Skip to main content
Advertisement
Browse Subject Areas
?

Click through the PLOS taxonomy to find articles in your field.

For more information about PLOS Subject Areas, click here.

  • Loading metrics

A p53-like transcription factor, BbTFO1, contributes to virulence and oxidative and thermal stress tolerances in the insect pathogenic fungus, Beauveria bassiana

  • Juan-Juan Wang ,

    Roles Conceptualization, Methodology, Software

    wjj880414@163.com (JJW); qiulei.2005@163.com (LQ)

    Affiliation School of Biological Science and Technology, University of Jinan, Jinan, China

  • Ya-Ping Yin,

    Roles Data curation, Writing – original draft

    Affiliation School of Biological Science and Technology, University of Jinan, Jinan, China

  • Ji-Zheng Song,

    Roles Writing – original draft

    Affiliations State Key Laboratory of Biobased Material and Green Papermaking, Qilu University of Technology, Shandong Academy of Sciences, Jinan, China, Maize Research Institute, Shandong Academy of Agricultural Sciences, Jinan, China

  • Shun-Juan Hu,

    Roles Writing – original draft

    Affiliation State Key Laboratory of Biobased Material and Green Papermaking, Qilu University of Technology, Shandong Academy of Sciences, Jinan, China

  • Wen Cheng,

    Roles Software

    Affiliation Maize Research Institute, Shandong Academy of Agricultural Sciences, Jinan, China

  • Lei Qiu

    Roles Conceptualization, Writing – review & editing

    wjj880414@163.com (JJW); qiulei.2005@163.com (LQ)

    Affiliation State Key Laboratory of Biobased Material and Green Papermaking, Qilu University of Technology, Shandong Academy of Sciences, Jinan, China

Abstract

The p53-like transcription factor (TF) NDT80 plays a vital role in the regulation of pathogenic mechanisms and meiosis in certain fungi. However, the effects of NDT80 on entomopathogenic fungi are still unknown. In this paper, the NDT80 orthologue BbTFO1 was examined in Beauveria bassiana, a filamentous entomopathogenic fungus, to explore the role of an NDT80-like protein for fungal pest control potential. Disruption of BbTFO1 resulted in impaired resistance to oxidative stress (OS) in a growth assay under OS and a 50% minimum inhibitory concentration experiment. Intriguingly, the oxidation resistance changes were accompanied by transcriptional repression of the two key antioxidant enzyme genes cat2 and cat5. ΔBbTFO1 also displayed defective conidial germination, virulence and heat resistance. The specific supplementation of BbTFO1 reversed these phenotypic changes. As revealed by this work, BbTFO1 can affect the transcription of catalase genes and play vital roles in the maintenance of phenotypes associated with the biological control ability of B. bassiana.

Introduction

The p53-like transcription factor (TF) superfamily includes the NDT80/PhoG-like DNA-binding family (http://pfam.xfam.org/family/PF05224), which has only been discovered in unikont lineages [1]. NDT80-like proteins have different functions in different fungi, such as regulation of sexual development, meiosis, filamentation, virulence, drug resistance, programmed cell death and the response to nutrient stress [1, 2]. The NDT80 protein of Saccharomyces cerevisiae is considered to be the founding member of the NDT80 family within the p53-like superfamily [1] and a key regulator of sporulation and meiosis [3]. XprG, also known as an NDT80-like protein, positively regulates mycotoxin generation, carbon depletion-induced apoptosis, and extracellular protease expression in Aspergillus nidulans [4]. This protein is similar to Neurospora crassa VIB-1, which plays a role in gene expression related to the positive regulation of extracellular protease generation as well as heterokaryon incompatibility-related apoptosis [2, 3, 5]. In Candida albicans, the roles of three NDT80-like proteins have been extensively studied. As regulators, CaNDT80 and CaRep1 control transporters (CDR1 and MDR1), which can eliminate drugs [6, 7]. Mutations in CaNDT80 and CaRON1 affect the induction of hyphal growth [7]. The CaNDT80 mutant also shows weakened virulence in a mouse model of infection [8]. NDT80 homologue mutations have certain impacts on secondary metabolism, including asexual or sexual spore pigmentation. N. crassa Δvib-1 mutants show pinkish conidia but not orange conidia (asexual spores); by contrast, mutants of XprG in A. fumigatus and A. nidulans possess pale conidia [3, 9]. Moreover, NDT80-like proteins are involved in fungal nutrition sensing. For example, suppressor of fusion (Suf), the Ndt80 homologue in Fusarium oxysporum, plays a vital role in regulating anastomosis depending on the nutrient [10]. As far as Trichoderma reesei is concerned, this protein also plays an important role in activating the N-acetylglucosamine (GlcNAc) gene cluster [2, 11]. C. albicans Rep1 and A. nidulans XprG are also required for GlcNAc catabolism [7].

The genus Beauveria, including the species B. bassiana, represents a strong insect pathogen with a wide range of hosts and has been prepared as a candidate mycoinsecticide with environmental friendliness [12]. As an important filamentous insect fungal pathogen, B. bassiana is the typical model for examining the pathogenic mechanism and growth of fungi together with their interactions with the host [13, 14]. Recently, some TFs in B. bassiana have been characterized. Typically, the BbMbp1 gene facilitates the transition of B. bassiana morphology with regard to saprophytic and pathogenic growth through a variety of genetic pathways [15]. In addition, in B. bassiana, the MADS-box TF Bbmcm1 plays a vital role in the regulation of the cell cycle, virulence and integrity [16]. Moreover, BbOhmm affects the homeostasis of reactive oxygen species (ROS) combined with the available oxygen to control BbSre1 transcriptional activity, thus regulating the homeostasis of mitochondrial iron, respiration-related genes and haem production to adapt to hypoxic conditions [14].

In our study, it is shown that the NDT80-like protein in B. bassiana plays an important role in oxidative stress (OS) resistance by impacting the expression of key catalase genes. Therefore, this protein was designated TFO1 (transcription factor regulating oxidation) in B. bassiana. This study was conducted to examine how BbTFO1 affects B. bassiana phenotypic changes related to its biological control ability through multi-phenotypic analyses of specific gene deletion or complementation strains combined with the parental WT. The results indicate that BbTFO1 plays an important role in regulating antioxidant effects, thermotolerance and pathogenesis in insect fungal pathogens.

Materials and methods

Phylogenetic analysis of BbTFO1

In NCBI databases, using the S. cerevisiae NDT80 sequence as a query, BLAST analysis based on NCBI was carried out to identify the genome from ARSEF 2860, the wild-type B. bassiana strain (WT), and some typical filamentous fungi. Next, each NDT80 homologous protein sequence found in the fungi was aligned to compare the structures and analyse the phylogenetics using MEGA7 (http://www.megasoftware.net).

Construction of BbTFO1 deletion and complementation mutants

BbTFO1 deletion together with a complementation mutant was created via Agrobacterium-mediated transformation [17]. The synthetic plasmids for homologous transformation were transformed by Agrobacterium tumefaciens and mobilized into B. bassiana according to a previous procedure [18]. Using the paired primers (Table 1), PCR was carried out to clone BbTFO1 5′ or 3′ coding/flanking fragment sequences based on the WT genome, with TaKaRa La Taq® (TaKaRa, Japan) used as the catalyst, followed by specific enzyme digestion and insertion to the target plasmid to form target gene deletion plasmids. BbTFO1 was knocked out in the WT by homogenously recombining the 5′ and 3′ fragments with a bar marker. The gene deletion mutant was initially screened by herbicide (200 μg ml−1 phosphinothricin) resistance [19].

thumbnail
Table 1. Paired primers used for gene cloning, deletion, complement and expression.

https://doi.org/10.1371/journal.pone.0249350.t001

For the reverse complement of the BbTFO1 deletion mutants, using Gateway BP Clonase II Enzyme Mix (Shanghai, China), primers (Table 1) were used to clone the entire BbTFO1 coding sequence based on the WT genome, followed by insertion into the p0380-sur-gateway for exchange for gateway fragments, and finally, p0380-sur-BbTFO1 was formed [20]. The plasmid was transformed in deletion via Agrobacterium-mediated transformation to form complementation strains of BbTFO1, and a putative complementary strain was selected by chlorimuron ethyl (10 μg ml−1) tolerance. Quantitative real-time PCR (qPCR) and PCR were used to detect the expected deletion and complementation strains.

Multi-phenotypic assays

Phenotypic assays were performed to detect the potential heterogeneities in the phenotype of ΔBbTFO1 compared with controls (WT and ΔBbTFO1/BbTFO1). All assays were repeated three times.

An appropriate amount of conidial suspension (1 μl, 107 conidia ml−1, identical concentration hereafter) harvested from SDAY (4% glucose, 1% yeast extract, 1% peptone plus 1.5% agar) plates was spotted onto media containing CPZ (3% sucrose, 0.1% K2HPO4, 0.3% NaNO3, 0.05% MgSO4, 0.001% FeSO4, 0.05% KCl combined with 1.5% agar), SDAY and 1/4 SDAY (containing 1/4 of every SDAY nutrient). The diameter of each colony was determined, and photos were taken at day 7 after incubation at temperatures of 25, 30 and 35°C. Additionally, the cross-sectional size of the colony was determined to be the growth rate index of diverse plates.

For chemical stresses, an appropriate amount of conidium suspension (1 μl) was added to CPZ plates containing different chemical stresses, including oxidative (2 mM H2O2 along with 0.2 mM menadione), cell wall (30 μg ml−1 Congo red) and osmotic (0.5 M NaCl) stressors. The colony diameters were measured and photographed as described earlier, and a CPZ plate without any chemical stress was used as the control.

To better assess how OS affected deletion mutant growth, an appropriate amount of conidium suspension (1 μl) was added to CPZ media containing gradient menadione doses (5–50 μM) and H2O2 (1–4 mM). The diameters of all colonies under OS and the control were measured to calculate the colony area after 7 days of incubation at 25°C. A minimum inhibitory concentration (MIC50) value of H2O2 and menadione that suppressed 50% colony growth was adopted as the OS resistance index in every strain.

Several conidial properties related to biocontrol potential were detected. The conidiation on SDAY was measured according to a related description [19]. Briefly, an appropriate amount of conidial suspension (100 μl) was uniformly added onto SDAY media, followed by incubation at 25°C. Then, three rounded cultures were collected from every plate every day by a cork borer (6 mm diameter) from day 5 onwards. The conidia on rounded cultures were completely added to 0.02% Tween-80 (1 ml). Later, conidial suspension content was assessed, followed by conversion to conidia count in the plate culture per unit area. In germination broth (GB: 0.5% peptone and 2% sucrose in 0.02% Tween-80), the median germination time (GT50, h) necessary to achieve 50% conidial germination under shaking at 110 rpm and 25/30°C was evaluated to be the conidial viability index at different temperatures [21].

The yield of blastospores was quantified in nitrogen-limited broth (NLB: 0.4% NH4NO3, 4% glucose, 0.3% MgSO4, 0.3% KH2PO4). Briefly, 40 μl conidial suspensions were added to 20 ml of NLB and cultured at 25°C with shaking (110 rpm). The number of blastospores was counted after 3 days, and then the blastospore production was transformed into cell count per ml culture medium.

Conidial thermotolerance was evaluated using a previously described method [22]. Conidia tolerance to high temperature was defined as the median lethal time (LT50, min) when conidia were treated with 0–120 min heat stress exposure at 45°C. In this process, 100 μl samples were taken from each tube at 15 min intervals, and then 1 ml of GB was added. Afterwards, the sample was subjected to 24 h of shaking at 110 rpm and 25°C, and a cytometer was used to evaluate the conidial germination rate microscopically.

Fungal virulence

Galleria mellonella larvae were used to assay fungal virulence according to immersion and injection approaches. The larvae were soaked in conidia suspension for 10 seconds and dry paper towels were used to remove excess liquid from the larvae. For intrahaemocoel injection, 1 μl conidial suspension (5 × 105 conidia ml−1) was added to the larval haemocoel. The two control groups were exposed to sterile 0.02% Tween-80. Each experiment was carried out three times, with 30 larvae being adopted in every experiment. Moreover, the number of dead larvae was measured at intervals of half a day. After analysing the time-mortality trend, the LT50 (d) for every strain was determined against every larval group. After curing under the optimal conditions for 6 days, the fungal growth on the surface of the mummified cadavers was photographed. During fungal virulence assessment, haemolymph samples were extracted from larvae that survived 72 h after injection, and hyphal bodies were formed in vivo, as seen by laser scanning confocal microscopy (LEICA DMi8, Germany).

Transcript profile analysis of antioxidant-related genes

The CAT family, SOD family and Bbohmm involved in OS responses were analysed by qPCR. In brief, 100 μl of conidial suspensions were inoculated into CPZ medium containing H2O2 (2 mM). After growing at 25°C for 5 days, an RNAiso Plus Kit (TaKaRa, Dalian, China) was utilized to extract total RNA from ΔBbTFO1 and the WT strain. Then, a PrimeScriptRT reagent kit (TaKaRa) was utilized for cDNA synthesis by reverse transcription. The resultant cDNA was then diluted at 10 ng ml-1 and adopted as a template in qPCR by the use of paired primers (Table 1) [23]. For B. bassiana, 18S rRNA was applied as the internal standard of each transcript. The gene transcriptional level was determined and repeated three times. The relative transcription level of the antioxidant-related genes was calculated as the fold change of the mutant strain transcript relative to the WT strain transcript in stress culture.

Statistical methods

One-way ANOVA, as well as Tukey’s HSD test, were used to analyse phenotype heterogeneities in the WT compared with control strains.

Results

BbTFO1 characteristics in and generation of BbTFO1 mutants

Using the NDT80 sequence of S. cerevisiae as a query, the annotated B. bassiana ARSEF 2860 strain genome was searched against the NCBI database, generating the most highly related sequence of 502 amino acids (identity: 31%; E-value: 3e-05). Phylogenetic analysis (S1A Fig) showed that the BbTFO1 sequence had the closest relationship with NDT80 from N. crassa (XP 011394327) and T. reesei (XP 006961996) because they were located on the same branch. Similar to the transcription factor NDT80 of S. cerevisiae, BbTFO1 had an NDT80/PhoG-like DNA-binding domain (residues 131–318). Successful BbTFO1 deletion and complementation were verified through PCR and qPCR using paired primers to examine correct recombination events (S1B–S1D Fig). Additionally, BbTFO1 transcription was not detected in the deletion strains, but similar transcription levels were detected in the WT and complementary strains (S1D Fig).

Contribution of BbTFO1 to chemical stress tolerance

The growth of deletion, complementation, and WT strains in different nutrient and chemical stress media is shown in Fig 1. The colony size of deletion strains was similar to that of control strains, and there was no significant difference with those in the rich medium SDAY, nutrient-deficient medium 1/4 SDAY, or minimal medium CPZ (Tukey’s HSD, P > 0.05) (Fig 1). Similar results were also found in CPZ medium containing NaCl and Congo red (Tukey’s HSD, P > 0.05). Nevertheless, ΔBbTFO1 showed markedly higher sensitivity to OS than the control strains (Tukey’s HSD, P < 0.05). The above findings indicate that BbTFO1 plays an essential role in maintaining OS tolerance but not in responding to other nutritional and chemical stresses.

thumbnail
Fig 1. The growth of fungal colonies on SDAY, 1/4 SDAY, CPZ and CPZ containing NaCl (0.5 M), H2O2 (2 mM), menadione (MND: 0.2 mM) and Congo red (CR: 3 μg ml−1).

The media were incubated for a period of 7 days at 25°C, and then photos were taken. Note that the ΔBbTFO1 strains of BbTFO1 grew smaller than the WT and ΔBbTFO1/BbTFO1 strains on CPZ containing H2O2 and menadione.

https://doi.org/10.1371/journal.pone.0249350.g001

Transcriptional changes in antioxidant-related genes in ΔBbTFO1

The colony size of deletion strains decreased by 41% and 63% compared with the control in CPZ medium containing menadione and H2O2, respectively (Fig 2A). However, the colony size of the deletion strain was not significantly different from that of the control in CPZ medium (Fig 2A). ΔBbTFO1 showed markedly high sensitivity to OS compared with controls. According to the MIC50 (Fig 2B), the deletion mutant also displayed high sensitivity to the oxidants H2O2 (19%) and menadione (23%). To explain the reduced tolerance of ΔBbTFO1 to oxidative stress, the transcriptional levels of some antioxidant-related genes were analysed by qPCR. Among them, four of six CAT family genes and Bbohmm showed transcriptional downregulation compared with WT. Significantly, two key catalases, cat2 and cat5, were transcriptionally repressed by 68% and 54%, respectively (Fig 2C), in response to OS in ΔBbTFO1 [23]. However, three of five SOD family genes showed transcriptional upregulation compared with WT. These results suggest that ΔBbTFO1 sensitivity to OS is accompanied by the transcriptional downregulation of these two key catalases.

thumbnail
Fig 2. Disruptions to BbTFO1 affect OS resistance and antioxidant-related enzyme gene expression in B. bassiana.

(A) Colony size was measured on the seventh day of culture at 25°C after inoculation. (B) MIC50 estimates for two oxidants (H2O2 and menadione) to inhibit 50% fungal growth in diverse strains on CPZ at 25°C. (C) Relative antioxidant-related gene transcript levels of ΔBbTFO1 strains compared with WT cultured for 5 days at 25°C on CPZ containing H2O2 (2 mM). The bar marked with an asterisk in each group shows a significant difference (Tukey’s HSD, P <0.05). Error bars: SD of 3 repeated assays.

https://doi.org/10.1371/journal.pone.0249350.g002

Role of BbTFO1 in conidiation and conidial quality

For conidiation, the ΔBbTFO1 strain was decreased by 44%, 22% and 21% on days 5, 6, and 7 (Fig 3A), respectively, which was significantly different from the controls (Tukey’s HSD, P < 0.05). However, final ΔBbTFO1 conidiation was similar to that of the control strain (Tukey’s HSD, P > 0.05). These results showed that the deletion of BbTFO1 affected conidiation growth but not conidiation ability. Moreover, the production of blastospores at 25°C showed no significant difference in ΔBbTFO1 compared with the corresponding controls (Tukey’s HSD, P > 0.05, Fig 3B).

thumbnail
Fig 3.

BbTFO1 disruption affects fungal conidiation and conidial quality (A) Conidiation was counted daily from 5 days to 8 days on SDAY plates, which was started with smearing an appropriate amount of conidial suspension (100 μl, 107 conidia ml−1). (B) The yield of blastospores was counted in NLB after 3 days of incubation at 25°C. (C) The GT50 (h) necessary to achieve 50% conidial germination at 25°C and 35°C. (D) Colony size in different strains grown at 30°C and 35°C for 7 days on CPZ medium. (E) The LT50 (min) of heat stress resistance in conidia at 45°C. The bar marked with an asterisk for every three-bar group showed a significant difference (Tukey’s HSD, P < 0.05). Error bars: SD of 3 replicates.

https://doi.org/10.1371/journal.pone.0249350.g003

Despite having no impact on conidiation capacity and the production of blastospores, deletion of BbTFO1 exerts a marked influence on conidial germination and thermotolerance. The ΔBbTFO1 mutant showed GT50 values of 2.1 h and 2.5 h longer than those of the WT at 25°C and 35°C, respectively (Fig 3C). The colony sizes of ΔBbTFO1 decreased by 26% and 65% compared with WT at 30°C and 35°C, respectively (Fig 3D). Moreover, heat stress resistance of ΔBbTFO1 at 45°C decreased significantly compared with that of the WT strain, and its LT50 decreased by 67% (Fig 3E). These data revealed the role of BbTFO1 in conidial germination and its contribution to thermotolerance but not in the production of blastospores.

Contribution of BbTFO1 to insect virulence

For normal infection via the cuticular penetration method, time-mortality trends and means (± SD) LT50 of G. mellonella larvae following conidial suspension immersion are shown in Fig 4A and 4B. On the ninth day after immersion, the mortality of the WT and complementation strain was close to 100%, while that of ΔBbTFO1 was approximately 80% (Fig 4A). The mean (± SD) LT50 values of the WT and complementation strains against larvae were 6.0 ± 0.2 and 5.9 ± 0.2 days via normal cuticle penetration (Fig 4B), respectively. The mean (± SD) LT50 values of the ΔBbTFO1 mutant against larvae were extended until 7.3 ± 0.3 days (Fig 4B). For intrahemocoel injection, the mortality of the WT and complementation strains were close to 100% on the fifth day after injection (Fig 4C); the mean (± SD) LT50 values of the WT and complementation were 3.6 ± 0.2 and 3.4 ± 0.1 days, respectively; and the mean LT50 of ΔBbTFO1 strains against the larvae increased to 4.0 ± 0.1 days (Fig 4D). After 6 days of preservation under the optimal conditions, the fungal growth on larvae infected with the control strain was greater than that on the larvae killed by ΔBbTFO1 (Fig 4E).

thumbnail
Fig 4. BbTFO1 deletion leads to defects in fungal pathogenicity.

(A, C) G. mellonella larval survival trend and (B, D) LT50 values after immersion in the 107 conidia ml−1 suspension to normally infect the cuticle and injection with ~500 conidia/larva through the haemocoel to achieve cuticle-bypassing infection. (E) Fungal outgrowth images of ΔBbTFO1 (DM), ΔBbTFO1/BbTFO1 (CM) and WT on the cadaver surface at 6 days after death. (F) Microscopic images (scale: 10 μm) of hyphal bodies within haemolymph samples collected from the surviving larvae at 72 h after injection. Filamentous cells were hyphal bodies, while subspherical and spherical cells were host haemocytes. The bar marked with an asterisk of the three-bar group showed a significant difference (Tukey’s HSD, P < 0.05). Error bars: SD of 3 replicates.

https://doi.org/10.1371/journal.pone.0249350.g004

To probe the potential reason for the reduced virulence by cuticle-bypassing infection with ΔBbTFO1, haemolymph samples obtained from larvae that survived 72 h after injection were analysed. As a result, there were rich hyphal bodies in the control strains but rare hyphal bodies within the haemolymph samples from the larvae following injection with ΔBbTFO1 conidia (Fig 4F).

Discussion

Compared with the controls, the ΔBbTFO1 strains showed significant defects in tolerance to OS, together with certain typical defects in phenotype related to the biological control potential of fungi, such as virulence and conidial quality. Typically, two catalase genes were markedly downregulated in the deletion mutants. The complicated phenotypic changes caused by the deletion of BbTFO1 are discussed below.

First, B. bassiana was subjected to biotic and abiotic OS in the field [24], which negatively affects the insect host infection capacity of fungi [25, 26]. Recently, more than 24 antioxidant enzymes showing dispersion within diverse families have been analysed for B. bassiana [27]. Some SODs in entomopathogenic fungi have been demonstrated to be crucial factors of fungal biocontrol potential [28]. Five members of the SOD family characterized in B. bassiana showed that Sod2 and Sod3 played a dominant role in fungal SOD activity [28, 29]. Each type of peroxide can be decomposed by members of the peroxidase (POD) family, including larger hydrogen peroxide, H2O2 and xenobiotics. KatG1, a bifunctional catalase-peroxidase, has been suggested to display POD and CAT activities, which play a vital role in chemical OS tolerance in Metarhizium acridum [24]. Both the GRX-GLR and TRX-TRR families have been examined to analyse their impacts on the antioxidation of B. bassiana [30, 31]. Some strong antioxidant genes have been adopted for entomopathogenic fungal genetic engineering for the sake of enhancing biological control abilities [32]. Overexpression of cytoplasmic manganese core superoxide dismutase (BbSod2) in B bassiana significantly increased antioxidant capacity, UV-B virulence and tolerance [33].

In addition, catalase family members decompose H2O2. Knockout mutants of catalases have shown discrete phenotypes in susceptibility to oxidative, heat, and UV-B stress in B. bassiana [34]. Among them, cat5, cat1 and cat4 are influential virulence factors, while cat1 and cat4 show equal importance to the regulation of UV-B resistance of conidia, with cat2 ranking second. In addition, cat1 is the most important regulator of conidial heat tolerance. It is believed that the CAT family affects fungal virulence by changing other virulence-related phenotypes of B. bassiana [23, 27]. Prior studies have shown that total catalase activity decreased by 89% and 56% in Δcat2 and Δcat5, respectively, while this decrease was only 9–12% in additional B. bassiana catalase knockout mutants [23]. cat2 and cat5 exert viral parts in B. bassiana antioxidation of catalase family [23]. Moreover, BbTFO1 deletion led to four of six CAT family genes showing transcriptional downregulation compared with the WT. In particular, the two key catalases cat2 and cat5 were transcriptionally repressed by 68% and 54%, respectively. In ΔBbTFO1, the defect of colony growth under OS may be due to the repression of catalase gene expression. Therefore, it was speculated that BbTFO1 can contribute to fungal antioxidation by affecting the expression of catalase genes. Beyond expectation, the absence of BbTFO1 did not increase the sensitivity to cell wall stressors in B. bassiana, while for additional fungi, such as C. albicans, NDT80 deletion resulted in enhanced sensitivity to Congo red and SDS [2]. The lack of NDT80 resulted in a denser biofilm relative to WT and might initiate the TEC1-dependent compensatory response, mainly through the TEC1-ROB1 pathway [35].

Second, according to a previous report in C. albicans [2], NDT80 deletion mutants also showed sensitivity to heat stress in B. bassiana. Heat stress could promote cellular ROS generation, thereby damaging certain biomolecules, such as DNA, lipids or proteins [33, 36]. Antioxidant enzymes can scavenge diverse intracellular ROS [27, 37]. Typically, SOD can decompose superoxide anions and produce H2O2 and oxygen [38], while H2O2 can be additionally decomposed to oxygen and water via catalase [39]. In the present work, 4 catalase gene expression levels, especially cat2 and cat5, in ΔBbTFO1 were repressed, resulting in a decrease in fungal ability to decompose H2O2, which naturally caused the reduced tolerance of ΔBbTFO1 conidia to high temperature. In addition, BbTFO1 knockout mutants were demonstrated to suffer defects in the development of conidiation and conidial germination. Similarly, the XprG mutant was slightly slower to initiate conidiation development in A. nidulans [4]. Conidial quality is also considered an essential virulence factor of B. bassiana [40]. In this study, defects in conidial quality may be one of the reasons for the decrease in virulence.

Finally, BbTFO1 mutants showed reduced virulence to hosts in B. bassiana by immersion and injection, as observed in NDT80 mutants of C. albicans. The virulence loss of NDT80 mutants in C. albicans may be associated with the critical part of NDT80 in the control of hyphal growth and stress resistance [2, 8]. An NDT80 homologue in F. oxysporum (Suf, suppressor of fusion) is associated with horizontal transport of virulence gene-carrying small chromosomes under conditions with limited nutrients [10]. For B. bassiana, the host immune response will produce cytotoxic molecules such as ROS when it infects insects [41]. Host phagocytes also generate reactive oxygen intermediates (ROIs), which are toxic to a variety of microorganisms [42]. Based on this experiment, it was believed that the disruption of BbTFO1 leads to a reduction in the transcription levels of four catalases and complex phenotypic changes, which are responsible for the reduction in fungal virulence. Catalase is considered a potential virulence factor in pathogenic fungi [23]. Overexpression of one catalase led to increased virulence of B. bassiana [43]. Catalase deletion within A. fumigatus mycelium and conidia leads to a significant decrease in virulence to mice [44]. In Magnaporthe grisea, the absence of cat2 enhances fungal sensitivity to OS and significantly decreases virulence to barley [45]. An NDT80-like protein was further examined in A. fumigatus, a pathogenic fungus in animals, showing that XprG makes no difference in the virulence of mice [9]. The deletion of BbTFO1 leads to defects in antioxidant capacity and germination defects, indicating that BbTFO1 is a critical virulence factor of B. bassiana.

Conclusion

Our results confirmed the previously reported functional diversity of NDT80-like proteins in different fungi [7]. In addition, BbTFO1 plays a crucial role in antioxidant activity, thermotolerance, conidiation, conidial quality and virulence in B. bassiana. According to transcript analysis, BbTFO1 can positively affect the expression of two key catalases to influence the antioxidant capacity of fungi, following the impact of the B. bassiana biological control potential.

Supporting information

S1 Fig. Bioinformatics description of BbTFO1 and generation of its mutant strains.

(A) Phylogenetic tree of B. bassiana NDT80 homologues and additional fungi. The NCBI accession codes for all proteins, together with the sequence identities in B. bassiana, are provided in brackets after the fungal names. (B) Sketch map of BbTFO1 deletion strategy. (C) BbTFO1 deletion identification (lanes 1–3) by PCR. Lane M: DNA marker (bp). Lane CK: blank control. Lane 1: WT. Lane 2: ΔBbTFO1. Lane 3: ΔBbTFO1/BbTFO1. (D) Identification of BbTFO1 deletion by qPCR.

https://doi.org/10.1371/journal.pone.0249350.s001

(TIF)

References

  1. 1. Katz ME, Cooper S. Extreme Diversity in the Regulation of Ndt80-Like Transcription Factors in Fungi. G3-Genes Genomes Genetics. 2015; 5(12): 2783–92. pmid:26497142
  2. 2. Min K, Biermann A, Hogan DA, Konopka JB. Genetic Analysis of NDT80 Family Transcription Factors in Candida albicans Using New CRISPR-Cas9 Approaches. Msphere. 2018; 3(6): e00545–18. pmid:30463924
  3. 3. Hutchison EA, Glass NL. Meiotic Regulators Ndt80 and Ime2 Have Different Roles in Saccharomyces and Neurospora. Genetics. 2010; 185(4): 1271–82. pmid:20519745
  4. 4. Katz ME, Braunberger K, Yi G, Cooper S, Nonhebel HM, Gondro C. A p53-like transcription factor similar to Ndt80 controls the response to nutrient stress in the filamentous fungus, Aspergillus nidulans. F1000Res. 2013; 2: 72. pmid:24358888
  5. 5. Dementhon K, Iyer G, Glass NL. VIB-1 is required for expression of genes necessary for programmed cell death in Neurospora crassa. Eukaryot Cell. 2006; 5(12): 2161–73. pmid:17012538
  6. 6. Chen CG, Yang YL, Shih HI, Su CL, Lo HJ. CaNdt80 is involved in drug resistance in Candida albicans by regulating CDR1. Antimicrob Agents Chemother. 2004; 48(12): 4505–12. pmid:15561818
  7. 7. Katz ME. Nutrient sensing-the key to fungal p53-like transcription factors? Fungal Genet Biol. 2019; 124: 8–16. pmid:30579885
  8. 8. Sellam A, Askew C, Epp E, Tebbji F, Mullick A, Whiteway M, et al. Role of transcription factor CaNdt80p in cell separation, hyphal growth, and virulence in Candida albicans. Eukaryot Cell. 2010; 9(4): 634–44. pmid:20097739
  9. 9. Shemesh E, Hanf B, Hagag S, Attias S, Shadkchan Y, Fichtman B, et al. Phenotypic and Proteomic Analysis of the Aspergillus fumigatus ΔPrtT, ΔprG and Δ prGPrtT Protease-Deficient Mutants. Front Microbiol. 2017; 8: 2490. pmid:29312198
  10. 10. Shahi S, Fokkens L, Houterman PM, Rep M. Suppressor of fusion, a Fusarium oxysporum homolog of Ndt80, is required for nutrient-dependent regulation of anastomosis. Fungal Genet Biol. 2016; 95: 49–57. pmid:27531696
  11. 11. Kappel L, Gaderer R, Flipphi M, Seidl-Seiboth V. The N-acetylglucosamine catabolic gene cluster in Trichoderma reesei is controlled by the Ndt80-like transcription factor RON1. Mol. Microbiol. 2016; 99(4): 640–57. pmid:26481444
  12. 12. Wang C, Wang S. Insect Pathogenic Fungi: Genomics, Molecular Interactions, and Genetic Improvements. Annu Rev Entomol. 2017; 62(1): 73–90. pmid:27860524
  13. 13. Zhang SZ, Xia YX, Kim B, Keyhani NO. Two hydrophobins are involved in fungal spore coat rodlet layer assembly and each play distinct roles in surface interactions, development and pathogenesis in the entomopathogenic fungus, Beauveria bassiana. Mol Microbiol. 2011; 80(3): 811–26. pmid:21375591
  14. 14. He Z, Zhao X, Gao Y, Keyhani NO, Wang H, Deng J, et al. The fungal mitochondrial membrane protein, BbOhmm, antagonistically controls hypoxia tolerance. Environ Microbiol. 2020; 22(7): 2514–35. pmid:31894607
  15. 15. Ding JL, Lin HY, Feng MG, Ying SH. Mbp1, a component of the MluI cell cycle box-binding complex, contributes to morphological transition and virulence in the filamentous entomopathogenic fungus Beauveria bassiana. Environ Microbiol. 2020; 22(2): 584–97. pmid:31743555
  16. 16. Zhao X, Yang X, Lu Z, Wang H, He Z, Zhou G, et al. MADS-box transcription factor Mcm1 controls cell cycle, fungal development, cell integrity and virulence in the filamentous insect pathogenic fungus Beauveria bassiana. Environ Microbiol. 2019; 21: 3392–416. pmid:30972885
  17. 17. Wang JJ, Cai Q, Qiu L, Ying SH, Feng MG. The histone acetyltransferase Mst2 sustains the biological control potential of a fungal insect pathogen through transcriptional regulation. Appl Microbiol Biotechnol. 2017; 102(3): 1343–55. pmid:29275430
  18. 18. Qiu L, Wang JJ, Chu ZJ, Ying SH, Feng MG. Phytochrome controls conidiation in response to red/far-red light and daylight length and regulates multistress tolerance in Beauveria bassiana. Environ Microbiol. 2014; 16(7): 2316–28. pmid:24725588
  19. 19. Qiu L, Wei XY, Wang SJ, Wang JJ. Characterization of trehalose-6-phosphate phosphatase in trehalose biosynthesis, asexual development, stress resistance and virulence of an insect mycopathogen. Pest Biochem Physiol. 2020; 163: 185–92. pmid:31973856
  20. 20. Qiu L, Wang JJ, Ying SH, Feng MG. Wee1 and Cdc25 control morphogenesis, virulence and multistress tolerance of Beauveria bassiana by balancing cell cycle-required cyclin-dependent kinase 1 activity. Environ Microbiol. 2015; 17(4): 1119–33. pmid:24910927
  21. 21. Shao W, Cai Q, Tong SM, Ying SH, Feng MG. Rei1-like protein regulates nutritional metabolism and transport required for the asexual cycle in vitro and in vivo of a fungal insect pathogen. Environ Microbiol. 2019; 21(8): 2772–86. pmid:30932324
  22. 22. Zhu XG, Tong SM, Ying SH, Feng MG. Antioxidant activities of four superoxide dismutases in Metarhizium robertsii and their contributions to pest control potential. Appl Microbiol Biotechnol. 2018; 102(21): 9221–30. pmid:30120522
  23. 23. Wang ZL, Zhang LB, Ying SH, Feng MG. Catalases play differentiated roles in the adaptation of a fungal entomopathogen to environmental stresses. Environ Microbiol. 2013; 15(2): 409–18. pmid:22891860
  24. 24. Li G, Fan A, Peng G, Keyhani NO, Xin J, Cao Y, et al. A bifunctional catalase-peroxidase, MakatG1, contributes to virulence of Metarhizium acridum by overcoming oxidative stress on the host insect cuticle. Environ Microbiol. 2017; 19(10): 4365–78. pmid:28925548
  25. 25. Keyser CA, Fernandes éKK, Rangel DEN, Roberts DW. Heat-induced post-stress growth delay: A biological trait of many Metarhizium isolates reducing biocontrol efficacy? J Invertebr Pathol. 2014; 120: 67–73. pmid:24909120
  26. 26. Fernandes EKK, Rangel DEN, Braga GUL, Roberts DW. Tolerance of entomopathogenic fungi to ultraviolet radiation: a review on screening of strains and their formulation. Curr Genet. 2015; 61(3): 427–40. pmid:25986971
  27. 27. Zhang LB, Feng MG. Antioxidant enzymes and their contributions to biological control potential of fungal insect pathogens. Appl Microbiol Biotechnol. 2018; 102(12): 4995–5004. pmid:29704043
  28. 28. Li F, Shi HQ, Ying SH, Feng MG. Distinct contributions of one Fe- and two Cu/Zn-cofactored superoxide dismutases to antioxidation, UV tolerance and virulence of Beauveria bassiana. Fungal Genet Biol. 2015; 81: 160–71. pmid:25263710
  29. 29. Xie XQ, Li F, Ying SH, Feng MG. Additive Contributions of Two Manganese-Cored Superoxide Dismutases (MnSODs) to Antioxidation, UV Tolerance and Virulence of Beauveria bassiana. PloS One. 2012; 7(1): e30298. pmid:22279579
  30. 30. Mashamaite LN, Rohwer JM, Pillay CS. The glutaredoxin mono- and di-thiol mechanisms for deglutathionylation are functionally equivalent: implications for redox systems biology. Biosci Rep. 2015; 35: 1–10. pmid:25514238
  31. 31. Zhang LB, Tang L, Ying SH, Feng MG. Distinct roles of two cytoplasmic thioredoxin reductases (Trr1/2) in the redox system involving cysteine synthesis and host infection of Beauveria bassiana. Appl Microbiol Biotechnol. 2016; 100(24): 10363–74. pmid:27344592
  32. 32. Lovett B, St Leger RJ. Genetically engineering better fungal biopesticides. Pest Manag Sci. 2018; 74(4): 781–9. pmid:28905488
  33. 33. Xie XQ, Wang J, Huang BF, Ying SH, Feng MG. A new manganese superoxide dismutase identified from Beauveria bassiana enhances virulence and stress tolerance when overexpressed in the fungal pathogen. Appl Microbiol Biotechnol. 2010; 86(5): 1543–53. pmid:20094711
  34. 34. Ortiz-Urquiza A, Keyhani NO. Stress response signaling and virulence: insights from entomopathogenic fungi. Curr Genet. 2015; 61: 239–249. pmid:25113413
  35. 35. Glazier VE, Murante T, Murante D, Koselny K, Liu Y, Kim D, et al. Genetic analysis of the Candida albicans biofilm transcription factor network using simple and complex haploinsufficiency. PLoS Genet. 2017; 13(8): 25. pmid:28793308
  36. 36. Perez-Guzman D, Montesinos-Matias R, Arce-Cervantes O, Gomez-Quiroz LE, Loera O, Garza-Lopez PM. Reactive oxygen species production, induced by atmospheric modification, alter conidial quality of Beauveria bassiana. J Appl Microbiol. 2016; 121(2): 453–60. pmid:27107399
  37. 37. Tong SM, Feng MG. Phenotypic and molecular insights into heat tolerance of formulated cells as active ingredients of fungal insecticides. Appl Microbiol Biotechnol. 2020; 104(13): 5711–24. pmid:32405755
  38. 38. Alscher RG, Erturk N, Heath LS. Role of superoxide dismutases (SODs) in controlling oxidative stress in plants. J Exp Bot. 2002; 53: 1331–41. https://doi.org/10.1093/jexbot/53.372.1331 pmid:11997379
  39. 39. Schouten A, Tenberge KB, Vermeer J, Stewart J, Wagemakers L, Williamson B, et al. Functional analysis of an extracellular catalase of Botrytis cinerea. Mol Plant Pathol. 2002; 3(4): 227–38. pmid:20569330
  40. 40. Keyhani NO. Lipid biology in fungal stress and virulence: Entomopathogenic fungi. Fungal Biol. 2018; 122(6): 420–9. pmid:29801785
  41. 41. Trevijano-Contador N, Zaragoza O. Immune Response of Galleria mellonella against Human Fungal Pathogens. J Fungi. 2019; 5(1): 13. pmid:30587801
  42. 42. Sheehan G, Kavanagh K. Analysis of the early cellular and humoral responses of Galleria mellonella larvae to infection by Candida albicans. Virulence. 2018; 9(1): 163–72. pmid:28872999
  43. 43. Chantasingh D, Kitikhun S, Keyhani NO, Boonyapakron K, Thoetkiattikul H, Pootanakit K, et al. Identification of catalase as an early up-regulated gene in Beauveria bassiana and its role in entomopathogenic fungal virulence. Biol Control. 2013; 67(2):85–93. http://dx.doi.org/10.1016/j.biocontrol.2013.08.004
  44. 44. Paris S, Wysong D, Debeaupuis JP, Shibuya K, Philippe B, Diamond RD, et al. Catalases of Aspergillus fumigatus. Infect Immun. 2003; 71(6): 3551–62. pmid:12761140
  45. 45. Skamnioti P, Henderson C, Zhang Z, Robinson Z, Jane Gurr S. A Novel Role for Catalase B in the Maintenance of Fungal Cell-Wall Integrity During Host Invasion in the Rice Blast Fungus Magnaporthe grisea. Mol Plant-Microbe Interact. 2007; 20: 568–80. pmid:17506334