Deciphering the Role of PIG1 and DHN-Melanin in Scedosporium apiospermum Conidia

Hélène Guegan, Wilfried Poirier, Kevin Ravenel, Sarah Dion, Aymeric Delabarre, Dimitri Desvillechabrol, Xavier Pinson, Odile Sergent, Isabelle Gallais, Jean-Pierre Gangneux, Sandrine Giraud, Amandine Gastebois, Hélène Guegan, Wilfried Poirier, Kevin Ravenel, Sarah Dion, Aymeric Delabarre, Dimitri Desvillechabrol, Xavier Pinson, Odile Sergent, Isabelle Gallais, Jean-Pierre Gangneux, Sandrine Giraud, Amandine Gastebois

Abstract

Scedosporium apiospermum is a saprophytic filamentous fungus involved in human infections, of which the virulence factors that contribute to pathogenesis are still poorly characterized. In particular, little is known about the specific role of dihydroxynaphtalene (DHN)-melanin, located on the external layer of the conidia cell wall. We previously identified a transcription factor, PIG1, which may be involved in DHN-melanin biosynthesis. To elucidate the role of PIG1 and DHN-melanin in S. apiospermum, a CRISPR-Cas9-mediated PIG1 deletion was carried out from two parental strains to evaluate its impact on melanin biosynthesis, conidia cell-wall assembly, and resistance to stress, including the ability to survive macrophage engulfment. ΔPIG1 mutants did not produce melanin and showed a disorganized and thinner cell wall, resulting in a lower survival rate when exposed to oxidizing conditions, or high temperature. The absence of melanin increased the exposure of antigenic patterns on the conidia surface. PIG1 regulates the melanization of S. apiospermum conidia, and is involved in the survival to environmental injuries and to the host immune response, that might participate in virulence. Moreover, a transcriptomic analysis was performed to explain the observed aberrant septate conidia morphology and found differentially expressed genes, underlining the pleiotropic function of PIG1.

Keywords: CRISPR-Cas9; DHN-melanin; PIG1; RNA-seq; Scedosporium apiospermum; cell wall.

Conflict of interest statement

The authors declare no conflict of interest.

Figures

Figure 1
Figure 1
Gene cluster of melanin biosynthesis in S. apiospermum (36.1 kb), from the published S. apiospermum genome [30]. The scytalone deshydratase encoding gene is located out of the cluster.
Figure 2
Figure 2
Generation of S. apiospermum ΔPIG1 mutants. (a) Restriction map of the S. apiospermum PIG1 locus and strategy for the construction of the disruption cassette. NcoI restriction sites are identified and the purple line indicates the hybridization site of the probe. The size of the expected fragments is indicated by the arrows. (b) Southern-blot analysis of PIG1 transformant strains. Genomic DNA of all strains (parental and mutant derivatives) were digested by NcoI and probed with a hybridization probe corresponding to a 600-bp fragment including the PIG1 3′ flanking region and 3′ end of the HPH gene. The expected signals were 1700 bp for the parental strains and 1123 bp for the M1 and M2 mutants. Two unexpected bands were observed for the T strain (2411 bp and 1700 bp), suggesting integration of the repair cassette upstream of the PIG1 gene. (c) Expression of PIG1 mRNA by the three HPH-resistant transformants relative to their respective parental strain (i.e., T vs. WT, M1 vs. ΔKU70, and M2 vs. ΔKU70). After reverse transcription, mRNA expression was quantified by quantitative PCR. Data were normalized to the expression of the UBC6 and FIS1 housekeeping genes and are representative of four independent experiments. Significant differences are indicated by asterisks: * p ≤ 0.05 (Mann-Whitney test). T: transformant derived from the WT strain; M1 and M2: transformants derived from the ΔKU70 strain.
Figure 3
Figure 3
Detection of melanin in parental and mutant S. apiospermum strains. (a) Macroscopic aspect of parental and mutant strains after nine days at 37 °C on PDA medium. WT-TRC indicates the WT strain grown on a PDA plate supplemented with 50 µg/mL tricyclazole as a specific inhibitor of melanin biosynthesis. (b) Melanin quantity per million conidia, extrapolated from a standard curve of synthetic melanin, determined by EPR (Graph representative of at least three independent measurements). T: transformant derived from the WT strain; M1 and M2: transformants derived from the ΔKU70 strain.
Figure 4
Figure 4
Analysis of the fluorescence labelling of conidia surface carbohydrates with FITC-conjugated lectins. (a,b) Fluorescence intensity of (a) ConA binding to mannosyl and glucosyl residues and (b) WGA binding to N-acetylglucosamine (NAG) residues on the surface of the conidia. Fluorescence analysis was performed using ImageJ (Fiji) software from at least 200 conidia. The graph is representative of three independent experiments. Significant differences are indicated: *** p < 0.001 (Kruskall-Wallis test). T: transformant derived from the WT strain; M1 and M2: transformants derived from the ΔKU70 strain.
Figure 5
Figure 5
Evaluation of cell-wall thickness and integrity. (a) TEM of ΔKU70 and M1 conidia cell walls. The image is representative of the three transformants. (b) Thickness determined from transmission electron microscopy (TEM) acquisition for 20 acquisitions. (c) Growth in contact with the cell-wall chemical stressors Congo Red and SDS. After an incubation of seven days at 30 °C, the colony diameter was measured and expressed according to the control growth on stress-free PDA. All graphs include results obtained from three independent experiments. Significant differences are indicated: * p < 0.05, ** p < 0.01, *** p < 0.001 (Kruskall-Wallis test). T: transformant derived from the WT strain; M1 and M2: transformants derived from the ΔKU70 strain.
Figure 6
Figure 6
Morphology of conidia. (a) Distribution of conidia morphotypes. At least 200 conidia were measured. (b) Distribution of conidia sizes according to cross-sectional area. At least 200 conidia were measured from images of conidia suspensions acquired with a widefield inverted fluorescence microscope (Olympus IX71) at 20× magnification and analyzed using ImageJ (Fiji) software. (c) Transmission electron microscopy (TEM) of ΔKU70 and M1 conidia. Images are representative of the three transformants. Scale bar: 2 µm. T: transformant derived from the WT strain; M1 and M2: transformants derived from the ΔKU70 strain.
Figure 7
Figure 7
Biofilm formation. The total biomass was monitored using a crystal violet assay. Results from three independent experiments are shown. Significant differences are indicated as follows: * p < 0.05, ** p < 0.01 (Kruskall-Wallis test). T: transformant derived from the WT strain; M1 and M2: transformants derived from the ΔKU70 strain.
Figure 8
Figure 8
Susceptibility of the T and the M1 and M2 deleted strains exposed to exogenous stresses. (a) Conidia survival after exposure to UV-C light. After exposure to a dose of 30 mJ/cm2 at 254 nm, conidia were stained with 25 µM propidium iodide (PI) before being analyzed on a LSRFortessaTM cytometer. The graph presents the percentage survival among exposed FITC+ conidia expressed relative to that of unexposed FITC+ conidia (live control), as follows: 100× ((% PI− FITC+ conidia among unexposed control conidia—% PI−FITC+ conidia among exposed conidia)/(% PI−FITC+ conidia among unexposed control conidia)). (b) Conidia survival after exposure to high temperature. After incubation at 40 °C or 50 °C for 15 min, 200 conidia were incubated on PDA plates at 37 °C for 72 h. The number of colonies reflect the number of live conidia. The graph presents the ratio between the number of colonies recovered after exposure of the conidia to heat and the number of colonies recovered after no heat exposure on PDA plates. (c,d) Growth under oxidative conditions. Mycelial growth in PDB medium supplemented with 200 µM cumene hydroperoxide (c) or 40 µM menadione (d) was monitored when incubated for 40 h at 37 °C by nephelometry. Results are expressed in relative nephelometric units resulting from the measurement of the intensity of scattered light through the wall as an indication of the increasing turbidity due to the mycelia. DMSO was used to dissolve the menadione powder. Graphs present the relative growth calculated as the ratio between the nephelometric signal obtained in the presence of an oxidizing compound and in oxidant-free PDB (growth control). All graphs include results obtained from three independent experiments: * p < 0.05, ** p < 0.01, *** p < 0.001 (Kruskall-Wallis test). T: transformant derived from the WT strain; M1 and M2: transformants derived from the ΔKU70 strain.
Figure 9
Figure 9
Functions of differentially expressed genes (DEGs) in M1 relative to ΔKU70. (a) Distribution of functions or metabolic pathways in DEGs with a known function (N = 198). The 80 DEGs with unknown function are not included. (b) Detailed functions of DEGs involved in an oxidoreduction process.
Figure 10
Figure 10
Fold-induction mRNA values for a selected panel of genes in the T, M1, and M2 strains relative to their parental ΔKU70 strain. Expression levels of target genes were normalized against the expression of the S. apiospermum UBC6 and FIS1 housekeeping genes. Results are expressed as 2−ΔΔCq, referring to the fold induction in the T, M1, and M2 strains relative to the mean quantification cycle obtained with the parental strains. Data are from four biological replicates: * p ≤ 0.05 (Mann-Whitney test). T: transformant derived from the WT strain; M1 and M2: transformants derived from the ΔKU70 strain.
Figure 11
Figure 11
Phagocytosis, killing, and cytokine release from human macrophages infected with T, M1, and M2 conidia. (a) Percentage of macrophages with adherent or ingested conidia after 2 or 6 h of incubation with FITC-stained conidia (MOI 5:1) at 37 °C in 5% CO2. Graphs present the percentage of FITC+ CD11b+ events among CD11b+ cells (macrophages). (b) Conidia killing after ingestion by macrophages. After 6 h of incubation (MOI 1:1), extracellular conidia were removed by three PBS washes and internalized conidia were released by macrophage lysis (cold water). The graph presents the decrease in conidia viability among ingested conidia relative to the viability of conidia not exposed to macrophages (live control). Conidia viability was determined as the percentage of FITC+ PI– events among FITC+ events (total conidia). The reduction in conidia survival was calculated relative to control conidia as: 100 × ((% of live conidia among control conidia − % of live conidia among cell-exposed conidia)/% of live conidia among control conidia). (c) Cytokine release in culture supernatants from macrophages incubated with conidia for 12 h (MOI 10:1). Concentrations were determined using a 13 plex-bead-based LEGENDplex® multianalyte flow assay. A non-infected (NI) condition was included in each experiment. Data are from three independent experiments. * p < 0.05, ** p < 0.01, *** p < 0.001 (Kruskall-Wallis test) T: transformant derived from the WT strain; M1 and M2: transformants derived from the ΔKU70 strain.

References

    1. Cortez K.J., Roilides E., Quiroz-Telles F., Meletiadis J., Antachopoulos C., Knudsen T., Buchanan W., Milanovich J., Sutton D.A., Fothergill A., et al. Infections caused by Scedosporium spp. Clin. Microbiol. Rev. 2008;21:157–197. doi: 10.1128/CMR.00039-07.
    1. Cimon B., Carrère J.F., Vinatier J.P., Chazalette J., Chabasse D., Bouchara J.P. Clinical Significance of Scedosporium apiospermum in Patients with Cystic Fibrosis. Eur. J. Clin. Microbiol. Infect. Dis. 2000;19:53–56. doi: 10.1007/s100960050011.
    1. Sedlacek L., Graf B., Schwarz C., Albert F., Peter S., Würstl B., Wagner S., Klotz M., Becker A., Haase G., et al. Prevalence of Scedosporium species and Lomentospora prolificans in patients with cystic fibrosis in a multicenter trial by use of a selective medium. J. Cyst. Fibros. 2015;14:237–241. doi: 10.1016/j.jcf.2014.12.014.
    1. Blyth C.C., Middleton P.G., Harun A., Sorrell T.C., Meyer W., Chen S.-A. Clinical associations and prevalence of Scedosporium spp. in Australian cystic fibrosis patients: Identification of novel risk factors? Med. Mycol. 2010;48:S37–S44. doi: 10.3109/13693786.2010.500627.
    1. Seidel D., Hassler A., Salmanton-García J., Koehler P., Mellinghoff S.C., Carlesse F., Cheng M.P., Falces-Romero I., Herbrecht R., Jover Sáenz A., et al. Invasive Scedosporium spp. and Lomentospora prolificans infections in pediatric patients: Analysis of 55 cases from FungiScope® and the literature. Int. J. Infect. Dis. 2020;92:114–122. doi: 10.1016/j.ijid.2019.12.017.
    1. Bronnimann D., Garcia-Hermoso D., Dromer F., Lanternier F., French Mycoses Study Group Characterization of the isolates at the NRCMA. Scedosporiosis/lomentosporiosis observational study (SOS): Clinical significance of Scedosporium species identification. Med. Mycol. 2021;59:486–497. doi: 10.1093/mmy/myaa086.
    1. Cornely O.A., Alastruey-Izquierdo A., Arenz D., Chen S.C.A., Dannaoui E., Hochhegger B., Hoenigl M., Jensen H.E., Lagrou K., Lewis R.E., et al. Mucormycosis ECMM MSG Global Guideline Writing Group, Global guideline for the diagnosis and management of mucormycosis: An initiative of the European Confederation of Medical Mycology in cooperation with the Mycoses Study Group Education and Research Consortium. Lancet Infect. Dis. 2019;19:e405–e421. doi: 10.1016/S1473-3099(19)30312-3.
    1. Latgé J.P. The pathobiology of Aspergillus fumigatus. Trends Microbiol. 2001;9:382–389. doi: 10.1016/S0966-842X(01)02104-7.
    1. Cunha C., Kurzai O., Löffler J., Aversa F., Romani L., Carvalho A. Neutrophil responses to aspergillosis: New roles for old players. Mycopathologia. 2014;178:387–393. doi: 10.1007/s11046-014-9796-7.
    1. Chamilos G., Carvalho A. Aspergillus fumigatus DHN-Melanin. Curr. Top. Microbiol. Immunol. 2020;425:17–28. doi: 10.1007/82_2020_205.
    1. Camacho E., Vij R., Chrissian C., Prados-Rosales R., Gil D., O’Meally R.N., Cordero R.J.B., Cole R.N., McCaffery J.M., Stark R.E., et al. The structural unit of melanin in the cell wall of the fungal pathogen Cryptococcus neoformans. J. Biol. Chem. 2019;294:10471–10489. doi: 10.1074/jbc.RA119.008684.
    1. Langfelder K., Streibel M., Jahn B., Haase G., Brakhage A.A. Biosynthesis of fungal melanins and their importance for human pathogenic fungi. Fungal Genet. Biol. 2003;38:143–158. doi: 10.1016/S1087-1845(02)00526-1.
    1. Wang T., Ren D., Guo H., Chen X., Zhu P., Nie H., Xu L. CgSCD1 Is Essential for Melanin Biosynthesis and Pathogenicity of Colletotrichum gloeosporioides. Pathogens. 2020;9:141. doi: 10.3390/pathogens9020141.
    1. Cho Y., Srivastava A., Ohm R.A., Lawrence C.B., Wang K.-H., Grigoriev I.V., Marahatta S.P. Transcription factor Amr1 induces melanin biosynthesis and suppresses virulence in Alternaria brassicicola. PLoS Pathog. 2012;8:e1002974. doi: 10.1371/journal.ppat.1002974.
    1. van de Veerdonk F.L., Gresnigt M.S., Romani L., Netea M.G., Latgé J.-P. Aspergillus fumigatus morphology and dynamic host interactions. Nat. Rev. Microbiol. 2017;15:661–674. doi: 10.1038/nrmicro.2017.90.
    1. Feng B., Wang X., Hauser M., Kaufmann S., Jentsch S., Haase G., Becker J.M., Szaniszlo P.J. Molecular cloning and characterization of WdPKS1, a gene involved in dihydroxynaphthalene melanin biosynthesis and virulence in Wangiella (Exophiala) dermatitidis. Infect. Immun. 2001;69:1781–1794. doi: 10.1128/IAI.69.3.1781-1794.2001.
    1. Xiao X., Li Y., Lan Y., Zhang J., He Y., Cai W., Chen Z., Xi L., Zhang J. Deletion of pksA attenuates the melanogenesis, growth and sporulation ability and causes increased sensitivity to stress response and antifungal drugs in the human pathogenic fungus Fonsecaea monophora. Microbiol. Res. 2021;244:126668. doi: 10.1016/j.micres.2020.126668.
    1. Chai L.Y.A., Netea M.G., Sugui J., Vonk A.G., van de Sande W.W.J., Warris A., Kwon-Chung K.J., Jan Kullberg B. Aspergillus fumigatus Conidial Melanin Modulates Host Cytokine Response. Immunobiology. 2010;215:915–920. doi: 10.1016/j.imbio.2009.10.002.
    1. Heinekamp T., Thywißen A., Macheleidt J., Keller S., Valiante V., Brakhage A.A. Aspergillus fumigatus melanins: Interference with the host endocytosis pathway and impact on virulence. Front. Microbiol. 2012;3:440. doi: 10.3389/fmicb.2012.00440.
    1. Jahn B., Koch A., Schmidt A., Wanner G., Gehringer H., Bhakdi S., Brakhage A.A. Isolation and characterization of a pigmentless-conidium mutant of Aspergillus fumigatus with altered conidial surface and reduced virulence. Infect. Immun. 1997;65:5110–5117. doi: 10.1128/iai.65.12.5110-5117.1997.
    1. van Duin D., Casadevall A., Nosanchuk J.D. Melanization of Cryptococcus neoformans and Histoplasma capsulatum Reduces Their Susceptibilities to Amphotericin B and Caspofungin. Antimicrob. Agents Chemother. 2002;46:3394–3400. doi: 10.1128/AAC.46.11.3394-3400.2002.
    1. Wang Y., Casadevall A. Growth of Cryptococcus neoformans in presence of L-dopa decreases its susceptibility to amphotericin B. Antimicrob. Agents Chemother. 1994;38:2648–2650. doi: 10.1128/AAC.38.11.2648.
    1. Paolo W.F., Dadachova E., Mandal P., Casadevall A., Szaniszlo P.J., Nosanchuk J.D. Effects of disrupting the polyketide synthase gene WdPKS1 in Wangiella [Exophiala] dermatitidis on melanin production and resistance to killing by antifungal compounds, enzymatic degradation, and extremes in temperature. BMC Microbiol. 2006;6:55. doi: 10.1186/1471-2180-6-55.
    1. Rollin-Pinheiro R., da Silva Xisto M.I.D., Rochetti V.P., Barreto-Bergter E. Scedosporium Cell Wall: From Carbohydrate-Containing Structures to Host-Pathogen Interactions. Mycopathologia. 2020;185:931–946. doi: 10.1007/s11046-020-00480-7.
    1. Rollin-Pinheiro R., Liporagi-Lopes L.C., de Meirelles J.V., de Souza L.M., Barreto-Bergter E. Characterization of Scedosporium apiospermum glucosylceramides and their involvement in fungal development and macrophage functions. PLoS ONE. 2014;9:e98149. doi: 10.1371/journal.pone.0098149.
    1. Rochetti V.P., Rollin-Pinheiro R., de Oliveira E.B., da Silva Xisto M.I.D., Barreto-Bergter E. Glucosylceramide Plays a Role in Fungal Germination, Lipid Raft Organization and Biofilm Adhesion of the Pathogenic Fungus Scedosporium aurantiacum. J. Fungi. 2020;6:345. doi: 10.3390/jof6040345.
    1. de Meirelles J.V., da Silva Xisto M.I.D., Rollin-Pinheiro R., Serrato R.V., Haido R.M.T., Barreto-Bergter E. Peptidorhamanomannan: A surface fungal glycoconjugate from Scedosporium aurantiacum and Scedosporium minutisporum and its recognition by macrophages. Med. Mycol. 2021;59:441–452. doi: 10.1093/mmy/myaa065.
    1. Ruiz-Díez B., Martínez-Suárez J.V. Isolation, characterization, and antifungal susceptibility of melanin-deficient mutants of Scedosporium prolificans. Curr. Microbiol. 2003;46:228–232. doi: 10.1007/s00284-002-3858-7.
    1. Al-Laaeiby A., Kershaw M.J., Penn T.J., Thornton C.R. Targeted Disruption of Melanin Biosynthesis Genes in the Human Pathogenic Fungus Lomentospora prolificans and Its Consequences for Pathogen Survival. Int. J. Mol. Sci. 2016;17:444. doi: 10.3390/ijms17040444.
    1. Vandeputte P., Ghamrawi S., Rechenmann M., Iltis A., Giraud S., Fleury M., Thornton C., Delhaès L., Meyer W., Papon N., et al. Draft Genome Sequence of the Pathogenic Fungus Scedosporium apiospermum. Genome Announc. 2014;2:e00988-14. doi: 10.1128/genomeA.00988-14.
    1. Tsuji G., Kenmochi Y., Takano Y., Sweigard J., Farrall L., Furusawa I., Horino O., Kubo Y. Novel fungal transcriptional activators, Cmr1p of Colletotrichum lagenarium and pig1p of Magnaporthe grisea, contain Cys2His2 zinc finger and Zn(II)2Cys6 binuclear cluster DNA-binding motifs and regulate transcription of melanin biosynthesis genes in a developmentally specific manner. Mol. Microbiol. 2000;38:940–954. doi: 10.1046/j.1365-2958.2000.02181.x.
    1. Eliahu N., Igbaria A., Rose M.S., Horwitz B.A., Lev S. Melanin biosynthesis in the maize pathogen Cochliobolus heterostrophus depends on two mitogen-activated protein kinases, Chk1 and Mps1, and the transcription factor Cmr1. Eukaryot. Cell. 2007;6:421–429. doi: 10.1128/EC.00264-06.
    1. Le Govic Y., Havlíček V., Capilla J., Luptáková D., Dumas D., Papon N., Le Gal S., Bouchara J.-P., Vandeputte P. Synthesis of the Hydroxamate Siderophore N α-Methylcoprogen B in Scedosporium apiospermum Is Mediated by sidD Ortholog and Is Required for Virulence. Front. Cell. Infect. Microbiol. 2020;10:587909. doi: 10.3389/fcimb.2020.587909.
    1. Al Abdallah Q., Ge W., Fortwendel J.R. A Simple and Universal System for Gene Manipulation in Aspergillus fumigatus: In Vitro-Assembled Cas9-Guide RNA Ribonucleoproteins Coupled with Microhomology Repair Templates. mSphere. 2017;2:e00446-17. doi: 10.1128/mSphere.00446-17.
    1. Pateau V., Razafimandimby B., Vandeputte P., Thornton C.R., Guillemette T., Bouchara J.-P., Giraud S. Gene Disruption in Scedosporium aurantiacum: Proof of Concept with the Disruption of SODC Gene Encoding a Cytosolic Cu,Zn-Superoxide Dismutase. Mycopathologia. 2018;183:241–249. doi: 10.1007/s11046-017-0204-y.
    1. Rollin-Pinheiro R., de Meirelles J.V., Vila T.V.M., Fonseca B.B., Alves V., Frases S., Rozental S., Barreto-Bergter E. Biofilm Formation by Pseudallescheria/Scedosporium Species: A Comparative Study. Front. Microbiol. 2017;8:1568. doi: 10.3389/fmicb.2017.01568.
    1. Mello T.P., Aor A.C., Gonçalves D.S., Seabra S.H., Branquinha M.H., Santos A.L.S. Assessment of biofilm formation by Scedosporium apiospermum, S. aurantiacum, S. minutisporum and Lomentospora prolificans. Biofouling. 2016;32:737–749. doi: 10.1080/08927014.2016.1192610.
    1. Joubert A., Calmes B., Berruyer R., Pihet M., Bouchara J.-P., Simoneau P., Guillemette T. Laser nephelometry applied in an automated microplate system to study filamentous fungus growth. Biotechniques. 2010;48:399–404. doi: 10.2144/000113399.
    1. Guinea J., Meletiadis J., Arikan-Akdagli S., Muehlethaler K., Arendrup M.C. the Subcommittee on Antifungal Susceptibility Testing (AFST) of the ESCMID European Committee for Antimicrobial Susceptibility Testing (EUCAST): New Version of Broth Microdilution of Moulds (vs. 9.4) 2022. [(accessed on 1 April 2022)]. Available online:
    1. Cokelaer T., Desvillechabrol D., Legendre R., Cardon M. “Sequana”: A Set of Snakemake NGS pipelines. J. Open Source Softw. 2017;2:352. doi: 10.21105/joss.00352.
    1. Köster J., Rahmann S. Snakemake—A scalable bioinformatics workflow engine. Bioinformatics. 2012;28:2520–2522. doi: 10.1093/bioinformatics/bts480.
    1. Chen S., Zhou Y., Chen Y., Gu J. fastp: An ultra-fast all-in-one FASTQ preprocessor. Bioinformatics. 2018;34:i884–i890. doi: 10.1093/bioinformatics/bty560.
    1. Dobin A., Davis C.A., Schlesinger F., Drenkow J., Zaleski C., Jha S., Batut P., Chaisson M., Gingeras T.R. STAR: Ultrafast universal RNA-seq aligner. Bioinformatics. 2013;29:15–21. doi: 10.1093/bioinformatics/bts635.
    1. Liao Y., Smyth G.K., Shi W. featureCounts: An efficient general purpose program for assigning sequence reads to genomic features. Bioinformatics. 2014;30:923–930. doi: 10.1093/bioinformatics/btt656.
    1. Ewels P., Magnusson M., Lundin S., Käller M. MultiQC: Summarize analysis results for multiple tools and samples in a single report. Bioinformatics. 2016;32:3047–3048. doi: 10.1093/bioinformatics/btw354.
    1. Love M.I., Huber W., Anders S. Moderated estimation of fold change and dispersion for RNA-seq data with DESeq2. Genome Biol. 2014;15:550. doi: 10.1186/s13059-014-0550-8.
    1. Staerck C., Yaakoub H., Vandeputte P., Tabiasco J., Godon C., Gastebois A., Giraud S., Guillemette T., Calenda A., Delneste Y., et al. The Glycosylphosphatidylinositol-Anchored Superoxide Dismutase of Scedosporium apiospermum Protects the Conidia from Oxidative Stress. J. Fungi. 2021;7:575. doi: 10.3390/jof7070575.
    1. Butler M.J., Day A.W. Fungal melanins: A review. Can. J. Microbiol. 1998;44:1115–1136. doi: 10.1139/w98-119.
    1. Cunha M.M.L., Franzen A.J., Alviano D.S., Zanardi E., Alviano C.S., De Souza W., Rozental S. Inhibition of melanin synthesis pathway by tricyclazole increases susceptibility of Fonsecaea pedrosoi against mouse macrophages. Microsc. Res. Tech. 2005;68:377–384. doi: 10.1002/jemt.20260.
    1. Liu X.-H., Ning G.-A., Huang L.-Y., Zhao Y.-H., Dong B., Lu J.-P., Lin F.-C. Calpains are involved in asexual and sexual development, cell wall integrity and pathogenicity of the rice blast fungus. Sci. Rep. 2016;6:31204. doi: 10.1038/srep31204.
    1. Wang X., Lu D., Tian C. Analysis of melanin biosynthesis in the plant pathogenic fungus Colletotrichum gloeosporioides. Fungal Biol. 2021;125:679–692. doi: 10.1016/j.funbio.2021.04.004.
    1. Huang P., Cao H., Li Y., Zhu S., Wang J., Wang Q., Liu X., Lin F.-C., Lu J. Melanin Promotes Spore Production in the Rice Blast Fungus Magnaporthe oryzae. Front. Microbiol. 2022;13:843838. doi: 10.3389/fmicb.2022.843838.
    1. Liu Y. RNA interference Pathways in Filamentous Fungi. Cell Mol Life Sci. 2010;67:3849–3863. doi: 10.1007/s00018-010-0471-y.
    1. Fernandes C., Mota M., Barros L., Dias M.I., Ferreira I.C.F.R., Piedade A.P., Casadevall A., Gonçalves T. Pyomelanin Synthesis in Alternaria alternata Inhibits DHN-Melanin Synthesis and Decreases Cell Wall Chitin Content and Thickness. Front. Microbiol. 2021;12:691433. doi: 10.3389/fmicb.2021.691433.
    1. Caesar-Tonthat T., Van Ommen K.F., Geesey G.G., Henson J.M. Melanin production by a filamentous soil fungus in response to copper and localization of copper sulfide by sulfide-silver staining. Appl. Environ. Microbiol. 1995;61:1968–1975. doi: 10.1128/aem.61.5.1968-1975.1995.
    1. Fetzner R., Seither K., Wenderoth M., Herr A., Correspondence R.F., Fischer R. Alternaria alternata transcription factor CmrA controls melanization and spore development. Microbiology. 2014;160:1845–1854. doi: 10.1099/mic.0.079046-0.
    1. Chattopadhyay A., Kushwaha C., Chand R., Srivastava J.S. Differential mode of action of tricyclazole in vitro and in planta on Bipolaris sorokiniana causing spot blotch in barley. Indian Phytopathol. 2012;66:155–158.
    1. Pinto M., Rodrigues M., Travassos L.R., Haido R., Wait R., Barreto-Bergter E. Characterization of glucosylceramides in Pseudallescheria boydii and their involvement in fungal differentiation. Glycobiology. 2002;12:251–260. doi: 10.1093/glycob/12.4.251.
    1. Fernandes C.M., Goldman G.H., Del Poeta M. Biological Roles Played by Sphingolipids in Dimorphic and Filamentous Fungi. mBio. 2018;9:e00642-18. doi: 10.1128/mBio.00642-18.
    1. Oura T., Kajiwara S. Disruption of the sphingolipid Delta8-desaturase gene causes a delay in morphological changes in Candida albicans. Microbiology. 2008;154:3795–3803. doi: 10.1099/mic.0.2008/018788-0.
    1. Rittershaus P.C., Kechichian T.B., Allegood J.C., Merrill A.H., Hennig M., Luberto C., Del Poeta M. Glucosylceramide synthase is an essential regulator of pathogenicity of Cryptococcus neoformans. J. Clin. Investig. 2006;116:1651–1659. doi: 10.1172/JCI27890.
    1. Fernandes C.M., de Castro P.A., Singh A., Fonseca F.L., Pereira M.D., Vila T.V.M., Atella G.C., Rozental S., Savoldi M., Del Poeta M., et al. Functional characterization of the Aspergillus nidulans glucosylceramide pathway reveals that LCB Δ8-desaturation and C9-methylation are relevant to filamentous growth, lipid raft localization and Psd1 defensin activity. Mol. Microbiol. 2016;102:488–505. doi: 10.1111/mmi.13474.
    1. Del Poeta M., Nimrichter L., Rodrigues M.L., Luberto C. Synthesis and Biological Properties of Fungal Glucosylceramide. PLoS Pathog. 2014;10:e1003832. doi: 10.1371/journal.ppat.1003832.
    1. Mor V., Rella A., Farnoud A.M., Singh A., Munshi M., Bryan A., Naseem S., Konopka J.B., Ojima I., Bullesbach E., et al. Identification of a New Class of Antifungals Targeting the Synthesis of Fungal Sphingolipids. mBio. 2015;6:e00647. doi: 10.1128/mBio.00647-15.
    1. McEvoy K., Normile T.G., Poeta M.D. Antifungal Drug Development: Targeting the Fungal Sphingolipid Pathway. J. Fungi. 2020;6:142. doi: 10.3390/jof6030142.
    1. Ruger-Herreros C., Corrochano L.M. Conidiation in Neurospora crassa: Vegetative reproduction by a model fungus. Int. Microbiol. 2020;23:97–105. doi: 10.1007/s10123-019-00085-1.
    1. Olmedo M., Ruger-Herreros C., Luque E.M., Corrochano L.M. A complex photoreceptor system mediates the regulation by light of the conidiation genes con-10 and con-6 in Neurospora crassa. Fungal Genet. Biol. 2010;47:352–363. doi: 10.1016/j.fgb.2009.11.004.
    1. Wang Y., Hu X., Fang Y., Anchieta A., Goldman P.H., Hernandez G., Klosterman S.J. Transcription factor VdCmr1 is required for pigment production, protection from UV irradiation, and regulates expression of melanin biosynthetic genes in Verticillium dahliae. Microbiology. 2018;164:685–696. doi: 10.1099/mic.0.000633.
    1. Lopes L.C.L., da Silva M.I.D., Bittencourt V.C.B., Figueiredo R.T., Rollin-Pinheiro R., Sassaki G.L., Bozza M.T., Gorin P.A.J., Barreto-Bergter E. Glycoconjugates and polysaccharides from the Scedosporium/Pseudallescheria boydii complex: Structural characterisation, involvement in cell differentiation, cell recognition and virulence. Mycoses. 2011;54:28–36. doi: 10.1111/j.1439-0507.2011.02105.x.
    1. Thywißen A., Heinekamp T., Dahse H.-M., Schmaler-Ripcke J., Nietsche S., Zipfel P.F., Brakhage A.A. Conidial Dihydroxynaphthalene Melanin of the Human Pathogenic Fungus Aspergillus fumigatus Interferes with the Host Endocytosis Pathway. Front. Microbiol. 2011;2:96. doi: 10.3389/fmicb.2011.00096.
    1. Akoumianaki T., Kyrmizi I., Valsecchi I., Gresnigt M.S., Samonis G., Drakos E., Boumpas D., Muszkieta L., Prevost M.-C., Kontoyiannis D.P., et al. Aspergillus Cell Wall Melanin Blocks LC3-Associated Phagocytosis to Promote Pathogenicity. Cell Host Microbe. 2016;19:79–90. doi: 10.1016/j.chom.2015.12.002.
    1. Liu D., Wei L., Guo T., Tan W. Detection of DOPA-melanin in the dimorphic fungal pathogen Penicillium marneffei and its effect on macrophage phagocytosis in vitro. PLoS ONE. 2014;9:e92610. doi: 10.1371/journal.pone.0092610.
    1. Staerck C., Vandeputte P., Gastebois A., Calenda A., Giraud S., Papon N., Bouchara J.P., Fleury M.J.J. Enzymatic Mechanisms Involved in Evasion of Fungi to the Oxidative Stress: Focus on Scedosporium apiospermum. Mycopathologia. 2018;183:227–239. doi: 10.1007/s11046-017-0160-6.
    1. Staerck C., Tabiasco J., Godon C., Delneste Y., Bouchara J.-P., Fleury M.J.J. Transcriptional profiling of Scedosporium apiospermum enzymatic antioxidant gene battery unravels the involvement of thioredoxin reductases against chemical and phagocytic cells oxidative stress. Med. Mycol. 2019;57:363–373. doi: 10.1093/mmy/myy033.
    1. Wang Y., Casadevall A. Decreased susceptibility of melanized Cryptococcus neoformans to UV light. Appl. Environ. Microbiol. 1994;60:3864–3866. doi: 10.1128/aem.60.10.3864-3866.1994.
    1. Dadachova E., Bryan R.A., Howell R.C., Schweitzer A.D., Aisen P., Nosanchuk J.D., Casadevall A. The radioprotective properties of fungal melanin are a function of its chemical composition, stable radical presence and spatial arrangement. Pigment Cell Melanoma Res. 2008;21:192–199. doi: 10.1111/j.1755-148X.2007.00430.x.
    1. Casadevall A., Cordero R.J.B., Bryan R., Nosanchuk J., Dadachova E. Melanin, Radiation, and Energy Transduction in Fungi. Microbiol. Spectr. 2017;5:5.02.05. doi: 10.1128/microbiolspec.FUNK-0037-2016.
    1. van de Sande W.W.J., de Kat J., Coppens J., Ahmed A.O.A., Fahal A., Verbrugh H., van Belkum A. Melanin biosynthesis in Madurella mycetomatis and its effect on susceptibility to itraconazole and ketoconazole. Microbes Infect. 2007;9:1114–1123. doi: 10.1016/j.micinf.2007.05.015.
    1. Mello T.P., Oliveira S.S.C., Branquinha M.H., Santos A.L.S. Decoding the antifungal resistance mechanisms in biofilms of emerging, ubiquitous and multidrug-resistant species belonging to the Scedosporium/Lomentospora genera. Med. Mycol. 2022;60:myac036. doi: 10.1093/mmy/myac036.
    1. Ninomiya Y., Suzuki K., Ishii C., Inoue H. Highly efficient gene replacements in Neurospora strains deficient for nonhomologous end-joining. Proc. Natl. Acad. Sci. USA. 2004;101:12248–12253. doi: 10.1073/pnas.0402780101.
    1. Villalba F., Collemare J., Landraud P., Lambou K., Brozek V., Cirer B., Morin D., Bruel C., Beffa R., Lebrun M.-H. Improved gene targeting in Magnaporthe grisea by inactivation of MgKU80 required for non-homologous end joining. Fungal Genet. Biol. 2008;45:68–75. doi: 10.1016/j.fgb.2007.06.006.

Source: PubMed

3
Abonnere