviving
In a study conducted by Lang et al, the mutation rate of a mutator strain various stressors such as fluconazole, oxidative stress induced by H2O2
of S. cerevisiae was observed to be a 225-fold higher than the wildtype. and osmotic stress induced by high NaNO3 concentrations. Further-
This was determined by genomic analysis of a strain harboring a more, their haploid progeny were shown to display higher fluconazole
knockout of the msh2 gene, a core component of the DNA mismatch MIC values than the progeny of regular cells (Gerstein et al., 2015). In
repair machinery (Lang et al., 2013). The same gene was found to have addition, it was demonstrated that Titan cells can be induced in vitro by
a single-base deletion at position 1689 in C. neoformans, which led to an exposure to FCS, a compound that induces capsule growth. Interest-
increased mutation rate. Although a decrease of genome stability could ingly, bronchial alveolar lavage (BAL) was also able to induce titan cell

7
M.W.J. Hokken, et al. Fungal Genetics and Biology 132 (2019) 103254

formation, due to the presence of biomarkers for bacterial pepti- generation of genetic variability through parasexual reproduction is
doglycan, which could facilitate adaptation to antifungal compounds hypothesized to occur more likely in older, long-lasting fungal colonies,
(Dambuza et al., 2018). if different genotypes are present in the population (Verweij et al.,
Although aneuploidy can also result in deleterious defects and fit- 2016a,b). Observations on the properties of fungal populations revealed
ness loss in fungi (Oromendia et al., 2008), these results also indicate that there was a high growth rate variability in clinical isolates from
that aneuploidy can lead to an increase in genetic variability and an- Candida spp. on fluconazole (15 μg/ml) and itraconazole (0.71 μg/ml)
tifungal resistance. (Schoofs et al., 1997), implying high genetic variation in the popula-
tion, which could explain the differences in antifungal drug tolerance. A
4.5. (Para)sexual reproduction genetic analysis of 2026 A. fumigatus isolates where 6% of all isolates
was azole resistant, showed that the azole susceptible populations ex-
Pathogenic fungi can have various methods of reproduction, which hibited little genetic differences, but the azole resistant isolates dis-
potentially aid the development and spread of resistance mechanisms. played a much higher level of genetic differences (Ashu et al., 2017).
The formation of asexual conidia provides an easy method for the Another study on A. fumigatus lineages derived from culturing in the
fungus to spread new mutations that originated during hyphal growth, presence of the fungicide difenoconazole, observed the rise of different
whereas the formation of sexual spores leads to increased genetic morphotypes in one lineage. Variation between morphotypes was found
variability due to reshuffling of chromosomes, which facilitates adap- in colony size, sporulation speed and in mycelial density. The mixed
tation to a changing environment (Heitman, 2015; Hoekstra, 2005; population of morphotypes was shown to confer faster cross-resistance
Verweij et al., 2016a,b). For sexual reproduction to occur, fusion of to medical azoles than isolated morphotypes from the population
hypha of two different mating types is usually required, although C. (Zhang et al., 2017b). As fungal isolates often encompass a population
neoformans and C. albicans evolved a form of sexual reproduction in with mixed morphotypes, this example shows that there could be fre-
which only one mating type is required (Alby et al., 2009; Bui et al., quency-dependent selection involved in a population resulting in
2008). Alternatively, parasexual reproduction can occur when fungal maintenance of various genotypes. This process leads to a more stable
cells or hypha undergo cellular fusion and form heterokaryons, where community which increases the chances of survival in stressful en-
potential recombination of genetic material occurs (Pontecorvo et al., vironments, making these fungal populations difficult to combat.
1953). Parasexual reproduction is hypothesized to occur in long lasting A study conducted by Huang et al. also demonstrated maintenance
mycelia in for instance CF patients, as there must be various genotypes of several genoypes in a population (Huang and Kao, 2012). Evolu-
present which are still clonally related (Verweij et al., 2016a,b). A re- tionary dynamics that were studied in an isogenic C. albicans population
cent study showed recovery of A. fumigatus heterokaryons from patients exposed to variable doses of fluconazole, showed that various adaptive
with chronic diseases. The heterokaryons were exposed to antifungals events occurred in the proximity of fluconazole. Most adaptive events
and had a higher growth rate than heterozygous diploids. These find- resulted in a higher MIC for fluconazole, but prior adaptations some-
ings implicate that the potency to develop heterokaryons results in a times persisted in the population.
higher phenotypic plasticity, and a higher chance to adapt to antifungal Notably, wildtype clinical C. albicans isolates can influence their
compounds (Zhang et al., 2019). population size by regulation of the transcription factor EFH1, which
Sexual ascospores of A. fumigatus were shown to be more resistant to determined the level of gastrointestinal (GI) tract colonization.
draught, oxidative stress, and heat (Wyatt et al., 2013), making them Overexpression of EFH1 resulted in a reduced colonization, which led
more resilient to several stresses. However, not all species show this to a more commensal growth behavior rather than an invasive one,
fitness advantage in sexual spores, as A. nidulans does not produce which would result in invasive candidiasis (White et al., 2007). Cells
stress-resistant ascospores (Schoustra et al., 2010). Nevertheless, a re- that overexpress EFH1 showed a reduced colonization of the GI tract,
cent study showed that sexual reproduction in A. fumigatus might be whereas null mutants were shown to lead to a higher level of coloni-
accountable for the increased length of the tandem repeat in isolates zation. The factors that influence fungal population sizes should be
collected from compost samples. Through unequal meiotic crossing monitored, as it could impose a survival mechanism that cells utilize
over, mispairing of the tandem repeat region can lead to even longer when encountering stressful environments.
repeats which result in a higher copy number of cyp51A transcripts and
increased resistance to azoles (Zhang et al., 2017a). In C. neoformans,
clonal cells of the α mating type were shown to mate with themselves in 4.7. Horizontal gene transfer in fungi
α-α unisexual reproduction cycle. This form of reproduction frequently
lead to offspring with aneuploid genomes, leading to fitness advantages Finally, evolution of resistance mechanisms through exchange of
against azole compounds when an extra copy of chromosome 9 or 10 genetic material within a population is shown to be highly mediated by
was obtained (Ni et al., 2013). In C. albicans, the parasexual cycle was horizontally transferred genes (HTG) in prokaryotes, although this
shown to create fusants of homozygous cells, which displayed a higher process also takes place in eukaryotes to a lesser extent (Collins and
growth rate then their parent cells (Zhang et al., 2015). Saville, 1990; Keeling and Palmer, 2008; Lerminiaux and Cameron,
Thus, sexual and parasexual reproduction are processes which 2018). A study conducted by Nguyen et al., identified 273 HTG from
generate genetic variability and have shown to lead to fitness increases prokaryotic origin in the A. fumigatus genome, and 542 HTG in the A.
in various examples. They promote potential adaptation, although to flavus genome (Nguyen et al., 2015). Although these genes are mostly
date a direct link with increased antifungal resistance was only found in involved in secondary metabolism, a study was done on an A. fumigatus
A. fumigatus and C. neoformans. strain which was identified to harbor AfuMGTC, a potential virulence
gene of bacterial origin, although this gene did not increase virulence in
4.6. Population dynamics A. fumigatus (Gastebois et al., 2011). In another study, an investigation
of the pan-genomes of fungal species elucidated that the majority of
The evolution of a drug-resistant strain from a drug-susceptible HTG found in A. fumigatus (391) and C. neoformans (420) came from
population depends on multiple processes described earlier, which lead other fungal species. These genes were identified to originate from
to genetic variation in a population. Additionally, the size of a popu- closely related species or species that share the same ecological niche
lation affects the evolution of antifungal resistance. In a large popula- (McCarthy and Fitzpatrick, 2019).
tion, the chances are higher that beneficial mutations arise per gen- Although no HGT of antifungal resistance has been reported yet,
eration, which can spread in the population through sporulation or these studies highlight the potential role HGT could play in the devel-
sexual recombination (Huang and Kao, 2012). Furthermore, the opment of antifungal resistance.

8
M.W.J. Hokken, et al. Fungal Genetics and Biology 132 (2019) 103254

5. Concluding remarks fumigatus C-5 Sterol Desaturases Erg3A and Erg3B : role in sterol biosynthesis and
antifungal drug susceptibility aspergillus fumigatus C-5 sterol desaturases Erg3A and
Erg3B: role in sterol biosynthesis and antifungal drug susceptibility. Antimicrob.
The emergence of antifungal drug resistance mechanisms results Agents Chemother. 5, 453–460. https://doi.org/10.1128/AAC.50.2.453.
from adaptation, and the ability to create genetic variation within a Alexander, B.D., Johnson, M.D., Pfeiffer, C.D., Jiménez-Ortigosa, C., Catania, J., Booker,
population. R., Castanheira, M., Messer, S.A., Perlin, D.S., Pfaller, M.A., 2013. Increasing echi-
nocandin resistance in candida glabrata: Clinical failure correlates with presence of
It is crucial that these mechanisms are detected quickly, to be able FKS mutations and elevated minimum inhibitory concentrations. Clin. Infect. Dis. 56,
to anticipate on resistant populations and adjust treatment strategies 1724–1732. https://doi.org/10.1093/cid/cit136.
accordingly. New resistance mechanisms are likely to continue to arise Altamirano, S., Fang, D., Simmons, C., Sridhar, S., Wu, P., Sanyal, K., Kozubowski, L.,
2017. Fluconazole-induced ploidy change in cryptococcus neoformans results from
against all antifungal compounds, and understanding the underlying the uncoupling of cell growth and nuclear division. mSphere 2, 1–18. https://doi.
evolutionary processes remains crucial to recognize situations that can org/10.1128/msphere.00205-17.
facilitate the development of resistance mechanisms. The rising re- Ashu, E.E., Hagen, F., Chowdhary, A., Xu, J., Meis, J.F., 2017. Global population genetic
analysis of Aspergillus fumigatus. MSphere 2, 1–13.
sistance rates around the world emphasize this and worldwide sur-
Aubin-Horth, N., Renn, S.C.P., 2009. Genomic reaction norms: Using integrative biology
veillance should be increased to monitor resistance rates (Chowdhary to understand molecular mechanisms of phenotypic plasticity. Mol. Ecol. 18,
and Meis, 2018; Lestrade et al., 2019; Verweij et al., 2016a,b). 3763–3780. https://doi.org/10.1111/j.1365-294X.2009.04313.x.
Awareness for environmental ‘hotspots’ that facilitate resistance de- Balashov, S.V., Park, S., Perlin, D.S., 2006. Assessing resistance to the echinocandin an-
tifungal drug caspofungin in Candida albicans by profiling mutations in FKS1.
velopment is important, as for instance, there is a strong link between Antimicrob. Agents Chemother. 50, 2058–2063. https://doi.org/10.1128/AAC.
the use of agricultural fungicides and the emergence of resistant fungal 01653-05.
populations (Brilhante et al., 2019; Fisher et al., 2018; Snelders et al., Balkis, M.M., Leidich, S.D., Mukherjee, P.K., Ghannoum, M.A., 2002. Mechanisms of
fungal resistance: an overview. [Review] [102 refs]. Drugs 62, 1025–1040.
2009; Verweij et al., 2009; Zhang et al., 2017b). It is critical to monitor Baracho, M.S., Baracho, I.R., 2003. An analysis of the spontaneous mutation rate mea-
the use of agricultural fungicides, to limit resistance development in the surement in filamentous fungi. Genet. Mol. Biol. 26, 83–87.
environment (Verweij et al., 2009). The use of the same or similar Basso Jr, L.R., Gast, C.E., Bruzual, I., Wong, B., 2015. Identification and properties of
plasma membrane azole efflux pumps from the pathogenic fungi Cryptococcus gattii
molecules for different applications, i.e. crop protection and medical and Cryptococcus neoformans. J. Antimicrob. Chemother. 70, 1396. https://doi.org/
treatments, increases the risk of resistance selection that may affect the 10.1093/JAC/DKU554.
effectiveness of the compounds for other applications. Berkow, E.L., Lockhart, S.R., 2018. Activity of novel antifungal compound APX001A
against a large collection of Candida auris. The Journal of Antimicrobial che-
The chances of resistance mechanisms originating in a population motherapy 73 (11), 3060–3062. https://doi.org/10.1093/jac/dky302.
that is exposed to antifungal compounds, would supposedly remain Bhatt, V.R., Viola, G.M., Ferrajoli, A., 2011. Invasive fungal infections in acute leukemia.
limited if a second antifungal from a different chemical class is present, Ther. Adv. Hematol. 2, 231–247. https://doi.org/10.1177/2040620711410098.
Bonaventura, G. Di, Pompilio, A., Picciani, C., Iezzi, M., Antonio, D.D., Piccolomini, R.,
as the mutational chances are much smaller that two resistance me-
2006. Biofilm formation by the emerging fungal pathogen trichosporon asahii: de-
chanisms in different pathways arise at the same instant. Although the velopment architecture, and antifungal resistance. Antimicrobial Agents and
combined use of amphotericin B and 5-flucytosine was shown to reduce Chemotherapy 50, 3269–3276. https://doi.org/10.1128/AAC.00556-06.
mortality in cryptococcal menigitis, solid evidence is lacking for the Bordallo-Cardona, M., Escribano, P., de la Pedrosa, E., Marcos-Zambrano, L., Cantón, R.,
Bouza, E., Guinea, J., 2017. Vitro exposure to increasing micafungin concentrations
effectiveness of the combined use of other compounds in fungal infec- easily promotes echinocandin resistance in Candida glabrata isolates. Antimicrob.
tions, as these studies require time and are costly. Combinations of Agents Chemother. 61, 1–5.
antifungals should less frequently lead to the emergence of resistant Bosco-Borgeat, M.E., Taverna, C.G., Vivot, W., Murisengo, O.A., Mazza, M., Córdoba, S.,
Davel, G., 2016. Amino acid substitution in Cryptococcus neoformans lanosterol 14-
populations. Synergistic effects between antifungal compounds should α-demethylase involved in fluconazole resistance in clinical isolates. Rev. Argent.
be looked into, as this could even lead to a drop in MICs, possibly Microbiol. 48, 137–142. https://doi.org/10.1016/j.ram.2016.03.003.
permitting the use of lower dosages (Campitelli et al., 2017). Bossche, H. Vanden, Bellens, D., Cools, W., Gorrens, J., Marichal, P., Verhoeven, H.,
Willemsens, G., Coster, R. De, Beerens, D., Haelterman, C., Coene, M., Lauwers, W.,
Finally, several studies suggest the influence of environmental stress Jeune, L. Le, 1986. Cytochrome P- 450: Target for ltraconazole. Drug Dev. Res. 298,
on mutational rates (Galhardo et al., 2012; Lamb et al., 2008). If in- 287–298.
creased cellular stress does indeed lead to the faster adaptation through Boyce, K.J., Wang, Y., Verma, S., Shakya, V.P.S., Xue, C., Idnurm, A., 2017. Mismatch
repair of DNA replication errors contributes to microevolution in the pathogenic
increased mutational rates, and mutation frequencies appear to differ
fungus cryptococcus neoformans. MBio 8. https://doi.org/10.1128/mbio.00595-17.
between compounds, it is essential to look into these mechanisms to Brilhante, R.S., de Alencar, L.P., Bandeira, S.P., Sales, J.A., de Jesus Evangelista, A.J.,
facilitate safe compound development. Serpa, R., de Aguiar, Cordeiro R., de Aquino, Pereira-Neto W., Sidrim, J.J., Castelo,
D.D., Rocha, M.F., 2019. Exposure of Candida parapsilosis complex to agricultural
Antifungal resistance develops when fungal species adjust their
azoles: An overview of the role of environmental determinants for the development of
physiology or develop beneficial mutations during therapy, or via en- resistance. Sci. Total Environ. 650, 1231–1238. https://doi.org/10.1016/j.scitotenv.
vironmental exposure. This inevitable process of adaptation should be 2018.09.096.
restrained as much as possible, by unraveling the mechanistic and Bromley, M., Johns, A., Davies, E., Fraczek, M., Gilsenan, J.M., Kurbatova, N., Keays, M.,
Kapushesky, M., Gut, M., Gut, I., Denning, D.W., Bowyer, P., 2016. Mitochondrial
evolutionary factors that contribute to the spread of resistant fungal complex I is a global regulator of secondary metabolism, virulence and azole sensi-
pathogens. Understanding these factors will ultimately contribute to the tivity in fungi. PLoS One 11, 1–22. https://doi.org/10.1371/journal.pone.0158724.
development of new therapies and may improve treatment options for Brun, S., Berges, T., Poupard, P., Vauzelle-Moreau, C., Renier, G., Chabasse, D., Bouchara,
J.-P., 2004. Mechanisms of azole resistance in petite mutants of Candida glabrata.
patients suffering from fungal diseases worldwide. Antimicrob. Agents Chemother. 48, 1788–1796. https://doi.org/10.1128/aac.48.5.
1788-1796.2004.
Formatting of funding sources Bui, T., Lin, X., Malik, R., Heitman, J., Carter, D., 2008. Isolates of Cryptococcus neo-
formans from infected animals reveal genetic exchange in unisexual, α mating type
populations. Eukaryot. Cell 7, 1771–1780. https://doi.org/10.1128/EC.00097-08.
This research did not receive any specific grant from funding Buil, J.B., Snelders, E., Denardi, L.B., Melchers, W.J.G., Verweij, P.E., 2019. Trends in
agencies in the public, commercial, or not-for-profit sectors azole resistance in aspergillus fumigatus, the Netherlands, 1994–2016. Emerg. Infect.
Dis. 25, 176–178. https://doi.org/10.3201/eid2501.171925.
Campitelli, M., Zeineddine, N., Samaha, G., Maslak, S., 2017. Combination antifungal
References therapy: a review of current data. J. Clin. Med. Res. 9, 451–456. https://doi.org/10.
14740/jocmr2992w.
Camps, S.M.T., Dutilh, B.E., Arendrup, M.C., Rijs, A.J.M.M., Snelders, E., Huynen, M.A.,
Alby, K., Schaefer, D., Bennett, R.J., 2009. Homothallic and heterothallic mating in the
Verweij, P.E., Melchers, W.J.G., 2012a. Discovery of a hapE mutation that causes
opportunistic pathogen Candida albicans. Nature 460, 890–893. https://doi.org/10.
azole resistance in aspergillus fumigatus through whole genome sequencing and
1038/nature08252.
sexual crossing. PLoS One 7, e50034. https://doi.org/10.1371/journal.pone.
Alcazar-Fuoli, L., Mellado, E., 2012. Ergosterol biosynthesis in Aspergillus fumigatus: Its
0050034.
relevance as an antifungal target and role in antifungal drug resistance. Front.
Camps, S.M.T., Van Der Linden, J.W.M., Li, Y., Kuijper, E.J., Van Dissel, J.T., Verweij,
Microbiol. 3, 1–6. https://doi.org/10.3389/fmicb.2012.00439.
P.E., Melchers, W.J.G., 2012b. Rapid induction of multiple resistance mechanisms in
Alcazar-fuoli, L., Mellado, E., Buitrago, M.J., Lopez, J.F., Joan, O., Cuenca-estrella, J.M.,
Aspergillus fumigatus during azole therapy: a case study and review of the literature.
Juan, L., Garcia-effron, G., Grimalt, J.O., Rodriguez-tudela, J.L., 2006. Aspergillus
Antimicrob. Agents Chemother. 56, 10–16. https://doi.org/10.1128/AAC.05088-11.

9
M.W.J. Hokken, et al. Fungal Genetics and Biology 132 (2019) 103254

Cannon, R.D., Lamping, E., Holmes, A.R., Niimi, K., Baret, P.V., Keniya, M.V., Tanabe, K., complex DNA repair sensitivity phenotype in baker’s yeast. PLoS Genet. 4. https://
Niimi, M., Goffeau, A., Monk, B.C., 2009. Efflux-mediated antifungal drug resistance. doi.org/10.1371/journal.pgen.1000123.
Clin. Microbiol. Rev. 22, 291–321. https://doi.org/10.1128/cmr.00051-08. Denning, D.W., Pashley, C., Hartl, D., Wardlaw, A., Godet, C., Del Giacco, S., Delhaes, L.,
Castanheira, M., Woosley, L.N., Diekema, D.J., Messer, S.A., Jones, R.N., Pfaller, M.A., Sergejeva, S., 2014. Fungal allergy in asthma–state of the art and research needs.
2010. Low prevalence of fks1 hot spot 1 mutations in a worldwide collection of Clin. Transl. Allergy 4, 1–23. https://doi.org/10.1186/2045-7022-4-14.
candida strains. Antimicrob. Agents Chemother. 54, 2655–2659. https://doi.org/10. Denning, D.W., Venkateswarlu, K., Oakley, K.L., Anderson, M.J., Manning, N.J., Stevens,
1128/AAC.01711-09. D.A., Warnock, D.W., Kelly, S.L., 1997. Itraconazole resistance in Aspergillus fumi-
Cavalheiro, M., Costa, C., Silva-Dias, A., Miranda, I.M., Wang, C., Pais, P., Pinto, S.N., gatus. Antimicrob. Agents Chemother. 41, 1364–1368.
Mol-Homens, D., Sato-Okamoto, M., Takahashi-Nakaguchi, A., Silva, R.M., Mira, dos Reis, T.F., Silva, L.P., de Castro, P.A., de Lima do Carmo, P.B.R.A., Marini, M.M., da
N.P., Fialho, A.M., Chibana, H., Rodrigues, A.G., Butler, G., Teixeira, M.C., 2018. A Silveira, J.F., Ferreira, B.H., Rodrigues, F., Malavazi, I., Goldman, G.H., 2017. The
transcriptomics approach to unveiling the mechanisms of in vitro evolution towards influence of genetic stability on aspergillus fumigatus virulence and azole resistance.
fluconazole resistance of a Candida glabrata clinical isolate. Antimicrob. Agents G3: Genes Genom. Genet. 8, 265–278. https://doi.org/10.1534/g3.117.300265.
Chemother. 63, 1–17. Drake, J.W., Charlesworth, B., Charlesworth, D., Crow, J.F., 1998. Rates of spontaneous
Chabasse, D., Planchenault, C., Defontaine, A., Hallet, J.-N., Declerk, P., Bouchara, J.-P., mutation. Genetics 78, 1209–1221. online ISSN: 1943-2631. Source: https://www.
2009. In-vitro resistance to azoles associated with mitochondrial DNA deficiency in genetics.org/content/148/4/1667.
Candida glabrata. J. Med. Microbiol. 48, 663–670. https://doi.org/10.1099/ Edlind, T.D., Katiyar, S.K., 2010. Mutational analysis of flucytosine resistance in Candida
00222615-48-7-663. glabrata. Antimicrob. Agents Chemother. 54, 4733–4738. https://doi.org/10.1128/
Chandra, J., Kuhn, D.M., Mukherjee, P.K., Hoyer, L.L., Cormick, T.M.C., Ghannoum, M.A., AAC.00605-10.
2001. Biofilm formation by the fungal pathogen Candida albicans: development, Eschenauer, G., DePestel, D.D., Carver, P.L., 2007. Comparison of echninocandin anti-
architecture, and drug resistance. J. Bacteriol. 183, 5385–5394. https://doi.org/10. fungals. Ther. Clin. Risk Manag. 3, 71–97. https://doi.org/10.2147/tcrm.2007.3.
1128/JB.183.18.5385. 1.71.
Charlier, C., Sissy, E., Bachelier-bassi, S., Scemla, A., Quesne, G., Sitterlé, E., Legendre, C., Escribano, P., Recio, S., Peláez, T., González-Rivera, M., Bouza, E., Guinea, J., 2012. In
Lortholary, O., Bougnoux, M., 2016. Acquired flucytosine resistance during combi- vitro acquisition of secondary azole resistance in Aspergillus fumigatus isolates after
nation therapy with caspofungin and flucytosine for Candida glabrata cystitis. prolonged exposure to itraconazole: Presence of heteroresistant populations.
Antimicrob. Agents Chemother. 60, 662–665. https://doi.org/10.1128/AAC.02265- Antimicrob. Agents Chemother. 56, 174–178. https://doi.org/10.1128/AAC.
15.Address. 00301-11.
Chen, C.A., Sorrell, T.C., 2007. Antifungal agents. Med. J. Aust. 187. Fan, X., Xiao, M., Zhang, D., Huang, J.-J., Wang, H., Hou, X., Zhang, L., Kong, F., Chen,
Chen, G., Bradford, W.D., Seidel, C.W., Li, R., City, K., Physiology, I., Boulevard, R., City, S.C.-A., Tong, Z.-H., Xu, Y.-C., 2018. Molecular mechanisms of azole resistance in
K., 2012. Induction of aneuploidy. Nature 482, 246–250. https://doi.org/10.1038/ Candida tropicalis isolates causing invasive candidiasis in China. Clin. Microbiol.
nature10795.Hsp90. Infect. https://doi.org/10.1016/j.cmi.2018.11.007.
Choi, M.J., Won, E.J., Shin, J.H., Kim, S.H., Lee, K., Kim, M.-N., Lee, K., Shin, M.G., Suh, Fanning, S., Mitchell, A.P., 2012. Fungal biofilms. PLoS Pathog. 8, 1–4. https://doi.org/
S.P., Ryang, D.W., Im, Y.J., 2016. Resistance mechanisms and clinical features of 10.1371/journal.ppat.1002585.
fluconazole-nonsusceptible Candida tropicalis isolates compared with fluconazole- Farlow, A., Long, H., Arnoux, S., Sung, W., Doak, T.G., Nordborg, M., Lynch, M., 2015.
less-susceptible isolates. Antimicrob. Agents Chemother. 60, 3653–3661. https://doi. The spontaneous mutation rate in the fission yeast Schizosaccharomyces pombe.
org/10.1128/aac.02652-15. Genetics 201, 737–744. https://doi.org/10.1534/genetics.115.177329.
Chotirmall, S.H., Mirkovic, B., Lavelle, G.M., McElvaney, N.G., 2014. Immunoevasive Fisher, M.C., Hawkins, N.J., Sanglard, D., Gurr, S.J., 2018. Health and food security.
aspergillus virulence factors. Mycopathologia 178, 363–370. https://doi.org/10. Science (80-.) 742, 739–742.
1007/s11046-014-9768-y. Flowers, S.A., Colón, B., Whaley, S.G., Schuler, M.A., David Rogers, P., 2015.
Chowdhary, A., Meis, J.F., 2018. Emergence of azole resistant Aspergillus fumigatus and Contribution of clinically derived mutations in ERG11 to azole resistance in Candida
One Health: time to implement environmental stewardship. Environ. Microbiol. 20, albicans. Antimicrob. Agents Chemother. 59, 450–460. https://doi.org/10.1128/
1299–1301. https://doi.org/10.1111/1462-2920.14055. AAC.03470-14.
Chowdhary, A., Sharma, C., Meis, J.F., 2017. Azole-resistant aspergillosis: epidemiology, Forastiero, A., Pelaez, T., Lopez, J.F., Grimalt, J.O., Cuesta, I., Zaragoza, O., Mellado, E.,
molecular mechanisms, and treatment. J. Infect. Dis. 216, S436–S444. https://doi. 2013. Candida tropicalis antifungal cross-resistance is related to different.
org/10.1093/infdis/jix210. Antimicrob. Agents Chemother. 57, 4769–4781. https://doi.org/10.1128/AAC.
Collins, R.A., Saville, B.J., 1990. Independent transfer of mitochondrial chromosomes and 00477-13.
plasmids during unstable vegetative fusion in Neurospora. Nature 345, 177–179. Galhardo, R.S., Hastings, P.J., Rosenberg, S.M., 2012. Mutation as a stress response and
https://doi.org/10.1038/345177a0. the regulation of evolvability Rodrigo. Crit. Rev. Biochem. Mol. Biol. https://doi.org/
Cordonnier, C., Cesaro, S., Maschmeyer, G., Einsele, H., Peter Donnelly, J., Alanio, A., 10.1080/10409230701648502.Mutation.
Hauser, P., Lagrou, K., Melchers, W.J.G., Helweg-Larsen, J., Matos, O., Bretagne, S., Gao, J., Wang, H., Li, Z., Wong, A.H.-H., Wang, Y.-Z., Guo, Y., Lin, X., Zeng, G., Wang, Y.,
Maertens, J., Agrawal, S., Kibbler, C., Pagliuca, A., Ward, K., Akova, M., Herbrecht, Wang, J., 2018. Candida albicans gains azole resistance by altering sphingolipid
R., Mallet, V., Ribaud, P., Aljurf, M., Averbuch, D., Engelhard, D., Berg, T., Cornely, composition. Nat. Commun. 9, 4495. https://doi.org/10.1038/s41467-018-06944-1.
O., Penack, O., van Boemmel, F., von Lilienfeld-Toal, M., Blennow, O., Ljungman, P., Garcia-Rubio, R., Cuenca-Estrella, M., Mellado, E., 2017. Triazole resistance in aspergillus
Bruggemann, R., Donnelly, P., Kullberg, B.J., Melchers, W., Calandra, T., Hirsch, H., species: an emerging problem. Drugs 77, 599–613. https://doi.org/10.1007/s40265-
Marchetti, O., Orasch, C., Tissot, F., Castagnola, E., Girmenia, C., Mikulska, M., 017-0714-4.
Pagano, L., Viscoli, C., De La Camara, R., Duarte, R., Munoz, P., Drgona, L., Gastebois, A., Blanc Potard, A.-B., Gribaldo, S., Beau, R., Latgé, J.P., Mouyna, I., 2011.
Hargreaves, R., Hubacek, P., Kouba, M., Racil, Z., Klyasova, G., Pettrikos, G., Roilides, Phylogenetic and functional analysis of aspergillus fumigatus MGTC, a fungal protein
E., Skiada, A., Rizzi-Puechal, V., Sinko, J., Slavin, M., Styczynski, J., Tweddle, L., homologous to a bacterial virulence factor. Appl. Environ. Microbiol. 77, 4700–4703.
Wood, C., 2016. Pneumocystis jiroveci pneumonia: Still a concern in patients with https://doi.org/10.1128/aem.00243-11.
haematological malignancies and stem cell transplant recipients. J. Antimicrob. Georgopapadakou, N.H., Dix, B.A., Smith, S.A., Freudenberger, J., Funke, P.T., 1987.
Chemother. 71, 2379–2385. https://doi.org/10.1093/jac/dkw155. Effect of antifungal agents on lipid biosynthesis and membrane integrity in Candida
Coste, A., Turner, V., Ischer, F., Morschhäuser, J., Forche, A., Selmecki, A., Berman, J., albicans. Antimicrob. Agents Chemother. 31, 46–51. https://doi.org/10.1128/AAC.
Bille, J., Sanglard, D., 2006. A mutation in Tac1p, a transcription factor regulating 31.1.46.
CDR1 and CDR2, is coupled with loss of heterozygosity at chromosome 5 to mediate Gerstein, A.C., Fu, M.S., Mukaremera, L., Li, Z., Ormerod, K.L., Fraser, J.A., Berman, J.,
antifungal resistance in Candida albicans. Genetics 172, 2139–2156. https://doi.org/ Nielsen, K., 2015. Polyploid titan cells produce haploid and aneuploid progeny to
10.1534/genetics.105.054767. promote stress adaptation. MBio 6, 1–14. https://doi.org/10.1128/mBio. 01340-15.
Coste, A.T., Karababa, M., Ischer, F., Bille, J., Sanglard, D., 2004. TAC1, transcriptional Editor.
activator of CDR genes, is a new transcription factor involved in the regulation of Ghannoum, M.A., Rice, L.B., 1999. Antifungal agents: mode of action, mechanisms of
Candida albicans ABC transporters CDR1 and CDR2. Eukaryot. Cell 3, 1639–1652. resistance, and correlation of these mechanisms with bacterial resistance. Clin.
https://doi.org/10.1128/EC.3.6.1639-1652.2004. Microbiol. Rev. 12, 501–517. Source: https://www.ncbi.nlm.nih.gov/pmc/articles/
Cowen, L.E., 2015. Hsp90 Potentiates the rapidevolution of new traits: drugresistance in PMC88922/.
diverse fungi. Science (80-.) 309, 1–6. https://doi.org/10.1126/science.1118370. Gordon, J.L., Byrne, K.P., Wolfe, K.H., 2011. Mechanisms of chromosome number evo-
Cowen, L.E., 2009. Hsp90 orchestrates stress response signaling governing fungal drug lution in yeast. PLoS Genet. 7. https://doi.org/10.1371/journal.pgen.1002190.
resistance. PLoS Pathog. 5, e1000471. https://doi.org/10.1371/journal.ppat. Gou, L., Bloom, J.S., Kruglyak, L., 2019. The genetic basis of mutation rate variation in
1000471. yeast. Genetics 211, 731–740. https://doi.org/10.1534/genetics.118.301609.
Cowen, L.E., Sanglard, D., Calabrese, D., Surjusingh, C., Anderson, J.B., Kohn, L.M., 2000. Graminha, M.A.S., Rocha, E.M.F., Prade, R.A., Martinez-Rossi, N.M., 2004. Terbinafine
Population genomics of drug resistance in experimental populations of Candida al- resistance mediated by salicylate 1-monooxygenase in Aspergillus nidulans.
bicans. J. Bacteriol. 99, 9284–9289. Antimicrob. Agents Chemother. 48, 3530–3535. https://doi.org/10.1128/AAC.48.9.
Cowen, L.E., Steinbach, W.J., 2008. Stress, drugs, and evolution: the role of cellular 3530-3535.2004.
signaling in fungal drug resistance. Eukaryot. Cell 7, 747–764. https://doi.org/10. Gsaller, F., Furukawa, T., Carr, P.D., Rash, B., Bertuzzi, M., Bignell, E.M., Bromley, M.J.,
1128/ec.00041-08. 2018. Mechanistic basis of ph-dependent 5-flucytosine resistance in aspergillus fu-
Dambuza, I.M., Drake, T., Chapuis, A., Zhou, X., Correia, J., Taylor-Smith, L., LeGrave, N., migatus. Antimicrob. Agents Chemother. 1–9.
Rasmussen, T., Fisher, M.C., Bicanic, T., Harrison, T.S., Jaspars, M., May, R., Brown, Guan, X.L., Souza, C.M., Pichler, H., Schaad, O., Kajiwara, K., Wakabayashi, H., Ivanova,
G., Yuecel, R., MacCallum, D.M., Ballou, E.R., 2018. The Cryptococcus neoformans T., Castillon, G.A., Piccolis, M., Abe, F., Loewith, R., Funato, K., Wenk, M.R.,
Titan cell is an inducible and regulated morphotype underlying pathogenesis. PLOS Riezman, H., 2009. Functional interactions between sphingolipids and sterols in
Pathog. 47, 377–384. https://doi.org/10.1097/00010694-193905000-00005. biological membranes regulating cell physiology. Mol. Biol. Cell 20, 2083–2095.
Demogines, A., Smith, E., Kruglyak, L., Alani, E., 2008. Identification and dissection of a https://doi.org/10.1091/mbc.E08.

10
M.W.J. Hokken, et al. Fungal Genetics and Biology 132 (2019) 103254

Hagiwara, D., 2018. Current status of azole-resistant Aspergillus fumigatus isolates in East 1201/b18707.
Asia: China, Japan, Korea and Taiwan. Med. Mycol. 59, 71–76. Ksiezopolska, E., Gabaldón, T., 2018. Evolutionary emergence of drug resistance in
Hagiwara, D., Miura, D., Shimizu, K., Paul, S., Ohba, A., Gonoi, T., Watanabe, A., Kamei, Candida opportunistic pathogens. Genes (Basel) 9, 461. https://doi.org/10.3390/
K., Shintani, T., Moye-Rowley, W.S., Kawamoto, S., Gomi, K., 2017. A novel Zn2-Cys6 genes9090461.
transcription factor AtrR plays a key role in an azole resistance mechanism of as- Kwon-chung, K.J., Chang, Y.C., 2012. Aneuploidy and drug resistance in pathogenic
pergillus fumigatus by co-regulating cyp51A and cdr1B expressions. PLoS Pathogens. fungi. PLoS Pathogens 8, 8–11. https://doi.org/10.1371/journal.ppat.1003022.
https://doi.org/10.1371/journal.ppat.1006096. Lamb, B.C., Mandaokar, S., Bahsoun, B., Grishkan, I., Nevo, E., 2008. Differences in
Hagiwara, D., Sakamoto, K., Abe, K., Gomi, K., 2016a. Signaling pathways for stress re- spontaneous mutation frequencies as a function of environmental stress in soil fungi
sponses and adaptation in Aspergillus species: stress biology in the post-genomic era. at “Evolution Canyon” Israel. Proc. Natl. Acad. Sci. U. S. A. 105, 5792–5796. https://
Biosci. Biotechnol. Biochem. 80, 1667–1680. https://doi.org/10.1080/09168451. doi.org/10.1073/pnas.0801995105.
2016.1162085. Lamb, D., Kelly, D., Kelly, S., 1999. Molecular aspects of azole antifungal action and
Hagiwara, D., Watanabe, A., Kamei, K., Goldman, G.H., 2016b. Epidemiological and resistance. Drug Resist. Updat. 2, 390–402. https://doi.org/10.1054/drup.1999.
genomic landscape of azole resistance mechanisms in aspergillus fungi. Front. 0112.
Microbiol. 7, 1382. https://doi.org/10.3389/fmicb.2016.01382. Lang, G.I., Murray, A.W., 2008. Estimating the per-base-pair mutation rate in the yeast
Händel, N., De La Sayette, S., Verweij, P.E., Brul, S., ter Kuile, B.H., 2015. De novo in- saccharomyces cerevisiae. Genetics 82, 67–82. https://doi.org/10.1534/genetics.
duction of resistance against voriconazole in Aspergillus fumigatus. J. Glob. 107.071506.
Antimicrob. Resist. 3, 52–53. https://doi.org/10.1016/j.jgar.2015.01.001. Lang, G.I., Parsons, L., Gammie, A.E., 2013. Mutation rates, spectra, and genome-wide
Healey, K.R., Zhao, Y., Perez, W.B., Lockhart, S.R., Sobel, J.D., Farmakiotis, D., distribution of spontaneous mutations in mismatch repair deficient yeast. G3: Genes
Kontoyiannis, D.P., Sanglard, D., Taj-Aldeen, S.J., Alexander, B.D., Jimenez-Ortigosa, Genom Genet. 3, 1453–1465. https://doi.org/10.1534/g3.113.006429.
C., Shor, E., Perlin, D.S., 2016. Prevalent mutator genotype identified in fungal pa- Lat, A., Thompson, G., 2011. Update on the optimal use of voriconazole for invasive
thogen Candida glabrata promotes multi-drug resistance. Nat. Commun. 7, 1–10. fungal infections. Infect. Drug Resist. 43. https://doi.org/10.2147/idr.s12714.
https://doi.org/10.1038/ncomms11128. Latgé, J.P., 2001. The pathobiology of Aspergillus fumigatus. Trends Microbiol. 9,
Heitman, J., 2015. Evolution of sexual reproduction: a view from the fungal kingdom 382–389. https://doi.org/10.1016/S0966-842X(01)02104-7.
supports an evolutionary epoch with sex before sexes. Fungal Biol. Rev. 29, 108–117. Laverdière, M., Lalonde, R.G., Baril, J.G., Sheppard, D.C., Park, S., Perlin, D.S., 2006.
https://doi.org/10.1016/j.fbr.2015.08.002. Progressive loss of echinocandin activity following prolonged use for treatment of
Hoekstra, R.F., 2005. Why sex is good. Nature 434, 571. https://doi.org/10.1021/ Candida albicans oesophagitis. J. Antimicrob. Chemother. 57, 705–708. https://doi.
jp0449511. org/10.1093/jac/dkl022.
Hoffman, A., Hercus, M.J., 2000. Environmental stress as an evolutionary force. Legrand, M., Chan, C.L., Jauert, P.A., Kirkpatrick, D.T., 2007. Role of DNA mismatch
Bioscience 50, 217–226. repair and double-strand break repair in genome stability and antifungal drug re-
Hokken, M.W.J., Zoll, J., Coolen, J.P.M., Zwaan, B.J., Verweij, P.E., Melchers, W.J.G., sistance in Candida albicans. Eukaryot. Cell 6, 2194–2205. https://doi.org/10.1128/
2019. Phenotypic plasticity and the evolution of azole resistance in Aspergillus fu- EC.00299-07.
migatus; an expression profile of clinical isolates upon exposure to itraconazole. BMC Lerminiaux, N.A., Cameron, A.D.S., 2018. Horizontal transfer of antibiotic resistance
Genom. 1–17. genes in clinical environments. Can. J. Microbiol. 65, 34–44. https://doi.org/10.
Holmes, A.R., Cardno, T.S., Strouse, J.J., Ivnitski-Steele, I., Keniya, M.V., Lackovic, K., 1139/cjm-2018-0275.
Monk, B.C., Sklar, L.A., Cannon, R.D., 2016. Targeting efflux pumps to overcome Lestrade, P.P., Bentvelsen, R.G., Schauwvlieghe, A.F.A.D., Schalekamp, S., van der
antifungal drug resistance. Future Med. Chem. 8, 1485–1501. https://doi.org/10. Velden, W.J.F.M., Kuiper, E.J., van Paassen, J., van der Hoven, B., van der Lee, H.A.,
4155/fmc-2016-0050. Melchers, W.J.G., de Haan, A.F., van der Hoeven, H.L., Rijnders, B.J.A., van der Beek,
Hortschansky, P., Haas, H., Huber, E.M., Groll, M., Brakhage, A.A., 2016. The CCAAT- M.T., Verweij, P.E., 2019a. Voriconazole resistance and mortality in invasive
binding complex (CBC) in Aspergillus species. Biochim. Biophys. Acta – Gene Regul. Aspergillosis: A multicenter retrospective cohort study. Clin. Infect. Dis. 68,
Mech. https://doi.org/10.1016/j.bbagrm.2016.11.008. 1463–1471. https://doi.org/10.1093/cid/ciy859.
Hoving, J.C., Kolls, J.K., 2017. New advances in understanding the host immune response Lestrade, P.P.A., Meis, J.F., Melchers, W.J.G., Verweij, P.E., 2019b. Triazole resistance in
to Pneumocystis. Curr. Opin. Microbiol. 40, 65–71. https://doi.org/10.1016/j.mib. Aspergillus fumigatus: recent insights and challenges for patient management. Clin.
2017.10.019. Microbiol. Infect. https://doi.org/10.1016/j.cmi.2018.11.027.
Hryncewicz-Gwóźdź, A., Kalinowska, K., Plomer-Niezgoda, E., Bielecki, J., Jagielski, T., Liu, M., Zheng, N., Li, D., Zheng, H., Zhang, L., Ge, H., Liu, W., 2016. Cyp51A-based
2013. Increase in resistance to fluconazole and itraconazole in trichophyton rubrum mechanism of azole resistance in Aspergillus fumigatus: Illustration by a new 3D
clinical isolates by sequential passages in vitro under drug pressure. Mycopathologia structural model of Aspergillus fumigatus CYP51A protein. Med. Mycol. 54, 400–408.
176, 49–55. https://doi.org/10.1007/s11046-013-9655-y. https://doi.org/10.1093/mmy/myv102.
Huang, M., Kao, K.C., 2012. Population dynamics and the evolution of antifungal drug Luzia, F.R., Lais, L., Maria, P., Moretti, L., Pham, C.D., Lockhart, S.R., Zaninelli, A., 2018.
resistance in Candida albicans. FEMS Microbiol. Lett. https://doi.org/10.1111/j. Surveillance for azoles resistance in Aspergillus spp. highlights a high number of
1574-6968.2012.02587.x. amphotericin B-resistant isolates. Mycoses 360–365. https://doi.org/10.1111/myc.
Huang, M., McClellan, M., Berman, J., Kao, K.C., 2011. Evolutionary dynamics of candida 12759.
albicans during in vitro evolution. Eukaryot. Cell 10, 1413–1421. https://doi.org/10. Malcolm, T.R., Chin-Hong, P.V., 2013. Endemic mycoses in immunocompromised hosts.
1128/ec.05168-11. Curr. Infect. Dis. Rep. 15, 536–543. https://doi.org/10.1007/s11908-013-0387-4.
Iwata, A., Watanabe, Y., Kumagai, N., Katafuchi-Nagashima, M., Sugiura, K., Pillai, R., Maligie, M.A., Selitrennikoff, C.P., 2005. Cryptococcus neoformans resistance to echi-
Tatsumi, Y., 2014. In vitro and in vivo assessment of dermatophyte acquired re- nocandins: (1,3)β-glucan synthase activity is sensitive to echinocandins. Antimicrob.
sistance to efinaconazole, a novel triazole antifungal. Antimicrob. Agents Chemother. Agents Chemother. 49, 2851–2856. https://doi.org/10.1128/AAC.49.7.2851-2856.
58, 4920–4922. https://doi.org/10.1128/AAC.02703-13. 2005.
Jenks, J.D., Salzer, H.J.F., Prattes, J., Krause, R., Buchheidt, D., Hoenigl, M., 2018. Martin, M.V., 1999. The use of fluconazole and itraconazole in the treatment of Candida
Spotlight on isavuconazole in the treatment of invasive aspergillosis and mucormy- abicans infections: a review. J. Antimicrob. Chemother. 44, 429–437. https://doi.
cosis: design, development, and place in therapy. Drug Des. Devel. Ther. 12, org/10.1093/jac/44.4.429.
1033–1044. https://doi.org/10.2147/DDDT.S145545. Martinez, L.R., Casadevall, A., 2015. Biofilm Formation by Cryptococcus neoformans.
Jensen, R.H., Johansen, H.K., Arendrup, M.C., 2013. Stepwise development of a homo- Microbiol. Spectr. 6, 513–521. https://doi.org/10.1128/microbiolspec.MB-0006-
zygous S80P substitution in Fks1p, conferring echinocandin resistance in Candida 2014.Correspondence.
tropicalis. Antimicrob. Agents Chemother. 57, 614–617. https://doi.org/10.1128/ Martins, M., Uppuluri, P., Thomas, D.P., Cleary, I.A., Henriques, M., Lopez-Ribot, J.L.,
AAC.01193-12. Oliveira, R., 2010. Presence of extracellular DNA in the Candida albicans biofilm
Jiménez-Ortigosa, C., Moore, C., Denning, D.W., Perlin, D.S., 2017. Emergence of echi- matrix and its contribution to biofilms. Mycopathologia 169, 323–331. https://doi.
nocandin resistance due to a point mutation in the fks1 gene of aspergillus fumigatus org/10.1007/s11046-009-9264-y.Presence.
in a patient with chronic. Pulmonary Aspergillosis 61, 1–6. McCarthy, C.G.P., Fitzpatrick, D.A., 2019. Pan-genome analyses of model fungal species.
Katiyar, S.K., Alastruey-Izquierdo, A., Healey, K.R., Johnson, M.E., Perlin, D.S., Edlind, Microb. Genom. 1–23. https://doi.org/10.1099/mgen.0.000243.
T.D., 2012. Fks1 and Fks2 are functionally redundant but differentially regulated in Mesa-Arango, A.C., Scorzoni, L., Zaragoza, O., 2012. It only takes one to do many jobs:
Candida glabrata: implications for echinocandin resistance. Antimicrob. Agents Amphotericin B as antifungal and immunomodulatory drug. Front. Microbiol. 3,
Chemother. 56, 6304–6309. https://doi.org/10.1128/aac.00813-12. 1–10. https://doi.org/10.3389/fmicb.2012.00286.
Kaur, S., Singh, S., 2015. Biofilm formation by Aspergillus fumigatus. J. Music Ther. 52, Moran, G.P., Sanglard, D., Donnelly, S.M., Shanley, D.B., Sullivan, D.J., Coleman, D.C.,
2–9. https://doi.org/10.3109/13693786.2013.819592. 1998. Identification and expression of multidrug transporters responsible for fluco-
Keeling, P.J., Palmer, J.D., 2008. Horizontal gene transfer in eukaryotic evolution. Nat. nazole resistance in Candida dubliniensis. Antimicrob. Agents Chemother. 42,
Rev. Genet. 9, 605–618. https://doi.org/10.1038/nrg2386. 1819–1830.
Kelly, S., Lamb, D., Kelly, D., Manning, N., Loeffler, J., Hebart, H., Schumacher, U., Moudgal, V., Little, T., Boikov, D., Vazquez, J.A., 2004. Multiechinocandin- and multi-
Einsele, H., 1997. Resistance to fluconazole and cross-resistance to amphotericin B in azole-resistant Candida parapsilosis isolates serially obtained during therapy for
Candida albicans from AIDS patients caused by defective sterol Δ5,6 -desaturation. prosthetic valve endocarditis. Antimicrob. Agents Chemother. 49, 767–769. https://
FEBS Lett. 400, 80–82. https://doi.org/10.1016/s0014-5793(96)01360-9. doi.org/10.1128/AAC.49.2.767.
Kelly, S.L., Lamb, D.C., Corran, A.J., Baldwin, B.C., Kelly, D.E., 1995. Mode of action and Mukherjee, P.K., Leidich, S.D., Isham, N., Leitner, I., Ryder, N.S., Ghannoum, M.A., 2003.
resistance to azole antifungals associated with the formation of 14 alpha-methy- Clinical Trichophyton rubrum strain exhibiting primary resistance to terbinafine.
lergosta-8,24(28)-dien-3 beta,6 alpha-diol. Biochem. Biophys. Res. Commun. https:// Antimicrob. Agents Chemother. 47, 82–86. https://doi.org/10.1128/AAC.47.1.82.
doi.org/10.1006/bbrc.1995.1272. Müller, F.M.C., Seidler, M., Beauvais, A., 2011. Aspergillus fumigatus biofilms in the
Khan, A., El-Charabaty, E., El-Sayegh, S., 2015. Fungal infection in renal transplant pa- clinical setting. Med. Mycol. 49, 96–100. https://doi.org/10.3109/13693786.2010.
tients. Med. Mycol. Curr. Trends Futur. Prospect. 7, 110–146. https://doi.org/10. 502190.

11
M.W.J. Hokken, et al. Fungal Genetics and Biology 132 (2019) 103254

Munro, C.A., Selvaggini, S., Bruijn, I De, Walker, L., Lenardon, M.D., Gerssen, B., Milne, Salzer, H.J.F., Burchard, G., Cornely, O.A., Lange, C., Rolling, T., Schmiedel, S., Libman,
S., Brown, A.J.P., Gow, N.A.R., 2007. The PKC, HOG and Ca2+ signalling pathways M., Capone, D., Le, T., Dalcolmo, M.P., Heyckendorf, J., 2018. Diagnosis and man-
co-ordinately regulate chitin synthesis in Candida albicans. Mol. Microbiol. 63, agement of systemic endemic mycoses causing pulmonary disease. Respiration 96,
1399–1413. https://doi.org/10.1111/j.1365-2958.2007.05588.x. 283–301. https://doi.org/10.1159/000489501.
Nelson, M.R., Fisher, M., Cartledge, J., Rogers, T., Gazzard, B.G., 1994. The role of azoles Sandherr, M., Maschmeyer, G., 2011. Pharmacology and metabolism of voriconazole and
in the treatment and prophylaxis of cryptococcal disease in HIV infection. AIDS 8, posaconazole in the treatment of invasive aspergillosis – review of the literature. Eur.
651–654. https://doi.org/10.1097/00002030-199405000-00011. J. Med. Res. 139–144.
Nguyen, M., Ekstrom, A., Li, X., Yin, Y., 2015. HGT-finder: a new tool for horizontal gene Sanglard, D., 2019. Finding the needle in a haystack: Mapping antifungal drug resistance
transfer finding and application to Aspergillus genomes. Toxins (Basel) 7, in fungal pathogen by genomic approaches. PLoS Pathogens 1–9.
4035–4053. https://doi.org/10.3390/toxins7104035. Sanglard, D., Ischer, F., Bille, J., 2001. Role of ATP-binding-cassette transporter genes in
Ni, M., Feretzaki, M., Li, W., Floyd-Averette, A., Mieczkowski, P., Dietrich, F.S., Heitman, high-frequency acquisition of resistance to azole antifungals in Candida glabrata.
J., 2013. Unisexual and heterosexual meiotic reproduction generate aneuploidy and Antimicrob. Agents Chemother. 45, 1174–1183. https://doi.org/10.1128/AAC.45.4.
phenotypic diversity de novo in the yeast cryptococcus neoformans. PLoS Biol. 11. 1174-1183.2001.
https://doi.org/10.1371/journal.pbio.1001653. Sanglard, D., Ischer, F., Monod, M., Bille, J., 1997. Cloning of Candida albicans genes
Nicolas, A.G., Serero, A., Legoix-Né, P., Jubin, C., Loeillet, S., 2014. Mutational landscape conferring resistance to azole antifungal agents: Characterization of CDR2, a new
of yeast mutator strains. Proc. Natl. Acad. Sci. 111, 1897–1902. https://doi.org/10. multidrug ABC transporter gene. Microbiology 143, 405–416. https://doi.org/10.
1073/pnas.1314423111. 1099/00221287-143-2-405.
Oromendia, A.B., Dodgson, S., Amon, A., 2008. Aneuploidy causes proteotoxic stress in Sanguinetti, M., Posteraro, B., Fiori, B., Ranno, S., Torelli, R., Fadda, G., 2005. Unknown –
yeast. Genetics 26, 2696–2708. https://doi.org/10.1101/gad.207407.112.Compton. Unknown – Zał_3B_TABELA PARAMETROW.pdf.pdf. Society 49, 668–679. https://
Osborne, C.S., Leitner, I., Favre, B., Neil, S., Osborne, C.S., Leitner, I., Favre, B., Ryder, doi.org/10.1128/AAC.49.2.668.
N.S., 2005. Amino acid substitution in trichophyton rubrum squalene epoxidase as- Sawaya, B.P., Briggs, J.P., Schnermann, J., 1995. Amphotericin B nephrotoxicity: the
sociated with resistance to terbinafine amino acid substitution in trichophyton ru- adverse consequences of altered membrane properties. J. Am. Soc. Nephrol. 6,
brum squalene epoxidase associated with resistance to terbinafine. Antimicrob. 154–164.
Agents Chemother. 49, 2840–2844. https://doi.org/10.1128/AAC.49.7.2840. Schlichting, C.D., Piglucci, M., 1993. Control of phenotypic plasticity via regulatory
Pais, P., Pires, C., Costa, C., Okamoto, M., Chibana, H., Teixeira, M.C., 2016. Membrane genes. Am. Nat. 142, 366–370.
proteomics analysis of the Candida glabrata response to 5-flucytosine: Unveiling the Schlichting, C.D., Smith, H., 2002. Phenotypic plasticity: linking molecular mechanisms
role and regulation of the drug efflux transporters CgFlr1 and CgFlr2. Front. with evolutionary outcomes. Evol. Ecol. 16, 189–211.
Microbiol. 7, 1–14. https://doi.org/10.3389/fmicb.2016.02045. Schmiedel, Y., Zimmerli, S., 2016. Common invasive fungal diseases: an overview of
Park, S., Kelly, R., Kahn, J.N., Robles, J., Hsu, M.J., Register, E., Li, W., Vyas, V., Fan, H., invasive candidiasis, aspergillosis, cryptococcosis, and Pneumocystis pneumonia.
Abruzzo, G., Flattery, A., Gill, C., Chrebet, G., Parent, S.A., Kurtz, M., Teppler, H., Swiss Med. Weekly 1–12. https://doi.org/10.4414/smw.2016.14281.
Douglas, C.M., Perlin, D.S., 2005. Specific substitutions in the echinocandin target Schoofs, A., Odds, F.C., Colebunders, R., Ieven, M., Wouters, L., Goossens, H., 1997.
Fks1p account for reduced susceptibility of rare laboratory and clinical Candida sp. Isolation of Candida species on media with and without added fluconazole reveals
isolates. Antimicrob. Agents Chemother. 49, 3264–3273. https://doi.org/10.1128/ high variability in relative growth susceptibility phenotypes. Antimicrob. Agents
AAC.49.8.3264-3273.2005. Chemother. 41, 1625–1635. https://doi.org/10.1128/aac.41.8.1625.
Paul, S., Diekema, D., Moye-Rowley, W.S., 2017. Contributions of both ATP-binding Schoustra, S., Rundle, H.D., Dali, R., Kassen, R., 2010. Fitness-associated sexual re-
cassette transporter and Cyp51A proteins are essential for azole resistance in production in a filamentous fungus. Curr. Biol. 20, 1350–1355. https://doi.org/10.
Aspergillus fumigatus. Antimicrob. Agents Chemother. 61, 1–10. 1016/j.cub.2010.05.060.
Paul, S., Moye-Rowley, W.S., 2014. Multidrug resistance in fungi: Regulation of trans- Schubert, S., Barker, K.S., Znaidi, S., Schneider, S., Dierolf, F., Dunkel, N., Aïd, M.,
porter-encoding gene expression. Front. Physiol. 5 (APR), 1–14. https://doi.org/10. Boucher, G., Rogers, P.D., Raymond, M., Morschhäuser, J., 2011. Regulation of efflux
3389/fphys.2014.00143. pump expression and drug resistance by the transcription factors Mrr1, Upc2, and
Perfect, J.R., 2018. The antifungal pipeline: a reality check. Nat. Rev. Drug Discovery 16, Cap1 in Candida albicans. Antimicrob. Agents Chemother. 55, 2212–2223. https://
603–616. https://doi.org/10.1038/nrd.2017.46.The. doi.org/10.1128/aac.01343-10.
Perfect, J.R., Cox, G.M., 1999. Drug resistance in Cryptococcus neoformans. Drug Resist. Selmecki, A., Forche, A., Berman, J., 2006. Aneuploidy and isochromosome formation in
Updat. 2, 259–269. https://doi.org/10.1054/drup.1999.0090. drug-resistant Candida albicans. Strain 313, 367–370.
Perlin, D.S., 2007. Resistance to echinocandin-class antifungal drugs. Drug Resist. Updat. Selmecki, A.M., Dulmage, K., Cowen, L.E., Anderson, J.B., Berman, J., 2009. Acquisition
https://doi.org/10.1016/j.drup.2007.04.002.Resistance. of aneuploidy provides increased fitness during the evolution of antifungal drug re-
Perlin, D.S., Shor, E., Zhao, Y., 2015. Update on antifungal drug resistance. Curr. Clin. sistance. PLoS Genet. 5, 1–16. https://doi.org/10.1371/journal.pgen.1000705.
Microbiol. Rep. 2, 84–95. https://doi.org/10.1007/s40588-015-0015-1. Shields, R.K., Kline, E.G., Healey, K.R., Kordalewska, M., Perlin, D.S., Nguyen, M.H.,
Perlin, M.H., Andrews, J., Toh, S.S., 2014. Essential letters in the fungal alphabet: ABC Clancy, C.J., 2019. Spontaneous mutational frequency and fks mutation rates vary by
and MFS transporters and their roles in survival and pathogenicity. Adv. Genet. echinocandin agent against Candida glabrata. Antimicrob. Agents Chemother. 1–9.
https://doi.org/10.1016/B978-0-12-800271-1.00004-4. Shoham, S., Marr, K.A., 2014. Invasive fungal infections in solid organ transplant re-
Pilmis, B., Garcia-Hermoso, D., Alanio, A., Catherinot, E., Scemla, A., Jullien, V., cipients. Future Microbiol. 71, 3831–3840. https://doi.org/10.1158/0008-5472.
Bretagne, S., Lortholary, O., 2018. Failure of voriconazole therapy due to acquired CAN-10-4002.BONE.
azole resistance in Aspergillus fumigatus in a kidney transplant recipient with chronic Singh, A., Masih, A., Khurana, A., Kumar, P., Gupta, M., Hagen, F., Meis, J.F., Chowdhary,
necrotizing aspergillosis. Am. J. Transplant. 18, 2352–2355. https://doi.org/10. A., 2018. High terbinafine resistance in Trichophyton interdigitale isolates in Delhi
1111/ajt.14940. India harbouring mutations in the squalene epoxidase gene. Mycoses 477–484.
Pontecorvo, G., Roper, J.A., Chemmons, L.M., Macdonald, K.D., Bufton, A.W.J., 1953. https://doi.org/10.1111/myc.12772.
The genetics of Aspergillus nidulans. Adv. Genet. 5, 141–238. https://doi.org/10. Sionov, E., Lee, H., Chang, Y.C., Kwon-Chung, K.J., 2010. Cryptococcus neoformans
1016/S0065-2660(08)60408-3. overcomes stress of azole drugs by formation of disomy in specific multiple chro-
du Pré, S., et al., 2018. Effect of the Novel Antifungal Drug F901318 (Olorofim) on mosomes. PLoS Pathog. 6, e1000848. https://doi.org/10.1371/journal.ppat.
Growth and Viability of Aspergillus fumigatus. Antimicrobial agents and 1000848.
Chemotherapy 62 (8). https://doi.org/10.1128/AAC.00231-18. Skelly, D.A., Magwene, P.M., Meeks, B., Murphy, H.A., Murphy, H.A., 2017. Known
Prigent, G., Nawel, A., Levesque, E., Fekkar, A., Costa, J., Anbassi, S. El, Duvoux, C., mutator alleles do not markedly increase mutation rate in clinical Saccharomyces
Merle, J., Dannaoui, E., Botterel, F., 2017. Echinocandin Resistance in Candida cerevisiae strains. Proc. R. Soc. B: Biol. Sci. 284 (1852).
Species isolates from Liver Transplant recipients. Antimicrob. Agents Chemother. 61, Smith, K.J., Warnock, D.W., Kennedy, C.T.C., Johnson, E.M., Hopwood, V., van Cutsem,
1–10. J., Vanden Bossche, H., 1986. Azole resistance in candida albicans. Med. Mycol. 24,
Rajendran, R., May, A., Sherry, L., Kean, R., Williams, C., Jones, B.L., Burgess, K.V., 133–144. https://doi.org/10.1080/02681218680000201.
Heringa, J., Abeln, S., Brandt, B.W., Munro, C.A., Ramage, G., 2016. Integrating Snelders, E., Camps, S.M.T., Karawajczyk, A., Schaftenaar, G., Kema, G.H.J., van der Lee,
Candida albicans metabolism with biofilm heterogeneity by transcriptome mapping. H.A., Klaassen, C.H., Melchers, W.J.G., Verweij, P.E., 2012. Triazole fungicides can
Sci. Rep. 6, 1–11. https://doi.org/10.1038/srep35436. induce cross-resistance to medical triazoles in Aspergillus fumigatus. PLoS One 7.
Ramage, G., Rajendran, R., Sherry, L., Williams, C., 2012. Fungal biofilm resistance. Int. https://doi.org/10.1371/journal.pone.0031801.
J. Microbiol. https://doi.org/10.1155/2012/528521. Snelders, E., Rijs, A.J., Kema, G.H., Melchers, W.J., Verweij, P.E., 2009. Possible en-
Renne, M.F., de Kroon, A.I.P.M., 2018. The role of phospholipid molecular species in vironmental origin of resistance of Aspergillus fumigatus to medical triazoles. Appl.
determining the physical properties of yeast membranes. FEBS Lett. 592, 1330–1345. Environ. Microbiol. 75, 4053–4057. https://doi.org/10.1128/AEM.00231-09.
https://doi.org/10.1002/1873-3468.12944. Snelders, E., Van Der Lee, H.A.L., Kuijpers, J., Rijs, A.J.M.M., Varga, J., Samson, R.A.,
Richardson, M.D., 1997. Effect of lamisil and azole antifungals in experimental nail in- Mellado, E., Donders, A.R.T., Melchers, W.J.G., Verweij, P.E., 2008. Emergence of
fection. Dermatology 302. azole resistance in Aspergillus fumigatus and spread of a single resistance me-
Rocha, E.M.F., Gardiner, R.E., Park, S., Martinez-Rossi, N.M., Perlin, D.S., 2006. A chanism. PLoS Med. 5, 1629–1637. https://doi.org/10.1371/journal.pmed.0050219.
Phe389Leu substitution in ErgA confers terbinafine resistance in Aspergillus fumi- Song, J., Zhang, S., Lu, L., 2018. Fungal cytochrome P450 protein Cyp51: What we can
gatus. Antimicrob. Agents Chemother. 50, 2533–2536. https://doi.org/10.1128/ learn from its evolution, regulons and Cyp51-based azole resistance. Fungal Biol. Rev.
AAC.00187-06. 32, 131–142. https://doi.org/10.1016/j.fbr.2018.05.001.
Rodero, L., Mellado, E., Rodriguez, A.C., Salve, A., Guelfand, L., Cahn, P., Cuenca-Estrella, Sorrell, T., Chen, S., 2007. Impact of antifungal resistance in Australia. Microbiol. Aust.
M., Davel, G., Rodriguez-Tudela, J.L., 2003. G484S amino acid substitution in la- 171–175.
nosterol 14-α demethylase (ERG11) is related to fluconazole resistance in a recurrent Stern, S., Dror, T., Stolovicki, E., Brenner, N., Braun, E., 2007. Genome-wide transcrip-
cryptococcus neoformans clinical isolate. Antimicrob. Agents Chemother. 47, tional plasticity underlies cellular adaptation to novel challenge. Mol. Syst. Biol. 3,
3653–3656. https://doi.org/10.1128/AAC.47.11.3653-3656.2003. 1–9. https://doi.org/10.1038/msb4100147.

12
M.W.J. Hokken, et al. Fungal Genetics and Biology 132 (2019) 103254

Tang, L., Yang, X.F., Qiao, M., Zhang, L., Tang, X.W., Qiu, H.Y., Wu, D.P., Sun, A.N., 2018. Webb, B.J.K., Ferraro, J.P., Rea, S., Kaufusi, S., Goodman, B.E., Spalding, J., 2018. epi-
Posaconazole vs. voriconazole in the prevention of invasive fungal diseases in pa- demiology and clinical features of invasive fungal infection in a US Health Care
tients with haematological malignancies: A retrospective study. J. Mycol. Med. 28, Network. Open Forum Infect. Dis. 5, 2–9. https://doi.org/10.1093/ofid/ofy187.
379–383. https://doi.org/10.1016/j.mycmed.2017.11.003. White, S.J., Rosenbach, A., Lephart, P., Nguyen, D., Benjamin, A., Tzipori, S., Whiteway,
Vale-Silva, L.A., Coste, A.T., Ischer, F., Parker, J.E., Kelly, S.L., Pinto, E., Sanglard, D., M., Mecsas, J., Kumamoto, C.A., 2007. Self-regulation of Candida albicans population
2012. Azole resistance by loss of function of the sterol Δ5,6- desaturase gene (ERG3) size during GI colonization. PLoS Pathog. 3, 1866–1878. https://doi.org/10.1371/
in Candida albicans does not necessarily decrease virulence. Antimicrob. Agents journal.ppat.0030184.
Chemother. 56, 1960–1968. https://doi.org/10.1128/AAC.05720-11. White, T.C., 1997. Increased mRNA levels of ERG16, CDR, and MDR1 correlate, with
Van Der Linden, J.W.M., Camps, S.M.T., Kampinga, G.A., Arends, J.P.A., Debets- increases in azole resistance in Candida albicans isolates from a patient infected with
Ossenkopp, Y.J., Haas, P.J.A., Rijnders, B.J.A., Kuijper, E.J., Van Tiel, F.H., Varga, J., human immunodeficiency virus. Antimicrob. Agents Chemother. 41, 1482–1487.
Karawajczyk, A., Zoll, J., Melchers, W.J.G., Verweij, P.E., 2013. Aspergillosis due to Wiederhold, N.P., 2017. Antifungal resistance: current trends and future strategies to
voriconazole highly resistant Aspergillus fumigatus and recovery of genetically re- combat. Infect. Drug Resist. 10, 249–259. https://doi.org/10.2147/IDR.S124918.
lated resistant isolates from domiciles. Clin. Infect. Dis. 57, 513–520. https://doi.org/ Wyatt, T.T., Wösten, H.A.B., Dijksterhuis, J., 2013. Fungal spores for dispersion in space
10.1093/cid/cit320. and time. Adv. Appl. Microbiol. https://doi.org/10.1016/B978-0-12-407672-3.
Vermes, A., Guchelaar, H.-J., Dankert, J., 2000. Flucytosine: a review of its pharma- 00002-2. 1st ed., Elsevier Inc.
cology, clinical indications, pharmacokinetics, toxicity and drug interactions. J. Yamada, T., Maeda, M., Alshahni, M.M., Tanaka, R., Yaguchi, T., Bontems, O., Salamin,
Antimicrob. Chemother. 171–179. K., Fratti, M., Monod, M., 2017. Terbinafine resistance of trichophyton clinical iso-
Vermeulen, E., Lagrou, K., Verweij, P.E., 2013. Azole resistance in Aspergillus fumigatus: lates caused by specific point mutations in the squalene epoxidase gene. Antimicrob.
a growing public health concern. Curr. Opin. Infect. Dis. 26, 493–500. https://doi. Agents Chemother. 61, 1–13. https://doi.org/10.1128/aac.00115-17.
org/10.1097/QCO.0000000000000005. Yang, F., Zhang, L., Wakabayashi, H., Myers, J., Jiang, Y., Cao, Y., Jimenez-Ortigosa, C.,
Verweij, P.E., Lestrade, P.P.A., Melchers, W.J.G., Meis, J.F., 2016a. Azole resistance Perlin, D.S., Rustchenko, E., 2017. Tolerance to caspofungin in Candida albicans is
surveillance in Aspergillus fumigatus: Beneficial or biased? J. Antimicrob. associated with at least three distinctive mechanisms that govern expression of FKS
Chemother. 71, 2079–2082. https://doi.org/10.1093/jac/dkw259. genes and cell wall remodelling. Antimicrob. Agents Chemother. 61, 1–13.
Verweij, P.E., Snelders, E., Kema, G.H., Mellado, E., Melchers, W.J., 2009. Azole re- Young, L.Y., Hull, C.M., Heitman, J., 2003. Disruption of ergosterol biosynthesis confers
sistance in Aspergillus fumigatus: a side-effect of environmental fungicide use? resistance to amphotericin B in Candida lusitaniae. Antimicrob. Agents Chemother.
Lancet Infect. Dis. 9, 789–795. https://doi.org/10.1016/S1473-3099(09)70265-8. 47, 2717–2724. https://doi.org/10.1128/AAC.47.9.2717-2724.2003.
Verweij, P.E., Voss, A., Meis, J.F., 1998. Resistance of aspergillus fumigatus to itraco- Yue-bin, Z., Yuan-shu, Q., Hong-ni, G.U., Epidemiological, S., 2007. Genome-wide ex-
nazole. Scand. J. Infect. Dis. 30, 642–643. pression profiling of the response to terbinafine in Candida albicans using a cDNA
Verweij, P.E., Zhang, J., Debets, A.J.M., Meis, J.F., van de Veerdonk, F.L., Schoustra, S.E., microarray analysis. Chin. Med. J. (Engl.) 120, 807–813.
Zwaan, B.J., Melchers, W.J.G., 2016b. In-host adaptation and acquired triazole re- Zhang, J., Snelders, E., Zwaan, B.J., Schoustra, S.E., Meis, J.F., Van Dijk, K., Hagen, F.,
sistance in Aspergillus fumigatus: a dilemma for clinical management. Lancet. Infect. Van Der Beek, M.T., Kampinga, G.A., Zoll, J., Melchers, W.J.G., Verweij, P.E., Debets,
Dis. 16, e251–e260. https://doi.org/10.1016/S1473-3099(16)30138-4. A.J.M., 2017a. A novel environmental azole resistance mutation in Aspergillus fu-
Vincent, B.M., Lancaster, A.K., Scherz-Shouval, R., Whitesell, L., Lindquist, S., 2013. migatus and a possible role of sexual reproduction in its emergence. MBio 8, 1–13.
Fitness Trade-offs Restrict the Evolution of Resistance to Amphotericin B. PLoS Biol. https://doi.org/10.1128/mBio. 00791-17.
11, e1001692. https://doi.org/10.1371/journal.pbio.1001692. Zhang, J., Snelders, E.E., Zwaan, B.J., Schoustra, S.E., Kuijper, E.J., Arendrup, M.C.,
Vu, K., Thompson, G.R., Roe, C.C., Sykes, J.E., Dreibe, E.M., Lockhart, S.R., Meyer, W., Melchers, W.J.G., Verweij, P.E., Debets, A.J.M., 2019. Relevance of heterokaryosis
Engelthaler, D.M., Gelli, A., 2018. Flucytosine resistance in Cryptococcus gattii is for adaptation and azole-resistance development in Aspergillus fumigatus. Proc. R.
indirectly mediated by the FCY2-FCY1-FUR1 pathway. Med. Mycol. 56, 857–867. Soc. B Biol. Sci. 286. https://doi.org/10.1098/rspb.2018.2886.
https://doi.org/10.1093/mmy/myx135. Zhang, J., Van Den Heuvel, J., Debets, A.J.M., Verweij, P.E., Melchers, W.J.G., Zwaan,
Walker, L.A., Gow, N.A.R., Munro, C.A., 2010. Fungal echinocandin resistance. Fungal B.J., Schoustra, S.E., 2017b. Evolution of cross-resistance to medical triazoles in
Genet. Biol. 47, 117–126. https://doi.org/10.1016/j.fgb.2009.09.003. Aspergillus fumigatus through selection pressure of environmental fungicides. Proc.
Walker, L.A., Lee, K.K., Munro, C.A., Gow, N.A.R., 2015. Caspofungin treatment of R. Soc. B Biol. Sci. 284. https://doi.org/10.1098/rspb.2017.0635.
Aspergillus fumigatus results in ChsG-dependent upregulation of chitin synthesis and Zhang, N., Magee, B.B., Magee, P.T., Holland, B.R., Rodrigues, E., Holmes, A.R., Cannon,
the formation of chitin-rich microcolonies. Antimicrob. Agents Chemother. 59, R.D., Schmid, J., 2015. Selective advantages of a parasexual cycle for the yeast
5932–5941. https://doi.org/10.1128/AAC.00862-15. Candida albicans. Genetics 200, 1117–1132. https://doi.org/10.1534/genetics.115.
Walsh, T.J., Petraitis, V., Petraitiene, R., Field-Ridley, A., Sutton, D., Ghannoum, M., Sein, 177170.
T., Schaufele, R., Peter, J., Bacher, J., Casler, H., Armstrong, D., Espinel-Ingroff, A., Zhao, H.Z., Wang, R.Y., Wang, X., Jiang, Y.K., Zhou, L.H., Cheng, J.H., Huang, L.P.,
Rinaldi, M.G., Lyman, C.A., 2003. Experimental pulmonary aspergillosis due to as- Harrison, T.S., Zhu, L.P., 2018. High dose fluconazole in salvage therapy for HIV-
pergillus terreus: pathogenesis and treatment of an emerging fungal pathogen re- uninfected cryptococcal meningitis. BMC Infect. Dis. 18, 1–8. https://doi.org/10.
sistant to amphotericin B. J. Infect. Dis. 188, 305–319. https://doi.org/10.1086/ 1186/s12879-018-3460-7.
377210.

13

You might also like