Next Article in Journal
Candida parapsilosis Colony Morphotype Forecasts Biofilm Formation of Clinical Isolates
Next Article in Special Issue
Host–Pathogen Interactions: Insects vs. Fungi
Previous Article in Journal
Characteristics, Risk Factors, and Survival Analysis of Candida auris Cases: Results of One-Year National Surveillance Data from Oman
Previous Article in Special Issue
Fungi Recorded on Folivorous Lepidoptera: High Diversity Despite Moderate Prevalence
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Screening and Virulence of the Entomopathogenic Fungi Associated with Chilo suppressalis Walker

by
Morteza Shahriari
1,
Arash Zibaee
1,*,
Seyyed Akbar Khodaparast
1 and
Mahmoud Fazeli-Dinan
2
1
Department of Plant Protection, Faculty of Agricultural Sciences, University of Guilan, Rasht 4199613779, Iran
2
Department of Medical Entomology and Vector Control, Health Sciences Research Center, Addiction Institute, School of Public Health, Mazandaran University of Medical Sciences, Sari 4815733971, Iran
*
Author to whom correspondence should be addressed.
J. Fungi 2021, 7(1), 34; https://doi.org/10.3390/jof7010034
Submission received: 21 October 2020 / Revised: 11 November 2020 / Accepted: 16 November 2020 / Published: 7 January 2021
(This article belongs to the Special Issue Host-Pathogen Interactions: Insects vs Fungi)

Abstract

:
The current study aimed to explore the entomopathogenic fungi associated with the larvae of Chilo suppressalis Walker, a serious pest of rice, in northern Iran. The collected specimens were cultured and identified through morphological and molecular methods. The 38 specimens were identified by microscopic examination and genetic sequencing of the ITS region as follows: twenty-one isolates of Beauveria bassiana, five isolates of Akanthomyces lecanii, four isolates of Akanthomyces muscarious, three isolates of Metarhizium anisopliae, two isolates of Hirsutella subulata, two isolates of Trichoderma sp. and one isolate of Aspergillus sp. All the identified isolates were treated on the larvae through bioassay, evaluating the amount of hydrophobin and the activities of proteases, chitinases and lipase to find their virulence. Moreover, the percentage of thermotolerant and cold activity of the isolates were tested to determine their environmental persistence. The overall results revealed the isolates of B. bassiana, including BBRR1, BBAL1 and BBLN1 as the most virulent and environmental adaptive isolates among the fungi associated with C. suppressalis.

1. Introduction

The rice striped stem borer, Chilo suppressalis Walker (Lepidoptera: Crambidae), is an economic pest of rice, annually causing significant losses in Asia, southern America and northern Africa [1]. The larvae feed intensively on rice stems and cause “whitehead” and “dead-heart” of the seedlings, which directly decrease the overall yield of rice [2]. The main control measure to suppress the C. suppressalis population is the wide-spraying of synthetic insecticides, including diazinon, Padan® and Reagent®. Nevertheless, C. suppressalis has developed resistance to these insecticides on one hand and resulted in environmental pollution, food residuals and toxicity on non-target organisms on the other hand [3,4]. These concerns should change the management strategies of chemical insecticides toward biocontrol agents like entomopathogens. Among the entomopathogens used to manage the population of insect pests, entomopathogenic fungi cause epizootics among insect pests and appear as the prevalent natural pathogens to regulate population fluctuations of pests and subsequent losses [5]. Their presence in almost all terrestrial and aquatic ecosystems, as well as way of infection by producing different extracellular enzymes and by releasing toxic secondary metabolites, has led to the success of entomopathogenic fungi to affect noxious arthropods in agriculture, forestry and livestock [5,6].
There are many reports on the efficacy of different entomopathogenic fungi, including Akanthomyces lecanii, Akanthomyces muscarious, Aspergillus spp., Beauveria bassiana, Isaria fumosorosea, Isaria sinclairii, Metarhizium anisopliae, Metarhizium rileyi, Nomuraea rileyi, Pecilomyces lilacinus and Purpureocillium lilacinum against lepidopteran pests such as Chilo suppressalis, Spodoptera litura, Spodoptera frugiperda, Spodoptera exigua, Ostrinia nubilalis, Helicoverpa armigera, Helicoverpa zea, Plutella xylostella, Duponchelia fovealis, Agrotis ipsilon, Pieris rapae, Trichoplusia ni, Ocinara varians, Galleria mellonella, Plodia interpunctella, Ephestia kuehniella [7,8,9,10,11,12,13,14,15,16,17,18,19]. These agents have generally shown to be safe for humans with the least effects on non-targets while they are relatively sensitive to environmental conditions, mainly heat, cold and UV radiation, so it is imperative to find isolates adaptable to these constraints for formulation and field application [20,21,22,23].
Exogenous isolates of the entomopathogenic fungi that were commercialized as pest biocontrol agents in different countries may be ineffective on some pests due to environmental suitability and strain differences related to the host [24]. Therefore, the application of local isolates may be a promising choice mainly in case of ecological suitability with pest species and lower hazards on non-target organisms compared to exotic strains [22,25,26,27]. These points were verified by several studies that demonstrate the virulence of isolates belonging to the same fungal species could be different because of genetic variations occurring in a specific geographical distribution [28,29,30]. The provinces of Guilan and Mazandaran are located in the north of Iran with high humidity, moderate annual temperatures and heavy rainfall, in which these conditions are appropriate for entomopathogenic fungi [5]. The rice fields of northern Iran, known as a reservoir of C. suppressalis [31], can represent ideal sites to study the existence of entomopathogenic fungi with natural enzootics to C. suppressalis, so the aims of our study were to; (a) isolate and identify different entomopathogenic fungi from fungus-infected C. suppressalis larvae, (b) evaluate the virulence of these fungi against the larvae of C. suppressalis, (c) examine the infection process of these isolates by the production of extracellular secretions and (d) compare the conidial germination of the fungal isolates after exposure to heat and cold.

2. Materials and Methods

2.1. Collection and Morphological Identification

The collection sites were all the municipal regions of Guilan and Mazandaran provinces in the north of Iran (Mazandaran and Guilan, Iran) with the highly cultivated area of rice. In each site, the remained stems of rice within the paddy fields were opened, and the infected larvae of C. suppressalis were collected and kept in sterile centrifuge tubes. The infected larvae were recognized according to the mycelial growth outside the larval body. Once the samples were transferred to the laboratory, the larvae were surface disinfected with sodium hypochlorite (2%) for 3 min and rinsed three times in sterile distilled water [27]. The larvae were then transferred on potato dextrose agar (PDA, Merck) plates and incubated at 25 °C for 2–4 days for fungal development. Afterward, the fungal mycelia were picked up and transferred to fresh PDA plates for purification. Finally, single-spore cultures were gathered according to the method described by Fang [32] and cultured on PDA slants. All collected specimens were inoculated on PDA plates and incubated at 25 °C in the dark for 14 days. For microscopic examination, mycelia and conidia from fungal specimens were mounted on a sliding glass and observed at 100× magnification on a phase-contrast microscope (Canon INC DS126311, Taiwan). Morphological identification of the specimens was made based on conidial morphology, shape, color and size based on the following literature: Akanthomyces spp. isolates [28,33,34], Beauveria spp. Isolates [28,34,35], Hirsutella spp. isolates [34,36,37,38] and Metarhizium spp. isolates [34,39].

2.2. Genomic DNA Extraction and PCR

DNA extraction was done using the protocol of Montero-Pau et al. [40]. Briefly, the mass mycelia of the specimen grown in PDA media were transferred to the 1.5 mL tubes containing 100 μL of alkaline lysis buffer (0.2 mM disodium ethylene diamide tetraacetic acid, 25 mM NaOH, pH 8.0, Merck) and centrifuged for 30 min at 2000× g. Then, the tubes were incubated at 95 °C for 30 min and cooled on ice for five min. Finally, 100 μL of Tris-HCl solution (Sigma-Aldrich, Vienna, Austria; 40 mM, pH 5.0) was added to the tubes, vortexed and maintained at −20 °C. The extracted solution was used as a template for PCR.
To amplify the internal transcribed spacers (ITS5-5.8S-ITS4), ITS5 (5′GGAAGTAAAAGTCGTAACAAGG3′) and ITS4 (5′-TCCTCCGCTTATTGATATGC-3′) primers were synthesized as previously described [41]. The PCR reaction mixture consisted of 12.5 μL of master mix (Including 10× PCR buffer. MgCl2, dNTPS TaqPolymerase, CinnaClone, Tehran, Iran), 7.5 μL of double-distilled H2O, 1 μL of each primer and 3 μL of DNA solution. PCR was carried out using a thermal cycler (Eppendorf Personal, Darmstadt, Germany) with the following reaction parameters: an initial denaturation for 2 min at 94 °C, 30 cycles of 94 °C for 30 s, 53 °C for 30 s and 72 °C for 1 min and a final extension for 5 min at 72 °C. Amplified PCR products were visualized by electrophoresis on 1% agarose gels. The PCR products were sent to a sequencing service company (Royan Zistagene Co., Tehran, Iran) for purification and sequencing. Finally, sequences were compared with other fungi using the BLAST search tool in NCBI (https://blast.ncbi.nlm.nih.gov/Blast.cgi).

2.3. Insect Rearing

The stock population of C. suppressalis was established by collecting the egg patches from rice fields of Amol, northern Iran. The eggs were kept in a growth chamber at 25 ± 2 °C, 85 ± 5% R.H. and 16:8 (L:D) h of photoperiod. The newly hatched larvae were transferred to the tubes (20 × 15 cm) supplied by rice seedlings of Hashemi variety. Quality of food was monitored every day, and the old cutting stems were replaced by the fresh ones [42]. Rearing was continued for three generations at the same controlled conditions.

2.4. Bioassay

Conidia of the two-week-old PDA cultures of the identified isolates were removed by a scalpel then suspended in sterile distilled water containing Tween-80 (0.02%). The concentrations of 102 to 108 conidia/mL from each isolate were separately prepared based on the preliminary tests. Early fourth instar larvae of C. suppressalis were randomly selected and separately dipped in the serial concentrations of each isolate while the control larvae were dipped in an aqueous solution of 0.02% Tween-80 alone. The bioassays were done in three replicates with ten larvae per replication, and the larvae were maintained at the rearing condition for the whole bioassay period. Mortality was recorded within 7 days, and LC50 values were determined using POLO-Plus software. For calculation of LT50, mortality was recorded until the death of all larvae at 108 conidia/mL concentration.

2.5. Hydrophobin Protein Extraction and Estimation

Hydrophobin content was determined according to the method described by Ying and Feng [43]. Conidia from the two-week-old cultures were added to 1 mL of 2% SDS aqueous solution containing β-mercaptoethanol (5%, Merck) and incubated in a boiling water bath for 10 min before being centrifuged at 22,000× g and 4 °C for 10 min. The supernatant was removed, and conidia were rinsed twice in distilled water to eliminate SDS (Merck) soluble proteins. Samples were incubated in 1 mL formic acid at zero temperature for 2 h before being centrifuged at 22,000× g and 4 °C for 10 min. The supernatant was transferred into fresh tubes, and 0.5 mL of distilled water was added to the samples. Afterward, 0.75 mL of 45% NaOH solution added to the mixture and maintained at 4 °C overnight. The proteins were separated from the supernatant by centrifugation at 22,000× g and 4 °C for 10 min. The extractable proteins of formic acid were rinsed twice with ethanol solution (3:1, v/v) and then dissolved in 2% SDS to quantitatively determine the amount of protein as mg/mL of conidia using the procedure of Lowry et al. [44].

2.6. Liquid Culture for Enzyme Production of the Isolates

The liquid media used for biochemical production of the extracellular enzymes contained; 0.02% of KH2PO4; 0.01% of CaCl2; 0.01% of MgSO4; 0.02% of Na2HPO4; 0.01% of ZnCl2 and 0.01% of yeast extract (Merck). The media were inoculated with 1 mL of 108 conidia/mL concentration of each isolate separately and 5% (weight) of larval cuticle was added to each flask containing liquid media. Then the flasks were kept on a rotatory shaker (70 rev/min) for 8 days at 25 ± 1 °C [3].

2.7. Sample Preparations for Enzymatic Assays

After 8 days, the culture flasks were harvested by centrifugation at 10,000× g for 30 min and washed in ice-cold Tris-HCl (25 mM, pH 8). Weighed mycelia were ground to a fine powder, suspended in DW, homogenized and centrifuged at 22,000× g and 4 °C for 30 min to obtain the supernatant of enzyme assay [12].

2.7.1. Assay of Proteases

Activities of subtilisin-like (Pr1) and trypsin-like (Pr2) as the two key fungal proteases were determined by the specific substrates of succinyl-(alanine) 2-prolinephenylalanine-p-nitroanilide and benzoylphenylalanine-valine-arginine-p-nitroanilide (Sigma-Aldrich, Co., Vienna, Austria), respectively. The reaction mixture contained 100 µL of Tris-HCl buffer (20 mM, pH 8), 30 µL of each substrate separately and 20 µL of enzyme solution. The mixture was incubated at 25 °C for 10 min, then 100 μL of trichloroacetic acid (TCA, 30%) was added, and the absorbance was recorded at 405 nm [3].

2.7.2. Lipase Assay

Lipase assay was done using the method of Tsujita et al. [45]. Fifty microliters of p-nitrophenyl butyrate (27 mM, Sigma-Aldrich, Co., Vienna, Austria) as substrate, 20 μL enzyme solution and 100 μL of Tris-HCl buffer (20 mM, pH 7) were incorporated and incubated at 37 °C for 5 min. Then, 100 μL of NaOH (1 N) was added to each tube, and the absorbance was recorded at 405 nm.

2.7.3. Endochitinase Assay

Twenty microliters of the enzyme solution were added to 50 μL of 0.5% colloidal chitin (Sigma-Aldrich, Co., Vienna, Austria) as substrate and 100 μL of Tris-HCl buffer (20 mM, pH 7). Then, the samples were incubated in a water bath of 30 °C for 60 min. Then, 100 μL of dinitrosalisylic acid (DNS, Sigma-Aldrich, Vienna, Austria) was added, the incubation was prolonged for 10 min at boiling water, and the absorbance was recorded at 545 nm [46].

2.7.4. Exochitinase Assay

The activity of exochitinase was assayed by 200 μL of p-nitrophenyl-N-acetyl-β-D-glucosaminide (pNPg; Sigma-Aldrich, Co., Vienna, Austria) solution (1 mg pNPg per mL of distilled water) as substrate, 25 μL of enzyme solution and 500 μL of Tris-HCl (25 mM, pH 7) which was incubated at 40 °C for 20 h. Then, the mixture was centrifuged at 13,000 rpm at 4 °C, and the supernatant was added to 200 μL of sodium tetraborate-NaOH buffer (125 mM, pH 10) before to read the absorbance at 400 nm. The extinction coefficient of 18.5 Mm−1 −cm−1 was considered for activity calculation based on the following formula:
Volume activity (U/mL) = [ΔOD (OD test–OD blank) × Vt × df]/(18.5 × t × 1.0 × Vs) where, Vt = total volume; Vs = sample volume; 18.5 = millimolar extinction coefficient of p-nitrophenol under the assay condition; 1.0 = lightpath length (cm); t = reaction time; and df = dilution factor [46].

2.8. Protein Assay

The method of Lowry et al. [44] was used to determine the content of protein in the provided samples. Twenty microliters of the enzyme solution were added into 100 µL of reagent (Ziest Chem. Co., Tehran, Iran) and incubated for 30 min before reading the absorbance at 545 nm.

2.9. Effects of Thermotolerance and Cold Activity on Conidial Germination

To measure thermotolerance for conidial germination, 100 μL of conidial suspensions (5 × 106 conidia/mL) from each isolate was transferred to 1.5 mL tubes and kept in a thermal cycler adjusted to 45 °C. After 1 h and 2 h, 20 μL of conidial suspensions were removed and plated (without spreading) on PDA. Finally, plates were maintained at 25 °C, and conidial germination was counted after 24 h by microscopic observation. Moreover, 20 μL of a conidial suspension (5 × 106 conidia/mL) was plated (without spreading) on PDA and kept at 5 °C to determine the germination after 7 and 14 days in cold condition. In both experiments, the conidial control suspensions were inoculated on PDA at 25 °C. The relative percent germination was estimated by comparing conidial germination to untreated isolates, and at least 100 conidia were counted for each treatment in every test [27].

2.10. Statistical Analysis

Probit analysis was done to determine LC50 and LT50 values at the corresponding 95% confidence interval (CI) values by using POLO-Plus software. Biochemical data and germination of conidia were compared by one-way analysis of variance (ANOVA) followed by Tukey’s test. Differences among control and treatments were statistically considered at a probability of less than 5% and marked by different letters.

3. Results and Discussion

3.1. Screening and Identification of Fungi

A total of 38 fungal specimens were collected from C. suppressalis larvae, which were naturally infected by fungi in the rice fields of northern Iran. The specimens were morphologically identified as Akanthomyces lecanii (×5 isolates), Akanthomyces muscarius (×4 isolates), Aspergillus sp. (×1 isolate), Beauveria bassiana (×21 isolates), Hirutella subulata (×2 isolates), Metarhizium anisopliae complex (×3 isolates) and Trichoderma sp. (×2 isolates) (Table 1, Figure 1). Among all specimens, it was the first report of the natural occurrence of H. subulata in Iran. In A. lecanii, conidiogenous cells were phialidic, phialides approximately small, length size of 11–16 μm and width size of 1.4–2 μm, aculeate and strongly tapering, solitary or in whorls 3–6, conidial shape ellipsoidal-cylindrical, length size 4.1–5.2 μm and width size of 1.3–2.1 μm (Table 1, Figure 1a). In A. muscarious, conidiogenous cells were phialidic; phialides burned straight on prostrate hyphae or on secondary branches, phialides generally tall and slender and longer than A. lecanii, length size of 28–35 μm and width size of 1.6–2 μm, conidial shape cylindrical and longer than A. lecanii, length size 6.5–9.5 μm and width size of 1.5–1.9 μm (Table 1, Figure 1b). The major difference between the species of Akanthomyces is the shape and size of phialides that our specimens matched accurately match the description given by Zare and Gams [33]. In Aspergillus sp. conidial shape was globose, length size 1.8–2.3 μm and width size of 1.8–2.3 μm (Table 1, Figure 1c). In addition, the Conidial shape of B. bassiana isolates was globose with a length size of 2.1–3.2 μm and width size of 2–3 μm; conidiogenous cells were phialidic; the phialides were flask-shaped, swollen at the base or near the base and tapering at the apex. In addition, the conidiogenous cells were usually solitary or in a cluster of up to five (Table 1, Figure 1d), which appeared typical of those described by other researchers [28,34,35]. The major difference between the species of Beauveria is the shape and size of conidia [28]. In H. subulata, conidiogenous cells were phialidic, phialides scattered, and the lower phialides were narrow ellipsoid; the conidial shape was ovoid and in a chain, length size 5.5–6.9 μm and width size of 3.9–5.1 μm (Table 1, Figure 1e). Our isolates were compared with Yoon et al. [37], and significant differences were not observed among them. The conidial shape of M. anisopliae isolates was oblong oval with a length size of 6.8–7.8 μm and a width size of 2.6–3.7 μm (Table 1, Figure 1f). Conidia were the only morphological particular that reliably distinguishes several Metarhizium species [39]. In Trichoderma sp. conidial shape was globose, length size 2.2–3.1 μm and width size of 2.1–2.9 μm (Table 1, Figure 1g).
The ITS5-5.8S-ITS4 region was used for molecular analysis, confirming identifications of the fungal isolates. The amplified ITS region from all specimens showed approximately 600 bp-sized fragments, and the samples were sequenced and compared in the GenBank database. Results of the ITS sequence data were consistent with those obtained using morphological studies. After submission to the GenBank database, the fungal isolates were renamed based on the given code (Table 1). In the current study, B. bassiana was the most frequent detected fungus in rice collected from fields similar to others studies that have shown B. bassiana as the most widespread entomopathogenic fungi in the endemic Moroccan forests of Argania spinosa, Switzerland, Spain, China and Southern Italy [22,47,48,49,50]. Moreover, B. bassiana has been reported as the major pathogen of insects in more than 200 species that have been identified from the soil and dead insects in nature [24,25,51,52].

3.2. Bioassay

Only the isolates of B. bassiana, M. anisopliae and H. subulata caused mortality against the larvae of C. suppressalis. A comparison of LC50 and LT50 values indicated the significant differences among the isolates. The most virulent, isolate BBLN1 (1 × 104 conidia/mL), had the least LC50 value, followed by BBAL1 (2.1 × 104 conidia/mL), BBRR1 (2.2 × 104 conidia/mL), BBLN2 (5.4 × 104 conidia/mL), BBAL3 (5.6 × 104 conidia/mL) and MASA (7.1 × 104 conidia/mL), while HSBL (1.6 × 106 conidia/mL), HSAL (7.9 × 105 conidia/mL) and BBLD5 (4.4 × 105 conidia/mL) showed the comparatively high LC50 values (Table 2). Moreover, the least LT50 values were obtained to be 2.71, 3.15, 3.45, 3.66 and 3.69 days for the larvae treated by BBRR1, BBLN1, BBAL1, BBAL4 and MASA, respectively (Table 3). These results revealed that BBRR1, BBLN1 and BBAL1 isolates of B. bassiana had higher efficacy than the other isolates on C. suppressalis larvae with a lesser concentration of conidia with a shorter time (days) to kill 50% of the larval population. Jandricic et al. [53] reported the higher virulence of B. bassiana isolates against the Myzus persicae Sulzer, Aphis gossypii Glover and Aulacorthum solani Kaltenbach (Hemiptera: Aphididae) compared to M. anisopliae isolates. Ramzi and Zibaee [12] showed that the two commercial isolates of B. bassiana and B. bassiana (BB1 and BB2) had the higher virulence against C. suppressalis larvae compared to A. lecanii, I. fumosoroseus and M. anisopliae. In addition, the higher virulence of the different isolates of B. bassiana and M. anisopliae has been observed on the boll weevil Anthonomus grandis Boheman (Coleoptera: Curculionidae) [29]. In our study, the least virulence of HSAL and HSBL as the two isolates of H. subulata were obtained compared to B. bassiana and M. anisopliae isolates, which may be correlated with low germination and sporulation rates in addition to the low activities of the extracellular enzymes of these isolates (see below) [54]. Finally, the isolates of A. lecanii and A. muscarious showed no mortality against C. suppressalis larvae. This case may be attributed to host–pathogen interaction between these isolates and the larvae of C. suppressalis, such as efficient attachment of conidia to the integument, negative impacts of integument composition with penetration tube of the fungi and immune responses of the larvae toward conidia. All these phenomena deserve detailed experiments to precisely elucidate the case.

3.3. Hydrophobin

The highest amounts of hydrophobin recorded in BBAL1, BBLD5, BBLD1, BBSI, BBBL1, BBLN1, BBLN2, HSAL and MASA, respectively (Table 4), while the least amounts of hydrophobin were in TSRT, ASAI and TSAH (Table 4). Entomopathogenic fungi achieve the nutrients at host bodies through the cuticle, so the first step of pathogenesis is adhesion to the integument. Therefore, the external surface of conidia has a fundamental protein with a hydrophobic rodlet layer that connects to the insect epicuticle [6]. Hydrophobins are a class of unique fungal proteins important in sporulation, pathogenesis, thermotolerant, growth and development of fungi [43,55,56]. Some studies reported that inhibition of hydrophobin gene expression negatively affected pigmentation, conidiation, hydrophobicity and virulence of entomopathogenic fungi [24,55,57]. Our findings revealed that the lesser amount of hydrophobin could be one of the reasons for no mortality of some isolates against the larvae of C. suppressalis. In fact, the proper attachment of the conidia to the host cuticle and subsequent germinations are the primarily important steps to effective infection by entomopathogenic fungi. The higher amounts of hydrophobin were obtained in the isolates with the more virulence-like BBLN1 and BBLN2. Although the higher amounts of hydrophobin were also found in HSAL with the least virulence, it should be noted that hydrophobin is not necessarily the main factor in the virulence of a fungus, but it only shows the better interaction with the host cuticle.

3.4. Extracellular Enzymes

The conidia of entomopathogenic fungi attach to the cuticle of host insects, germinate and penetrate to the hemocoel with the assistance of extracellular enzymes, such as chitinases, proteases and lipases [58]. Trypsin (Pr1) and subtilisin-like (Pr2) proteases are the primitive synthesized enzymes to simplify penetration of the hyphae into the host body. Then, synthesis of the chitinases increases the penetration efficiency [59], and finally, lipases involved in hydrolyzing lipid derivatives within the cuticle and facilitating the infection of host cells [60]. Our results revealed differences in the activities of extracellular enzymes between the fungal isolates. Isolates BBAL1, BBRR1, BBLN2, BBLN1 and BBLD2 demonstrated the highest activity of Pr1 while the least activity was observed in ASAI, TSRT and TSAH isolates (Figure 2). In the case of Pr2, BBRR1, BBLN2, BBLD2 and BBLN1, isolates showed the highest activity (Figure 2). The highest activity of lipase was recorded in BBLD4, BBSI, BBLN2 and BBRR1 isolates (Figure 3). BBRR1, BBLN2, BBAL1 and BBLN1 isolates showed the highest activity of exochitinase (Figure 4). In the case of endochitinase, the highest activity was obtained in BBRR1, BBLN2, BBBL1, BBAL1 and BBLN1 isolates (Figure 4). The higher Pr1 activity in the given isolates indicates the capability of protein digestion by these isolates in the initial stages of infection, so the efficiency of this enzyme may ensure the success of other enzymes to feasible penetration through insect cuticle. Charnley and St. Leger [61] believe in facilitating the cuticle infiltration by the proteases produced during invasion prior to chitinases during later steps. They concluded the major role of proteases in cuticle penetration compared to chitinases. Ramzi and Zibaee [12] demonstrated the different levels of proteinases, chitinase and lipase produced by B. bassiana, M. anisopliae, L. lecanii and I. fumosoroseus in the larvae C. suppressalis in which the isolates with the highest enzymatic activity led to the higher mortality Lu et al. [1] showed, the higher levels of protease and chitinase produced by ZJLSP09 isolate of Lecanicilium sp. in comparison with ZJLA07 and ZJLP08 isolates which were related to mortality in Diaphorina citri Kuwayana (Hemiptera: Psyllidae). Maqsoudi et al. [62] reported that the isolate of B. bassiana with the higher activity of proteases and chitinases led to the lower LC50 and LT50 values against Pseudococcus viburni Signoret (Hemiptera: Pseudococcidae). In our study, no clear correlation was obtained between lipase production and virulence of isolates, similar to earlier studies [12,23,63,64]. This conclusion on lipase may be more obvious in the case of BBLN2, which is the only isolate with higher virulence and lipase activity. Other isolates with higher virulence showed lower lipase activity. It seems that lipases are more important in the utilization of integument lipids for fungal development, not necessarily penetration. In contrast, the isolates with the higher virulence demonstrated the higher activity of proteases and chitinases, mainly BBLN1, BBLN2 and BBRR1. These findings apparently disclosure the correlation between efficiency of extracellular enzymes and higher virulence of the entomopathogenic fungi. Such isolates properly or rapidly penetrate through host cuticle with efficient cleavage of polypeptide and carbohydrate bonds then achieve hemocoel to continue the latter steps of infection. It should be mentioned that this process is accompanied by better production of blastospores and secondary metabolites within the host hemocoel to impose virulence on infected individuals.

3.5. Effects of Thermotolerance and Cold Activity on Conidial Germination

The inactivation and delay of conidia germination caused by heat and cold as the most important environmental factors reduce the efficiency of the entomopathogenic fungi as the biocontrol agents from both the virulence against host and persistence in ecosystems. Selection of the entomopathogenic fungi that tolerate thermal fluctuations te is necessary before field application [21,27,65]. The effect of thermotolerance on the germination rate of the conidia of the collected isolates in the current study have shown in Table 5. Fifteen isolates demonstrated a germination rate of more than 50% after 1 h exposure to 45 °C, while only one isolate exhibited a high tolerance after 2 h (Table 5). After 2 h of exposure, the isolate thermotolerance could be divided into three classes: low (below 30%), moderate (between 30% and 60%) and high tolerance (above 60%). Among the isolates, Aspergillus sp. (ASAI) showed high tolerance. Moderate tolerances were observed in four isolates of B. bassiana (BBLN1, BBAL1, BBLN2 and BBLD2), three isolates of M. anisopliae (MASA, MAAI and MAAL) and two isolates of Trichoderma sp. (TSAH and TSRT). The other isolates were low tolerance to the heat of 45 °C. Similar results have been reported by Lee et al. [27] as the rate of the conidial germination in B. bassiana, M. anisopliae and Lecanicillium attenuatum significantly reduced after 2 h at 45 °C. Rivas et al. [66] demonstrated the significant lower conidial germination of Lecanicillium isolates after incubation at 32 °C. The susceptibility of Metarhizium isolates to high temperatures (45 °C) was demonstrated by Rangel et al. [65]. Exposure to 35 °C for 10 min harmed the conidial germination of B. bassiana, but the M. anisopliae isolate germinated readily at this temperature [67]. Generally, the optimal temperature for conidia germination and growth of entomopathogenic fungi is between 23 and 28 °C. The growth was reduced above 30 °C, and it was totally inhibited above 34 °C [20,27,65,66,67]. Our results imply only Aspergillus sp. isolates as the highly thermotolerant isolate has although it had no virulence against the larvae of C. suppressalis. Finally, the cold activity of fungal isolates was examined through the treatment of the conidia at 5 °C for one and two weeks. All the isolates showed high activity (above 80%) at 5 °C for both time intervals except for Aspergillus sp. (ASAI) (Table 5). Lee et al. [27] reported the high cold activity at 7–14 days for almost all collected isolates. Based on earlier reports, most entomopathogenic fungi have high cold activity, although germination and sporulation may be delayed or stopped at a cold temperature [21,27,68,69]. Such a property may be important in the survival of entomopathogenic fungi in cold periods of the year.

4. Conclusions

Despite there are several commercial mycoinsecticide against major insect pests of agricultural products, exploration to native entomopathogenic fungi in each region may contribute to represent isolates or species with environmental and host adaptations. These adaptations may ensure virulence, environmental persistence and the least non-target effects of native isolates when they are used in field scale. The overall results of our study revealed the isolates of B. bassiana, including BBRR1, BBAL1 and BBLN1, were the most virulent and environmental adaptive isolates among the fungi associated with C. suppressalis based on bioassay, biochemical traits and thermal experiments. These isolates should undergo further studies considering field trials on the target pest and some predators and parasitoids of rice fields to better elucidate their role in pathogenesis.

Author Contributions

Methodology, investigation and writing M.S.; conceptualization, methodology and validation A.Z.; resources and data curation, S.A.K.; conceptualization and resources M.F.-D. All authors have read and agreed to the published version of the manuscript.

Funding

This research was supported by a grant from the Iran National Science Foundation (INSF) No. 96005511.

Acknowledgments

The authors would like to thank Hassan Hoda for his assistance in insect rearing and the collection of some specimens.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Lu, Z.; Zhu, P.; Gurr, G.M.; Zheng, X.; Chen, G.; Heong, K.L. Rice pest management by ecological engineering: A pioneering attempt in China. In Rice Planthoppers; Springer: Dordrecht, The Netherlands, 2015; pp. 161–178. [Google Scholar]
  2. Mirhaghparast, S.K.; Zibaee, A.; Hoda, H.; Sendi, J.J. Changes in cellular immune responses of Chilo suppressalis Walker (Lepidoptera: Crambidae) due to pyriproxyfen treatment. J. Plant Prot. Res. 2015, 55, 287–293. [Google Scholar] [CrossRef]
  3. Zibaee, A.; Sendi, J.; Alinia, F.; Ghadamyari, M.; Etebari, K. Diazinon resistance in different selected strains of Chilo suppressalis Walker (Lepidoptera: Pyralidae), rice striped stem borer, in the north of Iran. J. Econ. Entomol. 2009, 102, 1189–1196. [Google Scholar] [CrossRef] [PubMed]
  4. Cheng, X.; Chang, C.; Dai, S.M. Responses of striped stem borer, Chilo suppressalis (Lepidoptera: Pyralidae), from Taiwan to a range of insecticides. Pest Manag. Sci. 2010, 66, 762–766. [Google Scholar] [CrossRef] [PubMed]
  5. Lacey, L.A.; Grzywacz, D.; Shapiro-Ilan, D.I.; Frutos, R.; Brownbridge, M.; Goettel, M.S. Insect pathogens as biological control agents: Back to the future. J. Invertebr. Pathol. 2015, 132, 1–41. [Google Scholar] [CrossRef] [Green Version]
  6. Pedrini, N. Molecular interactions between entomopathogenic fungi (Hypocreales) and their insect host: Perspectives from stressful cuticle and hemolymph battlefields and the potential of dual RNA sequencing for future studies. Fungal Biol. 2018, 122, 538–545. [Google Scholar] [CrossRef]
  7. Uma Maheswara Rao, C.; Uma Devi, K.; Akbar, A.; Khan, P. Effect of combination treatment with entomopathogenic fungi Beauveria bassiana and Nomuraea rileyi (Hypocreales) on Spodoptera litura (Lepidoptera: Noctuidaeae). Biocontrol Sci. Technol. 2006, 16, 221–232. [Google Scholar] [CrossRef]
  8. Anand, R.; Tiwary, B.N. Pathogenicity of entomopathogenic fungi to eggs and larvae of Spodoptera litura, the common cutworm. Biocontrol Sci. Technol. 2009, 19, 919–929. [Google Scholar] [CrossRef]
  9. Godonou, I.; James, B.; Atcha-Ahowé, C.; Vodouhe, S.; Kooyman, C.; Ahanchédé, A.; Korie, S. Potential of Beauveria bassiana and Metarhizium anisopliae isolates from Benin to control Plutella xylostella L. (Lepidoptera: Plutellidae). Crop Prot. 2009, 28, 220–224. [Google Scholar] [CrossRef]
  10. Hussain, A.; Tian, M.Y.; He, Y.R.; Ahmed, S. Entomopathogenic fungi disturbed the larval growth and feeding performance of Ocinara varians (Lepidoptera: Bombycidae) larvae. Insect Sci. 2009, 16, 511–517. [Google Scholar] [CrossRef]
  11. Wraight, S.P.; Ramos, M.E.; Avery, P.B.; Jaronski, S.T.; Vandenberg, J.D. Comparative virulence of Beauveria bassiana isolates against lepidopteran pests of vegetable crops. J. Invertebr. Pathol. 2010, 103, 186–199. [Google Scholar] [CrossRef]
  12. Ramzi, S.; Zibaee, A. Biochemical properties of different entomopathogenic fungi and their virulence against Chilo suppressalis (Lepidoptera: Crambidae) larvae. Biocontrol Sci. Technol. 2014, 24, 597–610. [Google Scholar] [CrossRef]
  13. Baydar, R.; Güven, Ö.; Karaca, I. Occurrence of entomopathogenic fungi in agricultural soils from Isparta province in Turkey and their pathogenicity to Galleria mellonella (L.) (Lepidoptera: Pyralidae) larvae. Egypt. J. Biol. Pest Control. 2016, 26, 323. [Google Scholar]
  14. Duarte, R.T.; Gonçalves, K.C.; Espinosa, D.J.L.; Moreira, L.F.; De Bortoli, S.A.; Humber, R.A.; Polanczyk, R.A. Potential of entomopathogenic fungi as biological control agents of diamondback moth (Lepidoptera: Plutellidae) and compatibility with chemical insecticides. J. Econ. Entomol. 2016, 109, 594–601. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  15. Amatuzzi, R.F.; Cardoso, N.; Poltronieri, A.S.; Poitevin, C.G.; Dalzoto, P.; Zawadeneak, M.A.; Pimentel, I.C. Potential of endophytic fungi as biocontrol agents of Duponchelia fovealis (Zeller) (Lepidoptera: Crambidae). Braz. J. Biol. 2018, 78, 429–435. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  16. Karthi, S.; Vaideki, K.; Shivakumar, M.S.; Ponsankar, A.; Thanigaivel, A.; Chellappandian, M.; Vasantha-Srinivasan, P.; Muthu-Pandian, C.K.; Hunter, W.B.; Senthil-Nathan, S. Effect of Aspergillus flavus on the mortality and activity of antioxidant enzymes of Spodoptera litura Fab. (Lepidoptera: Noctuidae) larvae. Pestic. Biochem. Physiol. 2018, 149, 54–60. [Google Scholar] [CrossRef]
  17. Li, M.; Bai, Q.; Zang, L.; Ruan, C. Pathogenic fungi identified from the striped stem borer, Chilo suppressalis and their pathogenicity. Chin. J. Biol. Control. 2019, 35, 63–69. [Google Scholar]
  18. Fite, T.; Tefera, T.; Negeri, M.; Damte, T.; Sori, W. Evaluation of Beauveria bassiana, Metarhizium anisopliae, and Bacillus thuringiensis for the management of Helicoverpa armigera (Hubner) (Lepidoptera: Noctuidae) under laboratory and field conditions. Biocontrol Sci. Technol. 2020, 30, 278–295. [Google Scholar] [CrossRef]
  19. Mantzoukas, S.; Pettas, I.; Lagogiannis, I. Stored product pests as models for trapping entomopathogenic fungi from olive tree orchards in Western Greece. J. Stored Prod. Res. 2020, 87, 101584. [Google Scholar] [CrossRef]
  20. Ekesi, S.; Maniania, N.K.; Ampong-Nyarko, K. Effect of temperature on germination, radial growth and virulence of Metarhizium anisopliae and Beauveria bassiana on Megalurothrips sjostedti. Biocontrol Sci. Technol. 1999, 9, 177–185. [Google Scholar] [CrossRef]
  21. Fernandes, E.K.K.; Rangel, D.E.N.; Moraes, A.M.L.; Bittencourt, V.R.E.P.; Roberts, D.W. Cold activity of Beauveria and Metarhizium, and thermotolerance of Beauveria. J. Invertebr. Pathol. 2008, 98, 69–78. [Google Scholar] [CrossRef]
  22. Imoulan, A.; Alaoui, A.; El Meziane, A. Natural occurrence of soil-borne entomopathogenic fungi in the Moroccan Endemic forest of Argania spinosa and their pathogenicity to Ceratitis capitata. World J. Microbiol. Biotechnol. 2011, 27, 2619–2628. [Google Scholar] [CrossRef]
  23. Lu, L.; Cheng, B.; Du, D.; Hu, X.; Peng, A.; Pu, Z.; Zhang, X.; Huang, Z.; Chen, G. Morphological, molecular and virulence characterization of three Lencanicillium species infecting Asian citrus psyllids in Huangyan citrus groves. J. Invertebr. Pathol. 2015, 125, 45–55. [Google Scholar] [CrossRef] [PubMed]
  24. Sevim, A.; Demir, I.; Demirbağ, Z. Molecular characterization and virulence of Beauveria spp. from the pine processionary moth, Thaumetopoea pityocampa (Lepidoptera: Thaumetopoeidae). Mycopathologia 2010, 170, 269–277. [Google Scholar] [CrossRef] [PubMed]
  25. Abdo, C.; Nemer, N.; Nemer, G.; Abou Jawdah, Y.; Atamian, H.; Kawar, N.S. Isolation of Beauveria species from Lebanon and evaluation of its efficacy against the cedar web-spinning sawfly, Cephalcia tannourinensis. BioControl 2008, 53, 341–352. [Google Scholar] [CrossRef]
  26. Goettel, M.S.; Eilenberg, J.; Glare, T. Entomopathogenic Fungi and their Role in Regulation of Insect Populations. In Insect Control; Academic Press: Beijing, China, 2010; pp. 387–431. [Google Scholar]
  27. Lee, W.W.; Shin, T.Y.; Bae, S.M.; Woo, S.D. Screening and evaluation of entomopathogenic fungi against the green peach aphid, Myzus persicae, using multiple tools. J. Asia Pac. Entomol. 2015, 18, 607–615. [Google Scholar] [CrossRef]
  28. Kepler, R.M.; Luangsa-Ard, J.J.; Hywel-Jones, N.L.; Quandt, C.A.; Sung, G.H.; Rehner, S.A.; Aime, M.C.; Henkel, T.W.; Sanjuan, T.; Zare, R.; et al. A phylogenetically-based nomenclature for Cordycipitaceae (Hypocreales). IMA Fungus 2017, 8, 335–353. [Google Scholar] [CrossRef]
  29. Nussenbaum, A.L.; Lecuona, R.E. Selection of Beauveria bassiana sensu lato and Metarhizium anisopliae sensu lato isolates as microbial control agents against the boll weevil (Anthonomus grandis) in Argentina. J. Invertebr. Pathol. 2012, 110, 1–7. [Google Scholar] [CrossRef]
  30. Amatuzzi, R.F.; Poitevin, C.G.; Poltronieri, A.S.; Zawadneak, M.A.; Pimentel, I.C. Susceptibility of Duponchelia fovealis Zeller (Lepidoptera: Crambidae) to Soil-Borne Entomopathogenic Fungi. Insects 2018, 9, 70. [Google Scholar] [CrossRef] [Green Version]
  31. Zahiri, R.; Sarafrazi, A.; Salehi, L.; Kunkel, J.G. A geometric morphometric study on populations of the Rice Stem Borer, Chilo suppressalis Walker (Lepidoptera: Crambidae) in northern Iran. Zool. Middle East 2006, 38, 73–84. [Google Scholar] [CrossRef]
  32. Fang, Z.D. Research Method of Plant Disease, 3rd ed.; China Agriculture Press: Beijing, China, 1998. [Google Scholar]
  33. Zare, R.; Gams, W.J.N.H. A revision of Verticillium section Prostrata. IV. The genera Lecanicillium and Simplicillium gen. nov. N. Hedwig. 2001, 73, 1–50. [Google Scholar]
  34. Humber, R.A. Entomophthoromycota: A new phylum and reclassification for entomophthoroid fungi. Mycotaxon 2012, 120, 477–492. [Google Scholar] [CrossRef]
  35. Rehner, S.A.; Buckley, E. A Beauveria phylogeny inferred from nuclear ITS and EF1-α sequences: Evidence for cryptic diversification and links to Cordyceps teleomorphs. Mycologia 2005, 97, 84–98. [Google Scholar] [CrossRef] [PubMed]
  36. Mains, E.B. Entomogenous species of Hirsutella, Tilachlidium and Synnematium. Mycologia. 1951, 43, 691–718. [Google Scholar] [CrossRef]
  37. Yoon, C.S.; Kim, J.J.; Lee, M.H.; Yun, T.Y.; Yoo, J.K. First Report on Hirsutella subulata, a Pathogen of Rice Stem Borer, Chilo suppressalis in Korea. Korean J. Mycol. 1999, 27, 206–207. [Google Scholar]
  38. Pérez-González, O.; Rodríguez-Villarreal, R.A.; López-Arroyo, J.I.; Maldonado-Blanco, M.G.; Rodríguez-Guerra, R. Mexican strains of Hirsutella isolated from Diaphorina citri (Hemiptera: Liviidae): Morphologic and molecular characterization. Fla. Entomol. 2015, 290–297. [Google Scholar] [CrossRef]
  39. Bischoff, J.F.; Rehner, S.A.; Humber, R.A. A multilocus phylogeny of the Metarhizium anisopliae lineage. Mycologia 2009, 101, 512–530. [Google Scholar] [CrossRef] [Green Version]
  40. Montero-Pau, J.; Gómez, A.; Muñoz, J. Application of an inexpensive and high through put genomic DNA extraction method forthe molecular ecology of zooplanktonic diapausing eggs. Limnol. Oceanogra. Meth. 2008, 6, 218–222. [Google Scholar] [CrossRef] [Green Version]
  41. White, T.J.; Bruns, T.; Lee, S.; Taylor, J.W. Amplification and direct sequencing of fungal ribosomal RNA genes for phylogenetics. In PCR Protocols: A Guide to Methods and Applications; Innis, M.A., Gelfand, D.H., Sninsky, J.J., White, T.J., Eds.; Academic Press Inc.: New York, NY, USA, 1990; pp. 315–322. [Google Scholar]
  42. Shamakhi, L.; Zibaee, A.; Karimi-Malati, A.; Hoda, H. Effect of thermal stress on the immune responses of Chilo suppressalis walker (Lepidoptera: Crambidae) to Beauveria bassiana. J. Therm. Boil. 2019, 84, 136–145. [Google Scholar] [CrossRef]
  43. Ying, S.-H.; Feng, M.-G. Relationship between thermotolerance and hydrophobin like proteins in aerial conidia of Beauveria bassiana and Paecilomyces fumosoroseus and fungal biocontrol agents. J. Appl. Microbiol. 2004, 97, 323–331. [Google Scholar] [CrossRef]
  44. Lowry, O.H.; Rosenbrough, N.J.; Farr, L.L.; Randall, R.J. Protein measurement with the Folin phenol reagent. J. Biol. Chem. 1951, 193, 265–275. [Google Scholar]
  45. Tsujita, T.; Ninomiya, H.; Okuda, H. p-nitrophenyl butyrate hydrolyzing activity of hormone sensitive lipase from bovine adipose tissue. J. Lipid Res. 1989, 30, 997–1004. [Google Scholar] [PubMed]
  46. Miller, G.L. Use of dinitrosalicylic acid reagent for determination of reducing sugar. Anal. Chem. 1959, 31, 426–428. [Google Scholar] [CrossRef]
  47. Tarasco, E.; De Bievre, C.; Papierok, B.; Poliseno, M.; Triggiani, O. Occurrence of entomopathogenic fungi in southern Italy. Entomology 1997, 31, 157–166. [Google Scholar]
  48. Keller, S.; Kessler, P.; Schweizer, C. Distribution of insect pathogenic soil fungi in Switzerland with special reference to Beauveria brongniartii and Metarhizium anisopliae. Biocontrol 2003, 48, 307–319. [Google Scholar] [CrossRef]
  49. Quesada-Moraga, E.; Navas-Corte’s, J.A.; Maranhao, E.A.A.; Ortiz-Urquiza, A.; Santiago-Alvarez, C. Factors affecting the occurrence and distribution of entomopathogenic fungi in natural and cultivated soils. Mycol. Res. 2007, 111, 947–966. [Google Scholar] [CrossRef]
  50. Sun, B.D.; Liu, X.Z. Occurrence and diversity of insect-associated fungi in natural soils in China. Appl. Soil Ecol. 2008, 39, 100–108. [Google Scholar] [CrossRef]
  51. Feng, M.G.; Poprawski, T.J.; Khachatourians, G.G. Production, formulation and application of the entomopathogenic fungus Beauveria bassiana for insect control: Current status. Biocontrol Sci. Technol. 1994, 4, 3–34. [Google Scholar] [CrossRef]
  52. Thomas, K.C.; Khachatourians, G.G.; Ingledew, W.M. Production and properties of Beauveria bassiana conidia cultivated in submerged culture. Can. J. Microbiol. 1987, 33, 12–20. [Google Scholar] [CrossRef]
  53. Jandricic, S.E.; Filotas, M.; Sanderson, J.P.; Wraight, S.P. Pathogenicity of conidia-based preparations of entomopathogenic fungi against the greenhouse pest aphids Myzus persicae, Aphis gossypii, and Aulacorthum solani (Hemiptera: Aphididae). J. Invertebr. Pathol. 2014, 118, 34–46. [Google Scholar] [CrossRef]
  54. Kulkarni, R.; Kadam, J.R.; Mote, U.N. Efficiency of Verticilium lecanii against mealy bugs on pomegranate. J. App. Zool. Res. 2003, 14, 59–60. [Google Scholar]
  55. Zhang, S.; Xia, Y.X.; Kim, B.; Keyhani, N.O. 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, 811–826. [Google Scholar] [CrossRef] [PubMed]
  56. Jiang, Z.Y.; Ligoxygakis, P.; Xia, Y.X. HYD3, a conidial hydrophobin of the fungal entomopathogen Metarhizium acridum induces the immunity of its specialist host locust. Int. J. Biol. Macromol. 2020. [Google Scholar] [CrossRef] [PubMed]
  57. Kershaw, M.J.; Wakley, G.; Talbot, N.J. Complementation of the Mpg1 mutant phenotype in Magnaporthe grisea reveals functional relationships between fungal hydrophobins. EMBO J. 1998, 17, 3838–3849. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  58. Sevim, A.; Donzelli, B.G.; Wu, D.; Demirbaq, Z.; Gibson, D.M.; Turqueon, B.G. Hydrophobin genes of the entomopathogenic fungus, Metarhizium brunneum, are differentially expressed and corresponding mutants are decreased in virulence. Curr. Genet. 2012, 58, 79–92. [Google Scholar] [CrossRef] [PubMed]
  59. Pedrini, N.; Ortiz-Urquiza, A.; Zhang, S.; Keyhani, N.O. Targeting of insect epicuticular lipids by the entomopathogenic fungus Beauveria bassiana: Hydrocarbon oxidation within the context of a host-pathogen interaction. Front. Microbiol. 2013, 4, 24. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  60. Zibaee, A.; Sadeghi-Sefidmazgi, A.; Fazeli-Dinan, M. Properties of a lipase produced by Beauveria bassiana: Purification and biochemical studies. Biocontrol Sci. Technol. 2011, 21, 317–331. [Google Scholar] [CrossRef]
  61. St Leger, R.J.; Charnley, A.K.; Cooper, R.M. Kinetics of the digestion of insect cuticles by a protease (Pr1) from Metarhizium anisopliae. J. Invertebr. Pathol. 1991, 57, 146–147. [Google Scholar] [CrossRef]
  62. Maqsoudi, P.; Ramzi, S.; Zibaee, A.; Khodaparast, S.A. Virulence comparison of two Iranian isolates of Beauveria bassiana Vuillemin against Pseudococcus viburni Signoret (Hemiptera: Pseudococcidae). Trends Entomol. 2019, 14, 63–70. [Google Scholar]
  63. Sosa-Gómez, D.R.; Batista, S.; Tigano, M. Characterization and phenetic analysis of geographical isolates of Beauveria spp. Pesq. Agropec. Bras. 1994, 29, 401–409. [Google Scholar]
  64. Varela, A.; Morales, E. Characterization of some Beauveria bassiana isolates and their virulence toward the coffee berry borer Hypothenemus hampei. J. Invertebr. Pathol. 1996, 67, 147–152. [Google Scholar] [CrossRef]
  65. Rangel, D.E.N.; Butler, M.J.; Torabinejad, J.; Anderson, A.J.; Braga, G.U.L.; Day, A.W.; Roberts, D.W. Mutants and isolates of Metarhizium anisopliae are diverse in their relationships between conidial pigmentation and stress tolerance. J. Invertebr. Pathol. 2006, 93, 170–182. [Google Scholar] [CrossRef] [PubMed]
  66. Rivas, F.; Nuñez, P.; Jackson, T.; Altier, N. Effect of temperature and water activity on mycelia radial growth, conidial production and germination of Lecanicillium spp. isolates and their virulence against Trialeurodes vaporariorum on tomato plants. BioControl 2014, 59, 99–109. [Google Scholar] [CrossRef]
  67. Liu, H.; Skinner, M.; Brownbridge, M.; Parker, B.L. Characterization of Beauveria bassiana and Metarhizium anisopliae isolates for management of tarnished plant bug, Lygus lineolaris (Hemiptera: Miridae). J. Invertebr. Pathol. 2003, 82, 139–147. [Google Scholar] [CrossRef]
  68. Santos, M.P.; Dias, L.P.; Ferreira, P.C.; Pasin, L.A.; Rangel, D.E. Cold activity and tolerance of the entomopathogenic fungus Tolypocladium spp. to UV-B irradiation and heat. J. Invertebr. Pathol. 2011, 108, 209–213. [Google Scholar] [CrossRef] [PubMed]
  69. Shin, T.Y.; Bae, S.M.; Kim, D.J.; Yun, H.G.; Woo, S.D. Evaluation of virulence, tolerance to environmental factors and antimicrobial activities of entomopathogenic fungi against two-spotted spider mite, Tetranychus urticae. Mycoscience 2017, 58, 204–212. [Google Scholar] [CrossRef]
Figure 1. Morphological characteristics (conidium and phialid) of the fungus associated with the larvae of Chilo suppressalis. (a) Akanthomyces lecanii, (b) Akanthomyces muscarius, (c) Aspergillus sp., (d) Beauveria bassiana, (e) Hirutella subulate, (f) Metarhizium anisopliae complex, (g) Trichoderma sp. Bars are 20 micrometers.
Figure 1. Morphological characteristics (conidium and phialid) of the fungus associated with the larvae of Chilo suppressalis. (a) Akanthomyces lecanii, (b) Akanthomyces muscarius, (c) Aspergillus sp., (d) Beauveria bassiana, (e) Hirutella subulate, (f) Metarhizium anisopliae complex, (g) Trichoderma sp. Bars are 20 micrometers.
Jof 07 00034 g001
Figure 2. Activities of the proteases (U/mg protein, Mean ± SE) in the liquid culture media of the entomopathogenic fungi in the presence of C. suppressalis cuticle. Statistical differences are shown by different letters (Tukey’s test, p ≤ 0.05).
Figure 2. Activities of the proteases (U/mg protein, Mean ± SE) in the liquid culture media of the entomopathogenic fungi in the presence of C. suppressalis cuticle. Statistical differences are shown by different letters (Tukey’s test, p ≤ 0.05).
Jof 07 00034 g002aJof 07 00034 g002b
Figure 3. Activity of the lipase (U/mg protein, Mean ± SE) in the liquid culture media of the entomopathogenic fungi in the presence of C. suppressalis cuticle. Statistical differences are shown by different letters (Tukey’s test, p ≤ 0.05).
Figure 3. Activity of the lipase (U/mg protein, Mean ± SE) in the liquid culture media of the entomopathogenic fungi in the presence of C. suppressalis cuticle. Statistical differences are shown by different letters (Tukey’s test, p ≤ 0.05).
Jof 07 00034 g003
Figure 4. Activities of the chitinases (U/mg protein, Mean ± SE) in the liquid culture media of the entomopathogenic fungi in the presence of C. suppressalis cuticle. Statistical differences are shown by different letters (Tukey’s test, p ≤ 0.05).
Figure 4. Activities of the chitinases (U/mg protein, Mean ± SE) in the liquid culture media of the entomopathogenic fungi in the presence of C. suppressalis cuticle. Statistical differences are shown by different letters (Tukey’s test, p ≤ 0.05).
Jof 07 00034 g004
Table 1. Morphological characteristics and GenBank accession number of the collected fungi from the larvae of Chilo suppressalis.
Table 1. Morphological characteristics and GenBank accession number of the collected fungi from the larvae of Chilo suppressalis.
IdentificationIsolatesConidia Size (um)Shape of ConidiaColor of Conidia or ColonyGene Bank Accession No.Ident (%)
Akanthomyces lecaniiALRR4.3 ± 0.07 × 1.9 ± 0.04Ellipsoidal-CylindricalWhiteMW14352799.44
ALFN4.9 ± 0.1 × 1.7 ± 0.04Ellipsoidal-CylindricalWhiteMW14352899.08
ALAL4.5 ± 0.06 × 1.8 ± 0.04Ellipsoidal-CylindricalWhiteMW14352999.08
ALRT4.8 ± 0.06 × 1.6 ± 0.05Ellipsoidal-CylindricalWhiteMW14353199.44
ALLN4.5 ± 0.07 × 1.5 ± 0.03Ellipsoidal-CylindricalWhiteMW14353099.44
Akanthomyces muscariusAMRT8 ± 0.09 × 1.7 ± 0.04CylindricalWhiteMW14352399.82
AMAI7.8 ± 0.08 × 1.8 ± 0.03CylindricalWhiteMW14352498.40
AMAL8.2 ± 0.04 × 1.8 ± 0.05CylindricalWhiteMW14352599.80
AMBL7.5 ± 0.02 × 1.9 ± 0.06CylindricalWhiteMW14352699.61
Aspergillus sp.ASAI2 ± 0.05 × 2 ± 0.04GloboseDark greenMW14353299.82
Beauveria bassianaBBAL12.7 ± 0.07 × 2.5 ± 0.07GloboseWhiteMW14353799.81
BBAL22.8 ± 0.07 × 2.5 ± 0.05GloboseWhiteMW143538100
BBAL33 ± 0.05 × 2.8 ± 0.09GloboseWhiteMW14353999.81
BBAL42.9 ± 0.04 × 2.8 ± 0.08GloboseWhiteMW143540100
BBBL12.9 ± 0.06 × 2.6 ± 0.06GloboseWhiteMW14354199.61
BBBL23.3 ± 0.05 × 2.9 ± 0.09GloboseWhiteMW14354299.61
BBLN12.8 ± 0.08 × 2.6 ± 0.08GloboseWhiteMW14354699.63
BBLN22.9 ± 0.05 × 2.8 ± 0.07GloboseWhiteMW14354799.81
BBLN32.5 ± 0.06 × 2.4 ± 0.04GloboseWhiteMW14354899.63
BBLD12.8 ± 0.05 × 2.6 ± 0.05GloboseWhiteMW14354999.43
BBLD22.9 ± 0.04 × 2.7 ± 0.07GloboseWhiteMW14355099.44
BBLD32.6 ± 0.08 × 2.4 ± 0.05GloboseWhiteMW14355199.26
BBLD42.9 ± 0.04 × 2.5 ± 0.06GloboseWhiteMW14355299.26
BBLD52.8 ± 0.05 × 2.6 ± 0.09GloboseWhiteMW14355399.81
BBRT12.9 ± 0.05 × 2.7 ± 0.09GloboseWhiteMW143533100
BBRT22.8 ± 0.04 × 2.7 ± 0.04GloboseWhiteMW143534100
BBRR12.7 ± 0.08 × 2.4 ± 0.06GloboseWhiteMW143535100
BBRR22.9 ± 0.08 × 2.7 ± 0.07GloboseWhiteMW143536100
BBSI2.9 ± 0.06 × 2.7 ± 0.07GloboseWhiteMW14354498.90
BBFN2.9 ± 0.04 × 2.7 ± 0.08GloboseWhiteMW143543100
BBAI3 ± 0.06 × 2.9 ± 0.08GloboseWhiteMW143545100
Hirutella subulataHSAL5.9 ± 0.1 × 4.5 ± 0.06OvoidWhite-creamMW14355999.61
HSBL6 ± 0.2 × 4.7 ± 0.08OvoidWhite-creamMW14356099.61
Metarhizium anisopliae complexMASA7.6 ± 0.1 × 3.2 ± 0.07Oblong ovalBrown-greenMW143556100
MAAI7.4 ± 0.8 × 3.3 ± 0.05Oblong ovalBrown-greenMW14355799.81
MAAL7.7 ± 0.4 × 3.6 ± 0.05Oblong ovalBrown-greenMW14355899.81
Trichoderma spTSRT2.7 ± 0.08 × 2.7 ± 0.09GloboseDark greenMW143555100
TSAH2.5 ± 0.05 × 2.3 ± 0.06GloboseDark greenMW143554100
Note: final tested isolates were renamed after the submission of the ITS sequence to the GenBank database.
Table 2. LC50 values (conidia/mL) of the entomopathogenic fungi collected from rice fields against the fourth instar larvae of Chilo suppressalis.
Table 2. LC50 values (conidia/mL) of the entomopathogenic fungi collected from rice fields against the fourth instar larvae of Chilo suppressalis.
IsolatesNLC50 (Cl 95%) Conidia/mLX2 (df)Slope ± SE
BBAL11502.1 × 104 (1.1 × 103 − 1.9 × 105)3.253 (3)0.420 ± 0.068
BBAL21502.3 × 105 (4.7 × 104 − 1.4 × 106)0.377 (3)0.313 ± 0.059
BBAL31505.6 × 104 (1.1 × 104 − 2.4 × 105)0.689 (3)0.353 ± 0.062
BBAL41509.6 × 104 (1.9 × 104 − 4.8 × 105)0.820 (3)0.331 ± 0.060
BBBL11501.5 × 105 (3.1 × 104 − 7.9 × 105)0.453 (3)0.321 ± 0.059
BBBL21503.9 × 105 (7.2 × 104 − 3.2 × 106)0.327 (3)0.292 ± 0.058
BBLN11501 × 104 (2.9 × 103 − 9.9 × 104)3.084 (3)0.380 ± 0.065
BBNL21505.4 × 104 (1.1 × 104 − 2.4 × 105)2.285 (3)0.345 ± 0.061
BBLN31501.5 × 105 (4 × 104 − 6.7 × 105)2.455 (3)0.385 ± 0.064
BBLD11501.1 × 105 (2.3 × 104 − 5.5 × 105)1.280 (3)0.336 ± 0.060
BBLD21501 × 105 (2.6 × 104 − 4.5 × 105)1.126 (3)0.379 ± 0.063
BBLD31509.5 × 104 (2.3 × 104 − 3.9 × 105)1.156 (3)0.386 ± 0.064
BBLD41501.2 × 105 (2.7 × 104 − 5.8 × 105)2.365 (3)0.351 ± 0.061
BBLD51504.4 × 105 (9.9 × 104 − 2.7 × 106)1.126 (3)0.336 ± 0.061
BBRT11504.9 × 105 (6.8 × 104 − 6.8 × 106)1.320 (3)0.247 ± 0.056
BBRT21503.4 × 105 (7.8 × 104 − 1.9 × 106)0.534 (3)0.342 ± 0.061
BBRR11502.2 × 104 (4.6 × 103 − 8.8 × 104)1.477 (3)0.337 ± 0.064
BBRR21502.4 × 105 (6.1 × 104 − 1.1 × 106)0.776 (3)0.375 ± 0.063
BBSI1501.4 × 105 (3.2 × 104 − 7.6 × 105)0.513 (3)0.337 ± 0.060
BBFN1502.3 × 105 (4.5 × 104 − 1.5 × 106)1.470 (3)0.307 ± 0.059
BBAI1501.9 × 105 (4.2 × 104– 1 × 106)0.539 (3)0.332 ± 0.060
HSAL1507.9 × 105 (1.5×105 − 7.3 × 106)1.985 (3)0.309 ± 0.052
HSBL1501.6 × 106 (2.4 × 105 − 1.5 × 107)0.449 (3)0.297 ± 0.059
MASA1507.1 × 104 (1.6 × 104 − 2.9 × 105)2.325 (3)0.374 ± 0.062
MAAI1501.6 × 105 (3.4 × 104 − 9.4 × 105)1.743 (3)0.325 ± 0.060
MAAL1503.6 × 105 (7.4 × 104 − 2.4 × 106)0.820 (3)0.315 ± 0.059
Note: calculations were carried out by POLO-Plus software.
Table 3. LT50 values (days) of the entomopathogenic fungi collected from rice fields against the fourth instar larvae of Chilo suppressalis.
Table 3. LT50 values (days) of the entomopathogenic fungi collected from rice fields against the fourth instar larvae of Chilo suppressalis.
IsolatesLT50 (Cl 95%) DaysX2 (df)Slope ± SE
BBAL13.45 (2.55–4.43)11.948 (5)4.007 ± 0.496
BBAL24.16 (3.16–5.21)20.914 (8)3.270 ± 0.347
BBAL33.87 (3.28–4.45)9.303 (7)4.354 ± 0.465
BBAL44.03 (3.47–4.55)7.305 (8)4.12 ± 0.447
BBBL14.18 (3.50–4.86)12.719 (8)3.983 ± 0.410
BBBL24.63 (4.13–5.13)0.883 (8)3.803 ± 0.419
BBLN13.15 (2.36–4)13.895 (5)3.302 ± 0.423
BBNL23.70 (2.96–4.45)12.155 (7)3.545 ± 0.392
BBLN33.75 (2.96–4.54)14.866 (7)3.899 ± 0.416
BBLD14.05 (3.25–4.84)15.98 (8)3.654 ± 0.376
BBLD24.28 (3.30–5.23)27.697 (8)3.654 ± 0.376
BBLD33.81 (3.38–4.21)6.745 (7)4.248 ± 0.456
BBLD43.66 (3.24–4.05)5.112 (8)4.166 ± 0.420
BBLD54.23 (3.74–2.4.71)3.729 (8)3.648 ± 0.386
BBRT14.91 (4.03–5.78)15.756 (8)3.623 ± 0.396
BBRT24.77 (4.29–5.24)3.858 (8)4.141 ± 0.445
BBRR12.71 (2.29–3.10)5.961 (6)3.260 ± 0.383
BBRR24.16 (3.52–4.78)12.398 (8)4.373 ± 0.441
BBSI4.33 (3.77–4.87)8.335 (8)4.084 ± 0.419
BBFN4.48 (3.88–5.08)9.194 (8)3.99 ± 0.414
BBAI4.41 (3.67–5.14)14.25 (8)4.085 ± 0.419
HSAL4.65 (3.83–5.54)13.498 (7)4.070 ± 0.457
HSBL5.21 (4.72–5.74)2.075 (7)4.434 ± 0.526
MASA3.69 (3.01–4.38)9.280 (6)4.021 ± 0.470
MAAI4.91 (4.03–5.87)17.220 (7)3.723 ± 0.410
MAAL4.14 (3.64–4.64)1.887 (8)3.386 ± 0.366
Note: calculations were carried out by POLO-Plus software.
Table 4. Amount of the hydrophobin (mg/mL) in the collected entomopathogenic fungi from the larvae of Chilo suppressalis.
Table 4. Amount of the hydrophobin (mg/mL) in the collected entomopathogenic fungi from the larvae of Chilo suppressalis.
IsolatesAmount of Hydrophobin (mg/mL)
ALRR0.0603 ± 0.007 fghi
ALFN0.0628 ± 006 efghi
ALAL0.0535 ± 0.005 hi
ALRT0.0586 ± 0.002 ghi
ALLN0.0627 ± 0.003 efghi
AMRT0.0687 ± 0.003 cdefgh
AMAI0.0663 ± 0.005 cdefgh
AMAL0.0656 ± 0.001 defgh
AMBL0.0679 ± 0.002 cdefgh
ASAI0.038 ± 0.005 j
BBAL10.0953 ± 0.001 a
BBAL20.0745 ± 0.004 bcdefgh
BBAL30.0765 ± 0.002 abcdefg
BBAL40.0749 ± 0.002 bcdefgh
BBBL10.0846 ± 0.002 abcd
BBBL20.0780 ± 0.001 abcdefgh
BBLN10.822 ± 0.003 abcde
BBLN20.0803 ± 0.003 abcde
BBLN30.0756 ± 0.003 abcdefgh
BBLD10.0854 ± 0.004 abc
BBLD20.0796 ± 0.006 abcdefg
BBLD30.0704 ± 0.003 bcdefgh
BBLD40.0782 ± 0.002 abcdefg
BBLD50.0897 ± 0.002 ab
BBRT10.0704 ± 0.004 bcdefgh
BBRT20.0774 ± 0.002 abcdefg
BBRR10.0767 ± 0.002 abcdefg
BBRR20.0677 ± 0.003 cdefgh
BBSI0.0762 ± 0.001 abcdefg
BBFN0.0729 ± 0.001 bcdefgh
BBAI0.0832 ± 0.002 abcd
HSAL0.0816 ± 0.001 abcde
HSBL0.0684 ± 0.003 cdefgh
MASA0.0805 ± 0.002 abcde
MAAI0.0631 ± 0.002 ifghi
MAAL0.0714 ± 0.001 bcdefgh
TSRT0.0036 ± 0.002 j
TSAH0.043 ± 0.002 ij
Note: Statistical differences are shown by different letters (Tukey’s test, p ≤ 0.05).
Table 5. Thermotolerance and cold activity of the entomopathogenic fungi collected from the larvae of Chilo suppressalis.
Table 5. Thermotolerance and cold activity of the entomopathogenic fungi collected from the larvae of Chilo suppressalis.
IsolatesConidial Germination (%)
Exposure to 45 °CIncubating at 4 °C
1 h2 h7 Day14 Day
ALRR29.86 ± 1.2 mnopq10.90 ± 0.6 op84.36 ± 1.6 ghijk90.53 ± 1.5 abcdef
ALFN27.78 ± 1 nopq12.21 ± 0.3 mnop94.20 ± 0.9 abcde95.85 ± 0.9 abcd
ALAL37.75 ± 2.1 jklmn16.39 ± 0.5 ijklmno88.58 ± 1. abcdefghij93.98 ± 0.9 abcde
ALRT29.54 ± 1.5 mnopq11.98 ± 0.9 nop95.66 ± 1.6 abc96.28 ± 0.8 abc
ALLN35.46 ± 1.3 klmnop12.98 ± 0.9 klmnop86.39 ± 1.7 efghij90.72 ± 0.9 abcdef
AMRT22.67 ± 1.6 pq9.07 ± 0.6 op91.14 ± 1.4 abcdefgh96.49 ± 1.3 abc
AMAI27.13 ± 1.8 pq10.64 ± 0.6 pq94.98 ± 0.7 abcd97.07 ± 0.9 ab
AMAL29.95 ± 1.9 mnopq12.76 ± 0.7 klmnop87.13 ± 1.6 defghij94.30 ± 0.9 abcde
AMBL21.91 ± 1.7 pq8.93 ± 0.5 pq95.95 ± 1.3 a98.08 ± 0.6 a
ASAI97.33 ± 0.8 a92.41 ± 0.8 a0 l0 g
BBAL175.25 ± 1.6 bc38.55 ± 1.6 bcd93.19 ± 0.7 abcdef97.52 ± 0.7 ab
BBAL257.24 ± 1.4 fg23.55 ± 1.1 gh92.14 ± 0.8 abcdefg94.42 ± 0.5 abcde
BBAL338.14 ± 1.6 klm14.43 ± 0.9 klmnop88.45 ± 1.6 abcdefghij93.40 ± 0.9 abcde
BBAL436.36 ± 1.9 jklmno12.60 ± 1.0 lmnop87.80 ± 1.0 cdefghij91.94 ± 0.7 abcdef
BBBL150.41 ± 1.3 ghi18.93 ± 1.1 hijkl89.30 ± 1.1 abcdefghij94.65 ± 0.7 abcd
BBBL251.03 ± 1.8 ghi20.95 ± 1.0 ghij87.96 ± 1.4 bcdefghij91.07 ± 1 abcdef
BBLN176.29 ± 1.9 b41.58 ± 1.3 bc95.84 ± 0.7 ab97.29 ± 0.9 ab
BBLN269.56 ± 1.2 bcd35.83 ± 1.8 cde86.54 ± 0.9 efghij92.33 ± 0.8 abcde
BBLN342.47 ± 1.7 ijkl21.85 ± 1.2 ghi88.24 ± 1.6 abcdefghij93.19 ± 0.9 abcde
BBLD143.06 ± 1.5 hijk26.70 ± 1.1 fg93.78 ± 0.9 abcdef91.51 ± 6.1 abcde
BBLD257.79 ± 1.3 efg30.97 ± 1.1 ef86.48 ± 1.1 efghij91.68 ± 0.8 abcde
BBLD326.51 ± 1.4 pq11.27 ± 0.7 op94.98 ± 0.9 abcd96.86 ± 0.7 ab
BBLD437.00 ± 1.1 jklmno14.55 ± 1.1jkl mnop85.03 ± 1.1 fghijk90.64 ± 0.9 abcdef
BBLD533.88 ± 2.3 lmnop12.19 ± 1.1 mnop85.95 ± 1.2 fghijk91.11 ± 1 abcdef
BBRT151.95 ± 1.4 gh19.08 ± 1.1 hijk93.83 ± 0.9 abcdef96.50 ± 0.7 abc
BBRT235.18 ± 1.7 klmnop18.59 ± 0.8 hijklm90.53 ± 0.8 abcdefghi95.47 ± 0.8 abcd
BBRR172.63 ± 1.2 bc33.95 ± 1.6 de94.23 ± 0.9 abcde98.14 ± 0.5 a
BBRR228.27 ± 1.2 opq12.29 ± 1.6 mnop84.83 ± 1.2 fghijk93.03 ± 0.9 abcde
BBSI42.16 ± 1.7 ijkl18.04 ± 1.1 hijklmn82.78 ± 1.3 ijk90.66 ± 1.2 abcdef
BBFN44.60 ± 2.1 hij19.08 ± 1.1 hijk87.55 ± 1.2 defghij91.90 ± 1.2 abcde
BBAI42.76 ± 2.3 ijkl18.59 ± 0.8 hijklm87.60 ± 1.4 defghij93.38 ± 1.1 abcde
HSAL11.38 ± 0.9 r2.90 ± 0.7 q83.48 ± 1.6 hijk88.83 ± 1.5 cdef
HSBL9.77 ± 0.6 r1.59 ± 0.5 q82.95 ± 3.3 ijk86.87 ± 1.4 ef
MASA73.41 ± 1.2 bc36.91 ± 2.1 bcde89.87 ± 2.1 abcdefgij94.93 ± 1.7 abcd
MAAI67.65 ± 1.6 bcd33.82 ± 1.5 de83.93 ± 1.4 hijk89.85 ± 1.3 bcdef
MAAL63.04 ± 1.4 def31.31 ± 1.2 ef82.04 ± 1.5 jk88.30 ± 0.7 def
TSRT69.8 ± 1.7 bcd40.57 ± 1.7 bc88.11 ± 1.2 abcdefghij93.64 ± 0.8 abcde
TSAH75.05 ± 1.7 bc43.29 ± 1.2 b84.74 ± 2.3 ghijk89.89 ± 0.8 bcdef
Note: Statistical differences are shown by different letters.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Shahriari, M.; Zibaee, A.; Khodaparast, S.A.; Fazeli-Dinan, M. Screening and Virulence of the Entomopathogenic Fungi Associated with Chilo suppressalis Walker. J. Fungi 2021, 7, 34. https://doi.org/10.3390/jof7010034

AMA Style

Shahriari M, Zibaee A, Khodaparast SA, Fazeli-Dinan M. Screening and Virulence of the Entomopathogenic Fungi Associated with Chilo suppressalis Walker. Journal of Fungi. 2021; 7(1):34. https://doi.org/10.3390/jof7010034

Chicago/Turabian Style

Shahriari, Morteza, Arash Zibaee, Seyyed Akbar Khodaparast, and Mahmoud Fazeli-Dinan. 2021. "Screening and Virulence of the Entomopathogenic Fungi Associated with Chilo suppressalis Walker" Journal of Fungi 7, no. 1: 34. https://doi.org/10.3390/jof7010034

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop