Open Access
How to translate text using browser tools
1 May 2014 Selection of Beauveria Isolates Pathogenic to Adults of Nilaparvata lugens
Maoye Li, Shiguang Li, Amei Xu, Huafeng Lin, Dexin Chen, Hui Wang
Author Affiliations +
Abstract

The brown planthopper, Nilaparvata lugens Stål (Hemiptera: Delphacidae), is a destructive invasive pest and has become one of the most economically-important rice pests in China. Effective control measures are desperately needed. Entomopathogenic fungi, such as Beauveria bassiana (Balsamo-Crivelli) Vuillemin (Hypocreales: Clavicipitaceae) and B. brongniartii (Saccardo), have shown great potential for the management of some sucking pest species. In this study, to explore alternative strategies for sustainable control of the sucking pest population, nine isolates of Beauveria from different pests were bioassayed under the concentrated standard spray of 1000 conidia/mm2 in laboratory. The cumulative mortalities of adults ranged from 17.2 to 79.1% 10 days after inoculation. The virulence among all tested isolates exhibited significant differences (at p = 0.05). The highest virulent isolate was Bb09, which killed 79.1% of the treated insects and had a median lethal time of 5.5 days. Its median lethal concentration values were estimated as 134 conidia/mm2 on day 10. The chitinase activities of nine isolates were also assayed. The results showed that the chitinase activity (18.7 U/mg) of isolate Bbr09 was the highest among all tested isolates. The biological characteristics of these strains, including growth rate, sporulation, and germination rate, were further investigated. The results showed that strain Bbr09 exhibited the best biological characteristics with relatively higher hyphal growth rate, the highest spore production, and the fastest spore germination. The isolate of Bbr09 had strong pathogenicity and exhibited great potential for sustainable control of N. lugens.

Introduction

The brown planthopper, Nilaparvata lugens Stål (Hemiptera: Delphacidae), is one of the major rice pests throughout Asia. At present, chemical insecticides are still the only means utilized for the control of N. lugens. However the consequential problems of chemical insectides are of significant environmental concern. The extensive use of chemical insecticides can cause target insect resistance, detrimental impact on the natural enemies, health and environmental hazards, residue persistence, and development of tolerance (Liu et al. 2003; Jin et al. 2008; Kontsedalov et al. 2012). Therefore, there is a need for alternatives, and biocontrol may be an effective alternative to chemical control while also being friendly to environment (Faria and Wraight 2001).

Entomopathogenic fungi are important worldwide biological control agents, and a number of them are already commercially available (Peveling and Demba 1997; Faria and Wraight 2007; Rodriguez et al. 2009). At present, Beauveria bassiana (Balsamo-Crivelli) Vuillemin (Hypocreales: Clavicipitaceae), Metarhizium anisopliae (Metachnikoff) Sorokin, and Metarhizium flavoviride Gams and Rozsypal, which are studied intensively, have the potential to be used for biological control of sucking pests, such as aphids (Vandenberg et al. 2001; Shan and Feng 2010), whiteflies (Abdel-Raheem et al. 2009; Cuthbertson et al. 2011), leafhoppers (Feng et al. 2004; Pu et al. 2005), and planthoppers (Toledo et al. 2007, 2008; Kiran and Veeranna 2012). It was reported that these entomopathogenic fungi appeared to be the most efficient because of their ease of virulence, massive production, storage, and application. As for fungal insecticide development, virulence of fungal biological control agents is usually used as a primary criterion for isolate selection in addition to other biological characteris-characteristics (Samuels et al. 1989; Reay et al. 2008; Petlamul and Prasertsan 2012). Virulence is the most important indicator when measuring the potential of fungi against a pest and is the basis for choosing highly virulent fungi in laboratory bioassays (Roberts and Leger 2004; Jin et al. 2008). Entomopathogenic fungi cause infection by active penetration through the cuticle of insects and release extracellular cuticle-hydrolyzing enzymes that participate in the decomposition of protein, chitin, and lipids, which are the principal components of the cuticle (St. Leger et al. 1986; Pedrini et al. 2007; Schrank and Vainstein 2010). Different studies suggest that chitinases and proteases are major determinants of fungal virulence in the complex and multifactorial insect pathogen relationship (Fan et al. 2007; Pelizza et al. 2012). Therefore, more and more laboratory bioassays continue to be conducted in order to obtain a fungal isolate with higher efficacy for integrated N. lugens control.

In this study, nine strains of Beauveria collected from different pest species were selected, and their virulence on N. lugens adults and biological characteristics, including colony morphology and growth rate, conidial germination, and chitinase activity, were assessed for the selection as a suitable biological control agent candidate in the management of N. lugens.

Figure 1.

(A) Healthy adult Nilaparvata lugens. (B) Adult N. lugens (96 hr) after death caused by Beauveria (isolate Bbr09). Bar: 500 µm. High quality figures are available online.

f01_01.jpg

Table 1.

List of fungal species screened against adults of Nilaparvata lugens, original host, and geographic origin. Isolate codes follow fungal number of research center of entomopathogenic fungi.

t01_01.gif

Materials and Methods

Plant culture

Oryza sativa L. (Poales: Poaceae) cv. Nongfengyou, a rice cultivar susceptible to N. lugens, was prepared: 10-day-old rice seedlings in dishes (15 cm diameter) were transplanted in plastic pots (20 cm diameter, 20 cm high), which were filled with nutrition solution and cultured for another 10 days under controlled conditions (25–30°C; 14:10 L:D photoperiod; 80% RH) in the greenhouse at the School of Plant Protection, Anhui Agricultural University, Hefei, China.

N. lugens population rearing

Adult N. lugens (Figure 1A), originally collected from the rice fields in Anhui Agricultural University (No. 130 West Changjiang Rd, Hefei, Anhui, China), were reared on the rice seedlings described above. The condition stimulated the copula and oviposition by females. Nymphs that were born from these ovipositions and grew up to be adults were used in pathogenicity assays.

Sources of fungi

The fungal isolates of B. bassiana and B. brongniartii (Saccardo) used in the experiment were provided by Biocontrol Reserch Lab, Anhui Agricultural University, Hefei, China, and originally isolated from a broad spectrum of insects (Table 1). In order to confirm virulence of the isolates was not attenuated, original host insects were inoculated with these isolates and the fungi were reisolated on Sabouraud dextrose agar yeast medium. The conidia colonies were inoculated to Sabouraud dextrose agar yeast medium in Petri dishes (9 cm diameter) and maintained at 25 ± 1°C in darkness for 15 days. After being scraped from the plate, conidia were dried to a water content of 5% at ambient temperature with a vacuum drier (VirTis, SP Scientific,  www.spscientific.com). Dry conidia were preserved at 4°C in darkness for use as soon as possible in all tests, warranting ≥ 90% viability.

Conidia of each fungal isolate were transferred into corresponding test tubes (3 cm diameter, 20 cm length) and suspended in sterile distilled water by shaking in 10 mL flasks containing 60 to 72 glass beads (3 mm diameter). No surfactant [0.02% Tween-80 (polyoxyethylene sorbitan monolaurate)] was added, since 5 min of agitation at 700 oscillations on a mechanical shaker homogenize suspensions of viable single conidia. The suspensions were then adjusted to 1 × 108 conidia/mL by appropriate dilution based on hemocytometer counts.

Bioassays with N. lugens adults

Two different series of bioassays were conducted. For each fungal isolate, 60 adults (approximately 10 days old) from the rice seedlings in the greenhouse were aspirated into test tubes (3 cm diameter, 20 cm length). Afterwards, they were transferred into a transparent bell glass jar (6 cm diameter, 20 cm high) with a bottle cap that had small holes for air circulation to avoid adults escaping. The suspensions of 1 × 108 conidia/mL in sterile water with 0.05% (v/v) Tween-80 that was prepared previously were sprayed onto adults and rice seedlings in the glass jar by a hand-held Micro Ultra sprayer (Micron Sprayers,  www.micron.co.uk). The concentration of conidia deposited onto the adults and seedlings was measured as number of conidia/mm2 using microscopic (400 × magnification) counts of conidia collected onto four glass slips (20 × 20 mm) under each spray. Four replicates per isolate were used, and a blank control was sprayed with 0.05% (v/v) Tween-80 for each fungal isolate. All treatments were maintained at 25 ± 1°C, a 14:10 L:D photoperiod, and 80% RH, with adult mortality recorded daily for 10 days post-treatment. The dead adults were transferred every day into 70% ethanol for 10 sec, washed in sterile distilled water, then treated with 0.5% sodium hypochlorite for 30 sec, and washed again in sterile distilled water. After that, the specimens were placed in Petri dishes (9 cm diameter) with filter paper that was moistened with sterile distilled water and maintained in a humidity chamber (saturated atmosphere) for fungal sporulation at 25 ± 1°C for three to five days.

The second series of bioassays included only those isolates that caused greater than 50% adult mortality in the first series of experiment. Bioassays used aqueous B. brongniartii (Bbr03 and Bbr09) suspensions (1 × 106, 1 × 107 , and 1 × 108 conidia/mL) containing 0.02% Tween-80 in distilled water (the control). The bioassays were repeated four times using the same method described above.

Fungal characteristics, colony growth rates, and conidial yields

The isolates were cultured on Sabouraud dextrose agar yeast medium in Petri dishes (90 mm diameter), with four replicates for each isolate. The dishes were maintained in an incubation chamber at 25 ± 1°C and RH ≥ 80% in the dark. The diameters of the colonies were measured five, 10, and 15 days after incubation.

After 15 days, a culture disc from the center of each fungal colony was removed using a sterile 13 mm diameter punch and placed into a 50 mL flask with 20 mL sterile 0.05% Tween-80. Each isolate contained four flask cultures as replicates. The conidial suspensions of the nine isolates were prepared by stirring as described above. Conidial concentrations were determined using a hemocytometer according to the method by Zhang et al. (2011).

Conidial germination

Fifteen µL of each conidial suspension were transferred to a sterile glass slide smeared with Sabouraud dextrose agar yeast. The slide was incubated in at 25 ± 1°C and RH > 90%. Four replicates were set up for each isolate. Conidial germination was estimated at 24 hr by counting the germinated and non-germinated conidia in a random field of vision sample on each slide using a microscope at 400× magnification. The number of germinated and non-germinated conidia from four microscopic view fields (100 conidia/field) was counted. For each observation, 400 conidia were examined in each replicate slide. The mean percent germination was calculated for each sample.

Chitinase activity assay

The enzymatic reaction was initiated by addition of 200 µL isolated chitinase solution to the colloidal chitin (10% w/v chitin in 300 µL acetate buffer, pH 5.0) as substrate and continued for 60 min at 50°C. The hydrolysis of chitin was measured by the para-dimethyl-amino benzaldehyde reagent method according to Reissig et al. (1955).

The reaction was cooled to room temperature, centrifuged at 5,000 × g for 10 min, and the absorbance at 540 nm (A540) was taken against water as blank. The enzyme activity was measured according to a standard curve developed using known concentrations of N-acetyl-d-glucosamine. One enzyme unit (U) was defined as the amount of enzyme that produced 1 µM of N-acetyl glucosamine per minute under the above conditions.

Table 2.

Bioassay results for adults of Nilaparvata lugens representing isolates of Beauveria.

t02_01.gif

Statistical analyses

The mortalities (%) of N. lugens in treatments (M2) were corrected by the mortality (%) in blank controls (M1) according to Abbortt's formula: Mc = (M2 - Ml) / (100-M1). All corrected mortalities (%) data were subjected to an arc-sine transformation prior to analysis of variance to detected differences in mortalities of N. lugens. Mean separation was performed using Tukey's honestly significant difference test (p <0.05). Statistical analyses were performed using the DPS program (Tang and Feng 2007).

Results

Infection of fungal pathogens to N. lugens adults

It was found that N. lugens adults infected with Beauveria moved sluggishly in the early stage of infection, female adults stopped oviposition, and nymphs had difficulty in peeling. The infected N. lugens adults then died holding rice stems, which were not found in the control group (Figure 1A). As the disease progressed, N. lugens adults' bodies turned brown when infected with Beauveria for three days, and white flocculent hyphae grew out of the insect body from the segmacoria, somatic fold sag, foot joints, etc. In the following one to two days, the mycelium covered the whole body until white or light yellow conidia (Figure 1B) emerged. From inoculation to the completion of the entire process of intrusion infection, tissue lesions, host's death, and in vitro sporulation, the shortest time needed was four days, and the longest nine days.

Table 3.

Virulence of the selected isolates Bbr03 and Bbr09 towards Nilaparvata lugens adults.

t03_01.gif

Table 4.

Characteristics of the tested Beauveria bassiana (Bb) and B. brongniartii (Bbr) isolates on medium.

t04_01.gif

Virulence test on N. lugens adults

The cumulative corrective mortalities of N. lugens differed significantly among the 10 isolate treatments (F8, 27 = 269.5, p < 0.01) on the 10th day after spraying. Only two isolates, Bbr03 and Bbr09, killed more than 50% of the adults, of which Bbr09 had the lowest LT50 of (5.5 days). Cumulative adult mortality increased with conidial concentration. Based on the estimates of the LC50 and associated 95% CI (Table 3), the two isolates, Bbr09 and Bbr03, had LC50S of 134 and 443 conidia/mm2 on day 10, respectively.

Fungal characteristics, colony growth rates, and conidial yields

Viewing from biological characteristics of the original 15 days of culture (Table 4), colony growth rates (5 days: F8, 27 = 60.7, p < 0.01; 10 days: F8, 27 = 177.1, p < 0.01; 15 days: F8, 27 = 247.4, p < 0.01), conidial yields (F8, 27 = 477.7, p < 0.01), and spore germination (F8, 27 = 31.8, p < 0.01) of the nine isolates were tested, and the results differed significantly among them. The isolate Bbr09 generally had the fastest colony growth rates and the highest conidial yields. Spore germination of all tested isolates exceeded 80% within 24 hr.

Figure 2.

Chitinase activity assay of the examined strains. The letters on the top of columns show the significance of difference. High quality figures are available online.

f02_01.jpg

Chitinase activities

The chitinase activities of nine isolates demonstrated a significant variation (Figure 2; p < 0.01). The three strains Bbr09, Bbr03, and Bb255 had high chitinase activities, among which Bbr09 had the highest chitinase activity of (18.7 U/mL), whereas, the isolate Bb81 showed the least chitinase activity (4.3 U/mL).

Discussion

In this research, no relationship between pathogenicity and the isolates' original hosts was observed, even though most authors agree that isolates of entomopathogenic fungi are generally more pathogenic to the species of insect that they are isolated from or a closely related species (Zimmermann 1982; Poprawski et al. 1985). The results of our study do not support this point of view, since the most pathogenic isolate, B. brongniartii Bbr09, was isolated from larvae of Anomala corpulenta Motschulsky (Coleoptera: Rutelidae) living in corn crops. The results from this study suggest that isolates from natural environments have the potential to be highly pathogenic, and the screening of entomopathogenic fungi should not be limited only to isolates from the target host or its close relatives.

As the results of this study showed, isolates with higher chitinase activity had relatively higher virulence. The strain Bbr09, which had the highest chitinase activity among the test isolates, also caused the highest adult mortality. Studies have shown that the chitinase secreted from B. bassiana (Gupta et al. 1992; Havukkala et al. 1993; Pelizza et al. 2012), M. anisopliae (St. Leger et al. 1991), M. flavoviride (St. Leger et al. 1993), and Nomuraea rileyi (El-Sayed et al. 1989) was closely related to strain virulence. Fang et al. (2005) showed that highly expressed chitinase (Bbchit1) of B. bassiana could significantly improve the insecticidal virulence, and this phenomenon has been found in both fungi and bacteria. Research on Beauveria and Metarhizium infecting Peregrinus mai dis by Toledo et al. (2010) showed a similar phenomenon, as during the process of conidia germinating and penetrating epidermis, the secreted chitinase, esterase, and extracellular protease destroyed insect body surface and promoted the successful invasion of hyphae. During the process of entomopathogenic fungi infecting insects, chitinase not only participated in the degradation of the insect body wall by itself, but also played did so together with other enzymes, such as protease.

This research showed that biological characteristics of fungi strains were related to virulence. Zhang et al. (2011) found that the average growth speed, spore production quantity, and germination percentage of the strains were all related to virulence of the strains during the process of screening Beauveria, which had high virulence to Dendroctonns valens. Petlamul and Prasertsan (2012) argued that strains with higher germination rate had stronger virulence to Spodoptera litura. However, Liu et al. (2003) reported that no obvious correlation between germination rate and virulence of the fungal isolates was found in Beauveria and Metarhizium isolates against Lymantria xylina. In this study, colony growth rates, conidial germination rates, and spore production quantity of isolate Bbr09 were all superior to those of the other eight indoor saved strains. A shorter time needed for producing spores and a higher unit area of spore production quantity meant a higher production efficiency during the industrial production of fungal pesticides (Feng et al. 1994; Kassa et al. 2008); meanwhile, in the natural environment, strains would produce more spores for the next infection and diffusion cycle after finishing infecting the host insect.

In this study, bioassay results of the new strain B. brongniartii Bbr09, isolated from the field, showed it had the highest (79.1%) accumulative corrective mortality and the least LT50 (5.5 days) to N. lugens among the test strains. The study on chitinase activity, conidium infection ability, and biological characteristics showed the isolate Bbr09 was superior to those of other strains. The isolate Bbr09 was found to be efficienct in the control of eggs of N. lugens (Li et al. 2012). This strain had high insecticidal efficiency and was easily produced, showing that it has great application potential in field control of N. lugens.

Acknowledgements

We would like to thank Dr. Ian Riley, who is senior biologist in the South Australian Research and Development Institute, for a critical review of the manuscript. This research was supported by the Anhui Agricultural University Science Foundation for Young Scholars (2012zd013), the science and technology key projects of China tobacco corporation (110201202003), the science and technology projects of Sichuang company of China tobacco corporation (201202007), Anhui Province Natural Science Foundation (1408085QC52), and by the National High Technology Research and Development Program of China (863 Program) (2011AA10A205).

References

1.

AM Abdel-Raheem , KH Sabry , ZA. Ragab 2009. Effect of different fertilization rates on control of Bemisia tabaci (Genn.) by Verticillium lecanii and Beauveria bassiana in potato crop. Egyptian Journal of Biological Pest Control 19: 129–133. Google Scholar

2.

AGS Cuthbertson , LF Blackburn , DP Eyre , RJC Cannon , J Miller , P. Northing 2011. Bemisia tabaci: The current situation in the UK and the prospect of developing strategies for eradication using entomopathogens. Insect Science 18: 1–10. Google Scholar

3.

GN El-Sayed , TA Coudron , CM. Ignoffo 1989. Chitinolytic activity and virulence associated with native and mutant isolates of an entomopathogenic fungus Nomuraea rileyi. Journal of Invertebrate Pathology 54: 394– 403. Google Scholar

4.

YH Fan , WG Fang , SJ Gou , XQ Pei , YJ Zhang , YH Xiao , DM Li , K Jin , MJ Bidochka , Y. Pei 2007. Increased Insect Virulence in Beauveria bassiana Strains Over-expressing and Engineered Chitinase. Applied and Environmental Microbiology 73: 295–302. Google Scholar

5.

W Fang , B Leng , Y Xiao , K Jin , JC Ma , YH Fan , J Feng , XY Yang , YJ Zhang , Y. Pei 2005. Cloning of Beauveria bassiana chitinase gene Bbchitl and its application to improve fungal strain virulence. Applied and Environment Microbiology 71: 363–370. Google Scholar

6.

M Faria , SP. Wraight 2001. Biological control of Bemisia tabaci with fungi. Crop Protection 20: 767–778. Google Scholar

7.

M Faria , SP. Wraight 2007. Mycoinsecticides and mycoacaricides: a comprehensive list with worldwide coverage and international classification of formulation types. Biological Control 43: 237–256. Google Scholar

8.

MG Feng , TJ Poprawski , GG. Khachatourians 1994. Production, formulation and application of the entomopathogenic fungus Beauveria bassiana for insect control: current status. Biocontrol Science and Technology 4: 3–34. Google Scholar

9.

MG Feng , XY Pu , SH Ying , YG. Wang 2004. Field trials of an oil-based emulsifiable formulation of Beauveria bassiana conidia and low application rates of imidacloprid for control of false-eye leafhopper Empoasca vitis on tea in southern China. Crop Protection 23 : 489–496. Google Scholar

10.

SC Gupta , TD Leathers , GN El-Sayed , CM. Lgnoffo 1992. Insect cuticle-degrading enzymes from the entomogenous fungus Beauveria bassiana. Experimental Mycology 16: 132–137. Google Scholar

11.

I Havukkala , C Mitamura , S Hara , K Hirayae , Y Nishizawa , T. Hibi 1993. Induction and purification of Beauveria bassiana chitinolytic enzymes. Journal of Invertebrate Pathology 61: 97–102. Google Scholar

12.

SF Jin , MG Feng , JQ. Chen 2008. Selection of global Metarhizium isolates for the control of the rice pest Nilaparvata lugens (Homoptera: Delphacidae). Pest Management Science 64: 1008–1014. Google Scholar

13.

A Kassa , M Brownbridge , BL Parker , M Skinner , V Gouli , S Gouli , M Guo , F Lee , T. Hata 2008. Whey for mass production of Beauveria bassiana and Metarhizium anisopliae. Mycological Research 112: 583–591. Google Scholar

14.

R Kiran , R. Veeranna 2012. Evaluation of bio-pesticide Metarhizium anisoplea against brown plant hopper (Nilaparvatha lugens) and its efficiency on the improvement of the productivity of paddy. International Journal of Plant Protection 5:81–83. Google Scholar

15.

S Kontsedalov , F Abu-Moch , G Lebedev , H Czosnek , AR Horowitz , M. Ghanim 2012. Bemisia tabaci biotype dynamics and resistance to insecticides in Israel during the years 2008–2010. Journal of Integrative Agriculture 11: 312–320. Google Scholar

16.

MY Li , HF Lin , SG Li , AM Xu , MF. Feng 2012. Efficiency of entomopathogenic fungi in the control of eggs of the brown planthopper Nilaparvata lugens Stål. African Journal of Microbiology Research 6: 7162– 7167. Google Scholar

17.

HP Liu , M Skinner , M Brownbridge , BL. Parker 2003. Characterization of Beauveria bassiana and Metarhizium anisopliae isolates for management of tarnished plant bug, Lygus lineolaris (Hemiptera: Miridae). Journal of Invertebrate Pathology 82: 139–147. Google Scholar

18.

ZW Liu , ZJ Han , YC Wang , LC Zhang , HW Zhang , CJ. Liu 2003. Selection for imidacloprid resistance in Nilaparvata lugens: cross-resistance patterns and possible mechanisms. Pest Management Science 59: 1355–1359. Google Scholar

19.

N Pedrini , R Crespo , MP. Juarez 2007. Biochemistry of insect epicuticle degradation by entomopathogenic fungi. Comparative Biochemistry and Physiology 146: 124–137. Google Scholar

20.

SA Pelizza , LA Eliades , AC Scorsetti , MN Cabello , CE. Lange 2012. Entomopathogenic fungi from Argentina for the control of Schistocerca cancellata (Orthoptera: Acrididae) nymphs: fungal pathogenicity and enzyme activity. Biocontrol Science and Technology 22: 1119–1129. Google Scholar

21.

W Petlamul , P. Prasertsan 2012. Evaluation of strains of Metarhizium anisopliae and Beauveria bassiana against Spodoptera litura on the basis of their virulence, germination rate, conidia production, radial growth and enzyme activity. Mycobiology 40: 111–116. Google Scholar

22.

R Peveling , SA. Demba 1997. Virulence of the entomopathogenic fungus Metarhizium flavoviride Gams and Rozsypal and toxicity of diflubenzuron, fenitrothion-esfenvalerate and profenofos-cypermethrin to nontarget Arthropods in Mauritania. Archives of environmental contamination and toxicology 32: 69–79. Google Scholar

23.

TJ Poprawski , M Marchai , PH. Robert 1985. Comparative susceptibility of Otiorhynchus sulcatus and Sitona lineatus (Coleoptera: Curculionidae) early stage to five entomopathogenic hyphomycetes. Environment Entomology 14: 247–253. Google Scholar

24.

XY Pu , MG Feng , CH. Shi 2005. Impact of three application methods on the field efficacy of a Beauveria bassiana-based mycoinsecticide against the false-eye leafhopper, Empoasca vitis (Homoptera: Cicadellidae), in tea canopy. Crop Protection 24: 167–175. Google Scholar

25.

SD Reay , M Brownbridge , NJ Cummings , TL Nelson , B Souffre , C Lignon , TR. Glare 2008. Isolation and characterization of Beauveria spp. associated with exotic bark beetles in New Zealand Pinu sradiata plantation forests. Biological Control 46: 484– 494. Google Scholar

26.

JL Reissig , JL Strominger , LF. Leloir 1955. A modified colorimetric method for the estimation of n-acetyl amino sugars. Journal of Biological Chemistry 217: 959–966. Google Scholar

27.

DW Roberts , RJ. Leger 2004. Metarhizium spp., cosmopolitan insect-pathogenic fungi: mycological aspects. Advances in Applied Microbiology 54: 1–70. Google Scholar

28.

M Rodriguez , M Gerding , A France , R. Ceballos 2009. Evaluation of Metarhizium anisopliae var. anisopliae Qu-M845 isolate to control Varroa destructor (Acari: Varroidae) in laboratory and field trials. Chilean Journal of Agricultural Research 69: 541–547. Google Scholar

29.

KDZ Samuels , JB Heale , M. Llewellyn 1989. Characteristics relating to the pathogenicity of Metarhizium anisopliae toward Nilaparvata lu gens. Journal of Invertebrate Pathology 53: 25–31. Google Scholar

30.

A Schrank , MH. Vainstein 2010. Metarhizium anisopliae enzymes and toxins. Toxicon 56: 1267–1274. Google Scholar

31.

LT Shan , MG. Feng 2010. Evaluation of the biocontrol potential of various Metarhizium isolates against green peach aphid Myzus per si cae (Homoptera: Aphididae). Pest Management Science 66: 669–675. Google Scholar

32.

RJ Leger St , AK Charnley , RM. Cooper 1986. Cuticle-degrading enzymes of entomopathogenic fungi: mechanisms of interaction between pathogen enzymes and insect cuticle. Journal of Invertebrate Pathology 47: 295–302. Google Scholar

33.

RJ Leger St , RM Cooper , AK. Charnley 1991. Characterization of chitinase and chitobiase produced by the entomopathogenic fungus Metarhizium anisopliae. Journal of Invertebrate Pathology 58: 415–426. Google Scholar

34.

RJ Leger St , RC Staples , DW. Roberts 1993. Entomopathogenic isolates of Metarhizium anisopliae, Beauveria bassiana and Aspergillus flavus produce multiple extracellular chitinase isozymes. Journal of Invertebrate Pathology 61: 81–84. Google Scholar

35.

QY Tang , MG. Feng 2007. DPS date processing system: experimental design, statistical analysis and data mining. Science Press. Google Scholar

36.

AV Toledo , AMMD Lenicov , CCL. Lastra 2007. Pathogenicity of fungal isolates (Ascomycota: Hypocreales) against Peregrinus maidis, Delphacodes kuscheli (Hemiptera: Delphacidae), and Dalbulus maidis (Hemiptera: Cicadellidae), vectors of corn diseases. Mycopathology 163: 225–232. Google Scholar

37.

AV Toledo , L Giambelluca , AMMD Lenicov , CCL. Lastra 2008. Pathogenic fungi of planthoppers associated with rice crops in Argentina. International Journal of Pest Management 54: 163–168. Google Scholar

38.

AV Toledo , AMMD Lenicov , CCL. Lastra 2010. Histopathology caused by the entomopathogenic fungi, Beauveria bassiana and Metarhizium anisopliae, in the adult planthopper, Peregrinus maidis, a Maize Virus Vector. Journal of Insect Science 10:35. Available online:  www.insectscience.org/10.35 Google Scholar

39.

JD Vandenberg , LE Sandvol , ST Jaronski , MA Jackson , EJ Souza , SE. Halbert 2001. Efficacy of fungi for control of Russian wheat aphid (Homoptera: Aphididae) in irrigated wheat. Southwestern Entomology 26: 73–85. Google Scholar

40.

LW Zhang , Y J Liu , J Yao , B Wang , B Huang , ZZ Li , MZ. Fan 2011. Evaluation of Beauveria bassiana (Hyphomycetes) isolates as potential agents for control of Dendroctonus valens. Insect Science 18: 209– 216. Google Scholar

41.

G. Zimmermann 1982. Untersuchungen zur Wirkung von Metarhizium anisopliae (Metsch.) Sorok. auf Eier und Schlüpfende Eilarvenvon Otiorhynchus sulcatus F. (Col., Curculionidae). Journal of Applied Entomology 93: 476–482. Google Scholar
Copyright: This is an open access paper. We use the Creative Commons Attribution 3.0 license that permits unrestricted use, provided that the paper is properly attributed.
Maoye Li, Shiguang Li, Amei Xu, Huafeng Lin, Dexin Chen, and Hui Wang "Selection of Beauveria Isolates Pathogenic to Adults of Nilaparvata lugens," Journal of Insect Science 14(32), 1-12, (1 May 2014). https://doi.org/10.1673/031.014.32
Received: 29 January 2013; Accepted: 21 August 2013; Published: 1 May 2014
KEYWORDS
biological characteristics
chitinase
microbial control
virulence
Back to Top