Zacytuj

Introduction

Aspergillus flavus is a filamentous fungus, with worldwide distribution (Hedayati et al. 2007). After Aspergillus fumigatus this species is the second most important causative agent of invasive and non-invasive aspergillosis in patients (Baranyi et al. 2017). Three classes of antifungal agents form current treatments of invasive aspergillosis (IA): polyenes (amphotericin B (AMB)), azoles (voriconazole (VOR), itraconazole (IT), posaconazole (POS)) and more recently semi-synthetic echinocandins (caspofungin, micafungin, rezafungin and anidulafungin). Among these, azoles are the first choice drugs in managing and prophylaxis of aspergillosis (Patterson et al. 2016). The resistance to azoles has increased, especially in Aspergillus species. The overall azole resistance rate of A. fumigatus could reach 27.8%, according to Vermeulen et al. (2013). Prevalence rate variation between studies could be due to regional and disease differences (Vermeulen et al. 2013). Recently, Rhodes et al. (2022) explored 218 isolates collected from patients with different aspergillosis manifestations and from the environment. These A. fumigatus isolates were collected from different hospital centers (England, Wales, Scotland, and Ireland). They found that 49% showed resistance to at least one of the tested antifungal drugs. For specific azoles, 52% (n = 104) exceeded minimal inhibitory concentration (MIC) breakpoints to itraconazole (≥ 2 mg/l), 31% (n = 64) to voriconazole (≥ 2 mg/l), and 25% (n=44) to posaconazole (≥ 0.5 mg/l). They also found that 26 isolates (12%) exceeded MIC breakpoints to two or more azole drugs from both clinical (n = 23) and environmental (n = 3) sources (Rhodes et al. 2022). Shelton et al. (2023) explored 1,894 air samples collected from England, Wales, Scotland, and Northern Ireland. They identified 99 TEB-resistant A. fumigatus isolates. These isolates were resistant to itraconazole (86%), to voriconazole (64%), to posaconazole (18%) and to isavuconazole (83%). Thus, there was clear cross-resistance between the agrochemical TEB and the medical azole antifungals (Shelton et al 2023).

Knowing that antifungal susceptibility testing is not routine testing in many centers around the world, the azole resistance in A. flavus still needs to be fully explored. In addition, the molecular mechanism of azole resistance concerning this species of filamentous mold has yet to be fully explained.

Several studies have shown that azoles inhibit the ergosterol biosynthetic pathway by interacting with the two CYP51 proteins (CYP51A and CYP51B) and alter the metabolism of sterols in cells (Alcazar-Fuoli et al. 2008). However, a risk of decisive infection may occur if resistance is acquired (Howard and Arendrup 2011; Hadrich et al. 2012a). Various studies focus on azole-resistant Aspergillus species, with a particular interest in A. fumigatus isolates (Verweij et al. 2016; Chowdhary et al. 2017). Studying molecular mechanisms of resistance is an essential step and complementary to the in vitro identification of antifungal resistance (Buil et al. 2018). Over the last decade, there has been some progress in identifying genes responsible for antifungal resistance in A. fumigatus, but studies have been very rare in A. flavus (Hagiwara et al. 2016). A duplication of several genes (cyp51A and cyp51B) involved in the sterol biosynthesis pathway was noted in A. fumigatus (Mellado et al. 2001; Da Silva Ferreira et al. 2005; Nierman et al. 2005). In A. flavus, there are three CYP51 proteins: CYP51A, CYP51B, and CYP51C. Recently, rare studies have analyzed genetic alterations and/or overexpression of azole target genes and efflux pump genes (EPs) that reduce intracellular drug concentrations in azole-resistant isolates (Krishnan-Natesan et al. 2008; Liu et al. 2012; Paul et al. 2015). Furukawa et al. (2020) constructed a 484-member transcription factor null mutant library to provide a systematic evaluation of regulators that contribute to azole resistance in A. fumigatus. They identified 12 regulators that have a demonstrable role in itraconazole susceptibility. The loss of the negative cofactor 2 complex leads to resistance to itraconazole, amphotericin B, and terbinafine.

This study evaluated the mRNA expression levels of cyp51 (A and B) genes in clinical Aspergillus flavus isolates determined by RT-qPCR. A PCR sequencing of cyp51A and cyp51B genes was used to detect the presence of gene mutations.

Experimental
Materials and Methods
Patients and isolates

The 34 molecularly confirmed A. flavus sensu stricto clinical isolates included in the present study were collected from 14 patients hospitalized in the hematology unit at Hedi-Chaker Hospital, Sfax, Tunisia. Weekly samples (septum and nasal) were collected from immunocompromised patients with invasive aspergillosis. Eighteen isolates were cultured from sputum, ten from nasal swabs, five from bronchoalveolar lavage, and one from a pulmonary biopsy. All A. flavus strains were cultured on Sabouraud dextrose agar plates (SDA; AES™, bioMérieux, France). These A. flavus strains were identified based on morphological features and DNA sequences of the rRNA gene internal transcribed spaced regions ITS (de Hoog et al. 2007).

Antifungal resistance susceptibility by ETEST® method

The ETEST® strips (AES™, bioMérieux, France) were used as previously described to determine antifungal susceptibility profiles after 48 h of incubation (Hadrich et al. 2012b). Epidemiological cutoff values (ECVs) for POS (0.25 μg/ml), IT, and VOR (1 μg/ml) were previously described by five laboratories, according to the CLSI M38-A2 microdilution method (Espinel-Ingroff and Turnidge 2016).

Mechanisms of azole resistance. DNA and RNA extractions

DNA was extracted using a QIAamp® DNA Mini Kit (QIAGEN, Germany), as indicated by the manufacturer’s instructions.

The RNeasy® Mini Kit (QIAGEN, Germany) was used according to the manufacturer’s instructions to extract total RNA from the cell lysates. To avoid DNA contamination, RNA extracts were treated with RNase-free DNase (Promega, UK). The concentration and purity of the RNA were determined using a UV spectrophotometer by measuring the absorbance at 230 (A230), 260 (A260), and 280 nm (A280). The A260 nm/A280 nm ratio of the samples, reflecting the purity ranged from 2.06 to 2.21. The A260 nm/A230 nm ratio ranged from 1.90 to 2.50.

First strand cDNA synthesis

The transformation of the RNA extract from each isolate into cDNA was done by reverse transcription method (RT), according to the manufacturer’s instructions of the PrimeScript™ RT Reagent Kit (Perfect Real Time) from Takara Bio Inc. (Japan).

Quantitative real-time PCR (qPCR)

Specific primers and probes for gene expression analysis were designed using Primer3 software (https://bioinfo.ut.ee/primer3-0.4.0) (Koressaar and Remm 2007; Untergasser et al. 2012) and verified by sequence manipulation suite (https://www.bioinformatics.org/sms2) (Stothard 2000) (Table I).

The sequences of primers and probes used in RT-qPCR.

Gene Primers and probes
cyp51A F 5’-GCGCGCATGAGGGAGAT-3’
R 5’-CAATGCATGAGGTTCCAGATCA-3’
Probe HEX-TCATTAACGAGCGCCGCAAGAACC-MGB
cyp51B F 5’-ATTCGACTCGACATTTGCTGAA-3’
R 5’-GCATCACGCTTGCGGTTAT-3’
Probe FAM-CATGATCTCGACATGGGTTTTGCCC-MGB
ANXC4 F 5’-CCAACCCATAAACGCTCTGT-3’
R 5’-TGGTGGGAATCTTGGAGAAC-3’
Probe CY5-ATCGAAGCAGCCTGTCTCAT-MGB

The measurement of the expression level and the number of copies of each of the two genes cyp51 A and cyp51 B was performed by qPCR quantitative time PCR. In this work, the normalization was done in relation to housekeeping gene Annexin (ANXC4).

Gene expression was analyzed using 10 μl TaqMan™ Universal PCR Master Mix (Applied Biosystems™; Thermo Fisher Scientific, Inc., USA) and 1 μl template (DNA or cDNA) with 20 pmol forward and reverse primers, 7 pmol hydrolysis probe. The PCR amplification conditions were as follows. The first step is 50°C for 2 minutes, 95°C for 10 minutes, followed by 45 cycles of 95°C for 15 seconds, and the final step is 54°C for 1 minute. All reactions were performed in triplicate using the StepOne™ Real-Time PCR instrument (Applied Biosystems™; Thermo Fisher Scientific, Inc., USA). StepOne™ software version 2.1 was used to collect Cq data and calculate relative quantification (RQ). We then used the published comparative 2−ΔΔCq method to normalize the folding changes in target gene expression to the reference gene according to the following formula: RQ=2(CqTargetCqReference)Tested(CqTargetCqReference)Control $$\matrix{ {RQ = 2} \hfill & - \hfill & {\,{\rm{(}}CqTarget\, - \,Cq\,Reference{\rm{)}}\,Tested\,} \hfill \cr {} \hfill & - \hfill & {\,{\rm{(}}Cq\,Target\, - \,Cq\,Reference{\rm{)}}Control\,} \hfill \cr } $$

A 2.5-fold change was considered gene overexpression or an increase in gene copy number (Livak and Schmittgen 2001; VanGuilder et al. 2008).

Sequencing f PCR products

The A. flavus cyp51A gene, namely, cyp51A (NCBI accession number XM_002375082.1) and cyp51B (NCBI accession number XM_002379089.1) were obtained from the nucleotide databank NCBI BLAST (http://www.ncbi.nlm.nih.gov). Thus, we divided the gene of cyp51A into three overlapping fragments (cyp51AP1, cyp51AP2, and cyp51AP3) and the same for cyp51B (cyp51BP1, cyp51BP2, and cyp51BP3). According to these sequences, the primers for PCR sequencing assays were designed using Primer (version 3) software (https://frodo.wi.mit.edu) (Table II).

The sequences of primers used in PCR sequencing.

Gene Primers
cyp51A AP1F 5’-ATGGCATCCTTCACTCTCGT-3’
AP1R 5’-CG ATCAACTTCATGCTTCCG-3’
AP2F 5’-TCTGGAACCTCATGCATTGT-3’
AP2R 5’-TCCCTCGAAACCAGCAATTA-3’
AP3F 5’-GGCAGGGTCGAAATCACGGA-3’
AP3R 5’-GCCCGGATGGAAGAACCCTT-3’
cyp51B BP1F 5’-CTTTTATCGGAAGTACCATC-3’
BP1R 5’-CTTTATCAGGAACAACTTCG-3’
BP2F 5’-ACTCTTCGCATACATGCACC-3’
BP2R 5’-CCAACGTGCATGATATTGCC-3’
BP3F 5’-CGACTCGACATTTGCTGAAC-3’
BP3R 5’-CTCTTCGCATACATGCACCA-3’

Amplification reactions were performed with a final volume of 50 μl containing 10 μl 5 × reaction buffer (pH 8.5), 25 mM MgCl2, 0.2 mM (each) dATP, dCTP, dGTP and dTTP (Promega, UK), 20 pmol of each primer, 2.5 U of GOTaq® DNA Polymerase (Promega, UK) and 400 ng of genomic DNA. PCR was performed in a thermocycler (Eppendorf, Germany), with the following amplification conditions: 94°C for 5 min, followed by 30 cycles of 30 s at 94°C, 30 s at 58°C and 1 min at 72°C and a final extension at 72°C for 10 min. PCR products were analyzed by agarose gel electrophoresis and purified using the Purification PCR Wizard® Kit (Promega, UK). Both sequences of the triplicate PCR product were directly sequenced using the BigDye™ Terminator v3.1 Cycle Sequencing Kit (Applied Biosystems™; Thermo Fisher Scientific, Inc., USA) with primers and automated sequencer (ABI 3730; Applied Biosystems™; Thermo Fisher Scientific, Inc., USA), according to the manufacturer’s instructions. Product sequences were compared with those of wild-type A. flavus cyp51A and cyp51B sequences using the NCBI Align Sequence Nucleotide BLAST and ClustalW V2 tools (https://www.genome.jp/tools-bin/clustalw).

Homology modeling

The DNA sequences of cyp51A and cyp51 B genes for 11 strains of A. flavus (six resistant and five sensitive to azoles) were submitted to the NCBI GenBank database with the access number in Table III. Then, they were translated into the equivalent amino acid sequences by using the Geniegene software (http://acces.ens-lyon.fr/acces/logiciels/applications/geniegen). The BLAST server (NCBI, USA) was used for Protein Data Bank (PDB) similarity searches to find homologous sequences (models) that matched known experimental 3D structures (Altschul et al. 1997). To align the protein sequences, the Clustal Omega program server (EMBL-EBI, UK) was used (Sievers et al. 2011). The automated protein structure homologymodeling server SWISS-MODEL (SIB, Switzerland) has generated a homology model for each target protein (Biasini et al. 2014).

Antifungal susceptibility to the azoles, relative quantification of gene expression, gene copy number of cyp51A and cyp51B genes, and mutation in Aspergillus flavus.

Strain GenBank accession Isolation site Isolation date Itraconazole Posaconazole RNA relative quantification DNA relative quantification Cyp51A Cyp51B
MIC (μg/ml) R/S MIC (μg/ml) R/S cyp51A ±SD cyp51B ± SD cyp51A ±SD cyp51B ± SD GenBank accession Punctual mutation GenBank accession Punctual mutation
Patient 1 Invasive aspergillosis TN-1 MN964040 nasal 15/10/2018 0.125 S 0.125 S 0.90 ±0,02 1.16 ±0,04 0.69 ±0,05 0.57 ±0,02 ND ND ND ND
TN-2 MN964041 sputum 15/10/2018 0.125 S 0.125 S 2.15 ±0,03 0.73 ±0,02 0.89 ±0,02 1.98 ±0,04 ND ND ND ND
TN-3 MN964042 sputum 22/10/2018 0.125 S 0.125 S 0.58 ±0,01 0.90 ±0,02 1.37 ±0,02 0.97 ±0,02 ND ND ND ND
Patient 2 Invasive aspergillosis TN-4 MN964043 nasal 15/10/2018 0.125 S 0.19 S 1.53 ±0,02 1.05 ±0,04 1.27 ±0,03 1.93 ±0,02 ND ND ND ND
TN-5 MN964044 sputum 15/10/2018 0.125 S 0.064 S 0.51 ±0,04 0.05 ±0,02 1.05 ±0,02 1.92 ±0,06 ND ND ND ND
TN-6 MN964045 nasal 23/10/2018 0.032 S 0.094 S 0.95 ±0,01 0.89 ±0,02 1.69 ±0,03 0.93 ±0,02 ND ND ND ND
Patient 3 Invasive aspergillosis TN-7 MN964046 nasal 30/10/2018 1.5 R 0.125 S 1.65 ±0,02 0.73 ±0,03 2.51 ±0,02 1.98 ±0,02 MN964023 C-T*(183) MN964034 No mutation
TN-8 MN964047 sputum 30/10/2018 0.5 S 0.094 S 1.12 ±0,02 0.73 ±0,02 1.13 ±0,07 5.97 ±0,04 ND ND ND ND
Patient 4 Invasive aspergillosis TN-9 MN964048 sputum 30/10/2018 0.75 S 0.125 S 4.48 ±0,02 3.89 ±0,02 2.71 ±0,02 2.67 ±0,02 ND ND ND ND
TN-10 MN964049 nasal 30/10/2018 0.38 S 0.064 S 1.42 ±0,02 1.97 ±0,04 2.49 ±0,03 2.01 ±0,03 ND ND ND ND
Patient 5 Invasive aspergillosis TN-11 MN964050 sputum 10/11/2018 0.5 S 0.125 S 5.65 ±0,01 2.31 ±0,02 1.87 ±0,02 1.02 ±0,02 MN964018 No mutation MN964029 No mutation
TN-12 MN964051 BAL 17/11/2018 0.125 S 0.125 S 2.35 ±0,02 2.31 ±0,02 1.87 ±0,02 1.02 ±0,02 MN964019 C-T*(183) MN964030 No mutation
Patient 6 Invasive aspergillosis TN-13 MN964052 BAL 03/12/2018 0.75 S 0.19 S 5.68 ±0,02 2.35 ±0,05 2.83 ±0,05 3.95 ±0,02 ND ND ND ND
TN-14 MN964053 sputum 26/11/2018 0.5 S 0.125 S 0.63 ±0,03 0.88 ±0,06 6.95 ±0,07 1.06 ±0,02 ND ND ND ND
TN-15 MN964054 nasal 26/11/2018 1 R 0.19 S 0.90 ±0,02 0.76 ±0,02 0.53 ±0,02 1.65 ±0,05 MN964024 C-T*(183) MN964035 No mutation
TN-16 MN964055 sputum 10/11/2018 1 R 0.19 S 1.02 ±0,02 1.22 ±0,05 2.34 ±0,04 2.84 ±0,02 MN964025 C-T*(183) MN964036 No mutation
Patient 7 Invasive aspergillosis TN-17 MN964056 sputum 01/03/2017 0.75 S 0.125 S 0.48 ±0,02 1.10 ±0,02 0.77 ±0,02 2.07 ±0,06 ND ND ND ND
TN-18 MN964057 BAL 15/03/2017 0.5 S 0.125 S 2.18 ±0,03 1.81 ±0,02 2.11 ±0,02 2.17 ±0,02 ND ND ND ND
TN-19 MN964058 sputum 08/03/2017 0.75 S 0.125 S 0.92 ±0,02 0.76 ±0,02 0.57 ±0,02 1.11 ±0,02 ND ND ND ND
TN-20 MN964059 sputum 15/03/2017 0.38 S 0.064 S 2.27 ±0,02 2.88 ±0,05 0.64 ±0,04 1.09 ±0,02 ND ND ND ND
Patient 8 Invasive aspergillosis TN-21 MN964060 sputum 08/03/2017 0.38 S 0.125 S 0.45 ±0,02 0.88 ±0,02 6.87 ±0,02 1.82 ±0,02 MN964020 C-T*(183) MN964031 No mutation
Patient 9 Invasive aspergillosis TN-22 MN964061 sputum 01/03/2017 0.25 S 0.125 S 0.49 ±0,02 0.55 ±0,02 1.84 ±0,02 1.68 ±0,04 ND ND ND ND
Patient 10 Invasive aspergillosis TN-23 MN964062 sputum 15/03/2017 0.38 S 0.125 S 0.69 ±0,02 0.95 ±0,03 1.73 ±0,03 1.59 ±0,02 MN964021 C-T*(183) MN964032 No mutation
Patient 11 Invasive aspergillosis TN-24 MN964063 sputum 01/03/2017 0.38 S 0.125 S 0.51 ±0,02 0.87 ±0,02 1.62 ±0,02 1.66 ±0,09 ND ND ND ND
TN-25 MN964064 sputum 08/03/2017 0.75 S 0.125 S 0.81 ±0,02 0.75 ±0,01 1.62 ±0,01 1.48 ±0,02 ND ND ND ND
TN-26 MN964065 BAL 15/03/2017 0.75 S 0.125 S 2.21 ±0,02 2.51 ±0,02 0.67 ±0,02 1.42 ±0,02 ND ND ND ND
Patient 12 Invasive aspergillosis TN-27 MN964066 BAL 03/12/2018 0.25 S 0.125 S 0.65 ±0,02 0.78 ±0,04 5.95 ±0,03 1.03 ±0,08 ND ND ND ND
TN-28 MN964067 nasal 26/11/2018 0.25 S 0.125 S 1.73 ±0,02 0.88 ±0,02 1.95 ±0,02 1.06 ±0,02 ND ND ND ND
TN-29 MN964068 sputum 26/11/2018 0.5 S 0.094 S 2.11 ±0,02 0.48 ±0,02 4.39 ±0,02 1.95 ±0,02 MN964022 No mutation MN964033 No mutation
Patient 13 Invasive aspergillosis TN-30 MN964069 sputum 01/03/2017 0.5 S 0.094 S 2.37 ±0,02 2.27 ±0,08 0.68 ±0,07 1.44 ±0,03 ND ND ND ND
TN-31 MN964070 lung biopsy 17/04/2017 1.5 R 0.75 R 12.99 ±0,02 11.32 ±0,02 3.07 ±0,04 2.71 ±0,02 MN964026 C-T*(183) MN964037 A-G*(529)
TN-32 MN964071 nasal 01/03/2017 0.5 S 1 R 7.49 ±0,02 3.51 ±0,08 1.97 ±0,02 0.77 ±0,05 MN964027 C-T*(183) MN964038 No mutation
Patient 14 Invasive aspergillosis TN-33 MN964072 nasal 21/03/2018 1 R 0.75 R 16.89 ±0,02 2.48 ±0,02 2.57 ±0,06 1.71 ±0,02 MN964028 G-A*(6I6) MN964039 No mutation
TN-34 MN964073 nasal 28/03/2018 0.38 S 0.064 S 6.49 ±0,02 1.52 ±0,08 0.97 ±0,02 0.77 ±0,06 ND ND ND ND

* – Punctual mutation, S – susceptible, R – resistant, VOR – voriconazole, AMB – amphotericin B, ND – not done

The crystal structures of sterol 14-alpha demethylase (CYP51B) from a pathogenic filamentous fungus A. fumigatus in complex with different triazole deposited in the PDB under accession number 4UYL, 5FRB, and 6CR2 were used as the template.

The molecular docking studies were performed using open-source software (PatchDock) (http://bioinfo3d.cs.tau.ac.il/PatchDock) (Thenmozhi and Kannabiran 2013). For this, the input is two molecules of any type: proteins, DNA, peptides, and drugs. The output is a list of potential complexes sorted by shape complementarity criteria and a redirection of the results for refinement and scoring by the FireDock web server (http://bioinfo3d.cs.tau.ac.il/FireDock). A set of candidate complexes (consisting of receptors and ligands) is the input to the FireDock algorithm. All candidates are refined and ranked based on binding energies. The 3D visualization for observation and comparison of refined complexes is provided in the output (Mashiach et al. 2008).

Results
Antifungal susceptibility of A. flavus

The antifungal drug sensitivity profiles of the A. flavus strains were established by ETEST® strips. In all isolates, the MIC levels varied from 0.032 μg/ml to 1.5 μg/ml for IT and from 0.064 to 1 μg/ml for POS. According to the triazole ECVs established for A. flavus by the CLSI M38-A2 microdilution method, we noted the presence of 5 out of 34 A. flavus strains (14.71%) and 3 (8.82%) strains having IT and POS-resistant profiles, respectively. Only two strains (5.88%) could be considered cross-resistant to IT and POS (Table III). All isolates were sensitive to VOR.

Mechanisms of azole resistance. Levels of cyp5lA and cyp5lB expression by A. flavus isolates

Thirty-four clinical A. flavus isolates were analyzed by RT-qPCR to quantify expression and number of copies of the two genes (cyp51A and cyp51B) implicated in the molecular mechanisms associated with azole resistance.

So, the expression of cyp51A and B genes were tested on three IT resistant strains, two IT and POS resistant strains, one POS resistant strains, 28 susceptible strains compared to the ANXC4 gene, and the susceptible strain TN-855 (MIC IT: 0,032 μg/ml; MIC VOR: 0,025 μg/ml; MIC POS: 0,012 μg/ml).

The RT-qPCR results are presented in Fig. 1 and 2 as well as in Table III. In the 28 IT/POS sensitive isolates, cyp51A was overexpressed in four isolates and two isolates concerning cyp51B. When examining the 2 IT/POS resistant isolates, overexpression was observed in these two isolates for cyp51 A. The strain (TN-33) showed the highest level (16.89-fold), followed by the strain (TN-31), which had an overproduction of 12.99-fold (Table III). Furthermore, the overexpression was observed only for strain (TN-31) in the cyp51B gene (11.32 fold). We noted that three strains of resistant IT phenotype had no overexpression for the two genes cyp51A and cyp51B. However, the TN-32 strain of the resistant phenotype showed an overexpression of 7.49 and 3.51-fold for these two genes, respectively.

Fig. 1.

Study of the level of expression of cyp51A genes by relative quantification with RT-qPCR in Aspergillus flavus: 28 susceptible strains (striped bars) and 6 IT-POS resistant strains (white bars). Gene expression values are represented as bar plots with mean + SD. p-Values were calculated using the Mann-Whitney U test, p-values of statistical tests are shown within the graphs.

*p < 0.005

Fig. 2.

Study of the level of expression of cyp51B genes by relative quantification with RT-qPCR in Aspergillus flavus: 28 susceptible strains (striped bars) and 6 IT-POS resistant strains (white bars). Gene expression values are represented as bar plots with mean + SD. p-Values were calculated using the Mann-Whitney U test, p-values of statistical tests are shown within the graphs.

*p < 0.005

The overexpression of the cyp51A genes was associated with an increased copy number of genes in A. flavus strains TN-31 and TN-33. Additionally, the presence of cyp51B in multiple copies in the genome has been associated with overexpression in the TN-31 strain.

Sequencing of PCR products

The PCR sequencing showed the presence of three mutations for cyp51A and cyp51B.

The protein-coding regions of the cyp51A and cyp51B genes consist of 1,606 bp and 1,533 bp, respectively. A single short 67 bp intron was identified in the cyp51A gene, while three short 54, 53, and 160 bp introns were identified in the cyp51B gene. The protein coding sequence (1,539 bp) of cyp51A encodes a polypeptide of 513 amino acids, while the coding sequence of the cyp51B gene (1,266 bp) encodes a polypeptide of 422 amino acids.

The sequence of the cyp51A gene in 11 A. flavus isolates was compared with the sequence of wild-type A. flavus strain (NCBI access number XM_002375082.1) (Fig. 3). A nucleotide transition from “C” to “T” at codon 183 in 8 isolates (23.52%) was shown. However, this mutation was synonymous (P61P) and does not affect the structure of the CYP51A protein. This synonymous punctual mutation was also detected in 2 sensitive strains (TN-21, TN-23). We also showed another nonsynonymous mutation caused by the nucleotide transition from “G” to “A” at codon 616 (G206L) for the TN-33 strain with cross IT/POS resistance, causing an amino acid change in the protein sequence (Table III and Fig. 3).

Fig. 3.

Alignment of the DNA sequences of the cyp51A gene fragments from Aspergillus flavus, was compared with wild type A. flavus strain (Genbank ID: XM_002375082.1).

For cyp51B, Fig. 4 illustrates the alignment of our sequences A.flavus isolates of the cyp51B gene with wild type A. flavus strain (NCBI access number XM_002379089.1). Our results showed the presence of a nucleotide transition from “A” to “G” at codon 529 (L177G) for only the TN-31 strain, which in fact exhibits IT-POS cross resistance causing a change in amino acids in the protein sequence.

Fig. 4.

DNA sequences of the cyp51B gene fragments from Aspergillus flavus were compared with wild type A. flavus strain (Genbank ID: XM_002379089.1).

Homology modeling

Sequence alignment was performed using the Clustal Omega program for homology modeling, and open-source software (PatchDock) was used to generate the interaction between the structure of the protein and the drug automatically.

In addition, for all A. flavus isolates, the wild or modified structure of the protein CYP51A and CYP51B showed the interaction with VOR, IT and POS (Fig. 5 and 6), and these could be explained by the fact that these modifications did not affect the protein and antifungal interaction site.

Fig. 5.

Overall view of the 3D model of the Aspergillus flavus CYP51A protein in complex with three antifungal (IT, POS and VOR). Only TN33 strain presented the structure of CYP51A with the substitution of lysine for a glycine in the position 206 of the protein.

Fig. 6.

Overall view of the 3D model of the Aspergillus flavus CYP51B protein in complex with three antifungal (IT, POS and VOR). Only TN31 strain presented the structure of CYP51B with the substitution of glycine for a lysine in the position 177 of the protein.

Discussion

Several antifungal drugs can be used against Aspergillus spp. such as amphotericin B, triazoles and echinocandins. In patients with aspergillosis who require long-term treatment, triazoles become the most effective because they are essential therapeutic options and are of particular significance due to the availability of oral formulations (Denning and Perlin 2011; Arikan-Akdagli et al. 2018).

The triazole antifungals present a broad spectrum for the management of invasive aspergillosis. The treatment of A.flavus infection is similar to that caused by other species of Aspergillus (Meena et al. 2021).

Over the past two decades, many researchers have studied and identified that resistance mechanisms to azoles vary depending on the country and the studies of clinical and environmental strains of A.flavus (Sharma et al. 2018). However, most recently published studies refer to alterations in the structure of cyp51A, cyp51B, or cyp51C, or of these expression levels, as the primary resistance mechanisms of this fungus, which leads to the failure of the treatment in patients with aspergillosis infected with azole-resistant strains. It should be noted that there are two reported pathways for developing azole resistance: the first environmental, corresponding to the use of fungicides to protect crops, and a second medicine due to long-term azole therapies (Gonzalez-Jimenez et al. 2020).

A. flavus has three different enzymes, which are successfully inhibited by azole drugs, suggesting that these enzymes can play a role in azole resistance (Mellado et al. 2001; Howard et al. 2011). However, other Aspergillus spp. and most filamentous fungi possess two CYP51 proteins, CYP51A and CYP51B, although some species have only one (Hawkins et al. 2014). A CYP51A, A2, and B paralog pattern was observed in the azole-resistant species Aspergillus fumigatiaffinis. A new CYP51 paralog, CYP51D, was recently described in fungi, which predominantly occurs in genomes of intrinsically azole-resistant Eurotiomycetes (Van Rhijn et al. 2021). Celia-Sanchez et al. (2022) found four major CYP51 groups: CYP51, CYP51A, CYP51B, and CYP51C. They confirm that all filamentous Ascomycota had a CYP51B paralog, while only 50% had a CYP51A paralog. They also suggest that CYP51 is present in all fungi, including CYP51C, which appears to have diverged from the progenitor of the CYP51A and CYP51B groups. Several synonymous and non-synonymous point mutations were found in cyp51C gene of A.flavus among both susceptible and non-susceptible strains such as M54T, S240A, D254N, I285V, and N423D (Liu et al. 2012; Paul et al. 2015; Sharma et al. 2018), D254G, P276T, S399I, and a mutation responsible for a stop codon (R250ST) (Lucio et al. 2020). The point mutation (Y319H) in A. flavus cyp51C suggests a relation to azole resistance although its role still needs further confirmation (Paul et al. 2015; Sharma et al. 2018).

However, there is minimal data on A. flavus isolates in Tunisia, and less is known of their resistance mechanisms, despite being the most commonly isolated mold species from patients and their environment (Hadrich et al. 2010). Therefore, our study evaluated the quantification of expression and number of copies for the two genes (cyp51A and cyp51B) in clinical A. flavus isolates IT-resistant and POS-resistant.

Our findings showed overexpression in two IT/POS resistant A. flavus isolates for cyp51A with the highest level of 16.8-fold and in one IT/POS resistant isolate (TN-31) for cyp51B, suggesting that these two genes contribute to resistance to azoles. According to Fattahi et al. (2015) a significant up-regulation of cyp51A (6.89 to 7.93-folds) (p < 0.001) in four A. flavus VOR-resistant isolates was found. Several authors have agreed that the up-regulation of the cyp51A gene is most commonly associated with azole resistance in A. fumigatus and Aspergillus lentulus (Howard et al. 2010; Mellado et al. 2011). In addition, CYP51B has been shown to be a functional biosynthetic enzyme that is inhibited by azoles. However, Liu et al. indicated that the expression levels of the three genes cyp51A, cyp51B, and cyp51C were not associated with azole resistance in A. flavus (Liu et al. 2012).

However, Sharma et al. (2018), found that cyp51A and cyp51B were upregulated (> 3-fold) for a single VOR-resistant strain (VPCI195/P/10). Contrary to what was expected, the only strain with cross-resistance to antifungals did not show up-regulation of any target genes analyzed.

In this study, the option of overexpression in sensitive profile strains may be explained by the vast clinical use of this antifungal in our country, particularly in patients with chronic diseases. This suggests that the cyp51A and B gene overexpression does not necessarily result in azole resistance in this fungus.

Furthermore, the mechanism of azole resistance in most of the previously published studies is associated with changes that are often corresponding to mutations in the cyp51A gene (Nascimento et al. 2003; Chen et al. 2005; Garcia-Effron et al. 2005; Buil et al. 2018). Several publications have clearly described single amino acid substitutions at G54, P216, F219, M220, and G448 in the CYP51A protein to confer azole resistance in A. fumigatus (Mann et al. 2003; Mellado et al. 2004; Camps et al. 2012; Krishnan Natesan et al. 2012). Over time, besides these ‘hot spots’, amino acid changes at other sites (Y121, G138, and Y431) have also been found in the CYP51A protein of resistant strains (Albarrag et al. 2001; Lescar et al. 2014). Prolonged azole treatment in chronic aspergillosis is a significant cause of these resistance mutations (Howard et al. 2009), as is environmental exposure to azole fungicides (Snelders et al. 2008). Typically, combinations of tandem repeats and point mutations of the cyp51A gene (TR34/L98H and TR46/Y121F/T289A) have been observed in azoleresistant A. fumigatus isolated from patients regardless of their azole treatment history (Snelders et al. 2008).

Here, we described for A. flavus, three new point mutations observed in Cyp51A (a synonymous substitution (P61P) and a non-synonymous (G206L)) and Cyp51B (a non-synonymous mutation in cyp51B (L177G). To our knowledge, these point mutations have not been described before. In a recent report, susceptible and non-susceptible A. flavus strains presented several synonymous and nonsynonymous point mutations in cyp51A and cyp51C genes. However, two amino acid mutations were noted in only two resistant strains: one strain contained a P214L substitution in cyp51A, and another was H349R in cyp51C (Gonzalez-Jimenez et al. 2020).

As stated previously, Krishnan-Natesan et al. (2008) were among the first to identify different mutations of cyp51A and cyp51B in spontaneous mutant strains obtained after exposure to voriconazole. These mutations were found in both cyp51A (K197N, D282E, M288L, Y132N, and T469S) and cyp51B (H399P, D411N, T454P, and T486P). However, these substitutions were recently revealed in azoles-sensitive A. flavus strains, making their implication in drug resistance highly unlikely (Hagiwara et al. 2016; Lucio et al. 2020). Similarly, Lucio et al. (2020) found that cyp51B showed no amino acid changes in A. flavus strains, except for some polymorphisms responsible for synonymous mutations. In our study, synonym and nonsynonymous mutations of cyp51A (P61P and G206L) were detected in azole-sensitive and non-sensitive A. flavus strains, excluding the possibility that they are associated with azole resistance. Similarly, Lucio et al. (2020) reported in susceptible and resistant strains the presence of three synonymous and nonsynonymous point mutations.

In previous studies, the crystallographic structures of the CYP51 proteins of Saccharomyces cerevisiae complexed with the substrates lanosterol, itraconazole, voriconazole, and fluconazole have been described (Monk et al. 2014; Sagatova et al. 2015). So also, the structure of Candida albicans CYP51 complexed with posaconazole (5FSA) and the tetrazole-based antifungal drug candidate VT1161 was analyzed (Hargrove et al. 2017a). Hargrove et al. (2017b) have described the structure of CYP51B in complex with the NIV derivative and the tetrazole-based VT-1598 inhibitor for A. fumigatus. Upon the recently resolved crystal structure of the CYP51B protein of A. fumigatus, a new 3D structural model of A. flavus CYP51A and CYP51B was built. Docking analysis of three azole molecules to their target proteins, CYP51A and CYP51B was performed based on this structural model.

According to the models provided by the PatchDock software, three new mutations, a synonym and a nonsynonym in cyp51A (P61P and G206L) and a nonsynonymous mutation in cyp51B (L177G) do not affect the interaction between the drug and protein CYP51A or CYP51B. The resistance of our strains could be mainly due to the overexpression of the cyp51A and cyp51B genes. However, an in-depth study of azole resistance-related genes in a large population of clinical isolates will expand our knowledge.

Therefore, further examination of other resistance mechanisms is necessary, like the CCAAT-binding domain complex CBC (a heterotrimer comprising HapB, HapC, and HapE, which is a negative regulator of sterol biosynthesis directly binding the promoters of 14 ergosterol biosynthetic genes), overexpression of MFS (major facilitator superfamily) transporters or overexpression of ABC transporters (ATP binding cassette superfamily). A deeper understanding of azole resistance mechanisms will facilitate the development of new therapeutic drugs against azole-resistant strains of Aspergillus.

eISSN:
2544-4646
Język:
Angielski
Częstotliwość wydawania:
4 razy w roku
Dziedziny czasopisma:
Life Sciences, Microbiology and Virology