scholarly journals First Report of Wilt of Rubber Tree Caused by Chalaropsis thielavioides in China

Plant Disease ◽  
2020 ◽  
Author(s):  
Xue Li ◽  
Jie Li ◽  
Hua Yong Bai ◽  
Kecheng Xu ◽  
Ruiqi Zhang ◽  
...  

Rubber tree (Hevea brasiliensis (Willd. ex Adr. Juss) Müll. Arg.) is used for the extraction of natural rubber and is an economically and socially important estate crop commodity in many Asian countries such as Indonesia, Malaysia, Thailand, India, Sri Lanka, China and several countries in Africa (Pu et al, 2007). Xishuangbanna City and Wenshan City are the main rubber cultivation areas in Yunnan Province, China. In November 2012, rubber tree showing typical wilt symptoms (Fig. 1 A) and vascular stains (Fig. 1 B) were found in Mengla County, Xishuangbanna City. This disease was destructive in these trees and plant wilt death rate reached 5%. The diseased wood pieces (0.5cm long) from trunk of rubber was surface disinfected with 75% ethanol for 30s and 0.1% mercuric chloride (HgCl2) for 2min, rinsed three times with sterile distilled water, plated onto malt extract agar medium (MEA), and incubated at 28℃. After 7 days, fungal-like filaments were growing from the diseased trunk. Six cultures from 6 rubber trunk were obtained and incubated on MEA at 28℃, after 7 days to observe the cultural features. The mycelium of each culture was white initially on MEA, and then became dark green. Cylindrical endoconidia apices rounded, non-septate, smooth, single or borne in chains (8.9 to 23.6 × 3.81 to 6.3μm) (Fig. 1 C). Chlamydospores (Fig. 1 D) were abundant, thick walled, smooth, forming singly or in chains (11.1 to 19.2 × 9.4 to 12.0μm). The mould fungus was identifed as Chalaropsis based on morphology (Paulin-Mahady et al. 2002). PCR amplification was carried out for 3 isolates, using rDNA internal transcribed spacer (ITS) primer pairs ITS1F and ITS4 (Thorpe et al. 2005). The nucleotide sequences were deposited in the GenBank data base and used in a Blast search of GenBank. Blast analysis of sequenced isolates XJm8-2-6, XJm8-2 and XJm10-2-6 (accessions KJ511486, KJ511487, KJ511489 respectively) had 99% identity to Ch. thielavioides strains hy (KF356186) and C1630 (AF275491). Thus the pathogen was identified as Ch. thielavioides based on morphological characteristics and rDNA-ITS sequence analysis. Pathogenicity test of the isolate (XJm8-2) was conducted on five 1-year-old rubber seedlings. The soil of 5 rubber seedlings was inoculated by drenching with 40 ml spore suspension (106 spores / ml). Five control seedlings were inoculated with 40 ml of sterile distilled water. All the seedlings were maintained in a controlled greenhouse at 25°C and watered weekly. After inoculated 6 weeks, all the seedlings with spore suspension produced wilt symptoms, as disease progressed, inoculated leaves withered (Fig. 1 E) and vascular stains (Fig. 1 F) by 4 months. While control seedlings inoculated with sterile distilled water remained healthy. The pathogen re-isolated from all inoculated symptomatic trunk was identical to the isolates by morphology and ITS analysis. But no pathogen was isolated from the control seedlings. The pathogenicity assay showed that Ch. thielavioides was pathogenic to rubber trees. Blight caused on rubber tree by Ceratocystis fimbriata previously in Brazil (Valdetaro et al. 2015), and wilt by Ch. thielavioides was not reported. The asexual states of most species in Ceratocystis are “chalara” or “thielaviopsis” (de Beer et al. 2014). To our knowledge, this is the first report of this fungus causing wilt of rubber in China. The spread of this disease may pose a threat to rubber production in China.

Plant Disease ◽  
2022 ◽  
Author(s):  
Kecheng Xu ◽  
Ruiqi Zhang ◽  
Haixia Lu ◽  
Jinglong Zhang ◽  
Jing Yang ◽  
...  

Coffee (Coffea arabica L.) is one of the most important agricultural commodities in the world market. As an important cash crop in China, coffee is cultivated mainly in Yunnan and Hainan provinces. During October 2013 and September 2020, coffee trees showing typical dieback and wilt symptoms were found in Nanping town (N 22° 38', E 101° 0'), Pu’er, and Puwen town (N 22° 32', E 101° 4'), Xishuangbanna in Yunnan province, China. Symptomatic trees initially exhibited yellowing of foliage, expanding in size along the leaf margin, then became blighted and dry, and the internal xylem in main stem discolored brown to black. Infected trees eventually developed dieback and wilt. Disease incidence ranged from 10% to 22% and 25% to 40% of crown symptoms in the affected coffee trees. In extreme cases, 50% out of 380 trees were affected. The stems of coffee trees with typical symptoms were collected, and then the diseased tissues were surface disinfected with 75% ethanol for 30 s and 0.1% mercuric chloride (HgCl2) solution for 2 min, rinsed three times with sterile distilled water, plated onto potato dextrose agar (PDA) medium, and incubated at 25°C. After 6 days, fungal mycelium was observed growing from the tissue. Three isolates (C3-1, C3-2, and C3-2-1) were obtained by picking spore masses from the apices of perithecia and transferring them to malt extract agar (MEA) medium and incubated at 25°C for 10 days to observe the cultural features. In culture, colonies reaching 65 mm within 10 days, mycelium initially white, then becoming light blue-green. After 6 days of formation, perithecia were black, globose (123.8 - 173.4 μm × 138.2 - 180.6 μm), and showed a long black neck (414.2 - 650.0 μm). Ascospores with outer cell wall forming a brim, hat-shaped, accumulating in a mucilaginous mass at the tips of ostiolar hyphae (4.3 μm × 6.0 μm). Cylindrical endoconidia (14.1 - 45.2 μm × 3.5 - 5.7 μm) were hyaline. Chain of barrel-shaped conidia (6.6 - 10.2 μm × 6.8 - 8.8 μm) were found. Aleuroconidia (10.8 - 16.9 μm × 9.1 - 13.0 μm) were olive-brown, ovoid or obpyriform, and smooth. Morphological characteristics of the fungus were consistent with the description of Ceratocystis fimbriata Ellis & Halst. (Engelbrecht and Harrington 2005). The three isolates were used for molecular identification, and their genomic DNA was extracted using the chelex-100 method (Xu et al. 2020). The internal transcribed spacer (ITS) region of rDNA was sequenced using the procedures of Thorpe et al. (2005). Analysis of the ITS sequence data (GenBank accessions KY580836, KJ511480, and KJ511479) showed that the isolates were 100% homologous to isolates of C. fimbriata from Punica granatum, Camellia sinensis, and Cucumis sativus in China (GenBank accessions KY580891, KY580870, and MH535909, respectively) by BLAST analysis. Neighbor-joining (NJ) phylogenetic analysis was performed using MEGA 6.06 based on the ITS sequences. The three isolates were clustered on the same clade with other C. fimbriata isolates with a high bootstrap value (90%). Therefore, the fungus was identified as C. fimbriata based on both morphological and molecular characteristics. Pathogenicity of the three isolates was tested by inoculating one-year-old pot grown coffee seedlings (C. arabica) through drenching the loams with 30 ml spore suspension (1 × 106 spores/ml). Control plants were inoculated with 30 ml of sterile distilled water. The trees were kept in a controlled greenhouse at 25°C and watered weekly. One month after inoculation, all inoculated plants produced typical dieback and wilt symptoms, whereas the control trees showed no symptoms. The same fungus was isolated from the inoculated trees on PDA and identified as C. fimbriata according to the methods described above, and no fungal growth was observed in the controls, thus fulfilling the Koch's postulates. Coffee canker disease caused by C. fimbriata has been reported in Indonesia and Colombia (Marin et al. 2003). To our knowledge, this is the first report of C. fimbriata causing canker disease of coffee trees in China.


Plant Disease ◽  
2022 ◽  
Author(s):  
Kecheng Xu ◽  
Ruiqi Zhang ◽  
Xue Li ◽  
Jinglong Zhang ◽  
Hang Zhao ◽  
...  

Eucalyptus species are widely planted in the tropics and subtropics, and eucalyptus is among the most important cash crop in Southern China. One of the most important diseases on eucalyptus is Ceratocystis wilt, caused by the fungus Ceratocystis fimbriata Ellis & Halst., and the genus name Chalaropsis has been proposed for anamorphs of Ceratocystis species (de Beer et al. 2014). During April 2018, severely infected Eucalyptus robusta trees were observed in Kunming, Yunnan Province, China. Symptomatic trees initially exhibited yellowing and wilting of foliage on individual branches, then spread to the whole canopy, sometimes followed by death of the whole tree. Reddish-brown to dark-brown discoloration in the woody xylem of affected trees, sometimes a grayish white layer of fungal growth may be seen. The disease was observed on 16% of trees surveyed. The base of trunks with typical symptoms were collected, then the discolored xylem tissues were surface disinfected with 75% ethanol for 30 s and 0.1% mercuric chloride (HgCl2) solution for 2 min, rinsed three times with sterile distilled water, plated onto potato dextrose agar (PDA) medium, and incubated at 25°C. After 6 days, a fungus was consistently observed growing from the tissue. Three isolates were obtained. In culture, colonies reaching 54mm diam within 15 days, mycelium initially white, then becoming celadon. Endoconidia unicellular, smooth, cylindrical, straight, biguttulate, 11.21 - 32.26 × 4.12 - 5.25 μm. Phialides produced on short, septate, aerial hyphae, lageniform and chain of phialoconidia (3.62 - 5.89 × 31.39 - 65.76 μm) were also observed. Chlamydospores (11.45 - 14.26 × 10.06 - 12.22 μm) were single, dark, thick-walled. Morphological characteristics of the fungus were consistent with the description of Chalaropsis thielavioides (Paulin-Mahady et al. 2002). The two of three isolates were used for molecular identification and genomic DNA was extracted from isolates (EKY2-2-1, EKY2-2-2) using the chelex-100 method (Xu et al. 2020). The ITS region of rDNA was sequenced using the procedures of Thorpe et al. (2005). Analysis of ITS sequence data (GenBank accessions MW242701, MW242702) showed that the isolates were 99% - 100% homologous to isolates of C. thielavioides from Hevea rubber, Monstera deliciosa L. and ants in China and Rosa sp. in Australia (GenBank accessions KT963172, KJ511482, KT963173 and KX954598) by BLAST analysis. Neighbor-joining (NJ) phylogenetic analysis were performed using MEGA 6.06 based on ITS sequences (Fig 1), the evolutionary distances were computed using the Maximum Composite Likelihood method. Analyses showed that both isolates (EKY2-2-1, EKY2-2-2) located on the same clade with all C. thielavioides, and clustered with the C. thielavioides strains with high bootstrap support (97% - 100%). Therefore, the fungus was identified as C. thielavioides based on morphology and molecular evidences. Pathogenicity of C. thielavioides was tested by inoculation of six one-year-old pot grown Eucalyptus citriodora seedlings. The sterilized soil of six seedlings was inoculated by drenching with 20 ml spore suspension (2.0 × 106 spores / ml). Control plants were inoculated with 20 ml of sterile distilled water. The seedlings were kept in a controlled greenhouse at 25°C and watered weekly. After one month incubation, all the isolates produced wilt symptoms, whereas control trees showed no symptoms. The original fungus was successfully re-isolated from inoculated trees and identified as C. thielavioides according to the methods described above, and no fungal growth was observed in the controls, thus satisfying Koch's postulates. Although wilt and canker caused by Ceratocystis fimbriata on eucalyptus have been previously reported in Brazil, Uruguay, Uganda, China and Pakistan (Ferreira et al. 1999; Li et al. 2014; Alam et al. 2017), eucalyptus wilt caused by C. thielavioides has not been reported anywhere. Also, wilt of rubber tree and postharvest rot on carrot caused by C. thielavioides have been reported (Li et al. 2021; Xu et al. 2020). To our knowledge, this is the first report of eucalyptus wilt caused by C. thielavioides in China.


Plant Disease ◽  
2022 ◽  
Author(s):  
Kecheng Xu ◽  
Ruiqi Zhang ◽  
Jie Li ◽  
Xue Li ◽  
Jing Yang ◽  
...  

The rubber tree (Hevea brasiliensis) is an important economic resource for the rubber and latex industry. During November 2013 and June 2016, rubber trees showing typical wilt symptoms were found in Mengla, Xishuangbannan, Yunnan, China (N 21° 28', E 101° 33'). Symptomatic trees initially exhibited wilting of foliage on individual branches, then spread to the whole canopy, finally followed by death of the whole tree. Dark-blue to black discoloration was observed in the inner bark and affected xylem, a grayish layer of fungal growth and sporulation occasionally. The disease was detected on 20% of trees surveyed. The diseased tissues of three rubber trees were surface disinfected with 75% ethanol for 30 s and 0.1% mercuric chloride (HgCl2) for 2 min, rinsed three times with sterile distilled water, plated onto potato dextrose agar (PDA), and incubated at 25°C. After 7 days, a fungus was consistently observed growing from the tissue. Three single-spore isolates were obtained. In culture, colonies reaching 69 mm diam within 10 days, mycelium was initially white, then becoming celadon. After 5 days of perithecium formation, observed perithecia were black, globose (173.1 - 237.9 × 175.6 - 217.2 μm) and showed a long black neck (507.3 - 794.1 μm). Ascospore with outer cell wall forming a brim, hat-shaped at the tips of ostiolar hyphae (3.43 × 5.63 μm). Cylindrical endoconidia (10.5 - 39.7 × 3.5 - 6.6 μm) were hyaline. Chain of barrel-shaped conidia (7.2 - 9.5 × 4.1 - 6.2 μm) was found. Aleuroconidia were ovoid or obpyriform, and smooth (10.2 - 14.1 × 8.4 - 10.6 μm). Morphological characteristics of the fungus were consistent with the description of Ceratocystis fimbriata (Engelbrecht and Harrington 2005). The genomic DNA was extracted from isolates (XJm10-2-5, XJm8-2-5, XJm4) using the Chelex-100 method (Xu et al. 2020). The ITS region of rDNA was sequenced using the procedures of Thorpe et al. (2005). Analysis of ITS sequence data (GenBank accessions KJ511488, KJ511485, KT963149) showed that the isolates were 100% homologous to those of the isolates on Punica granatum and Colocasia esculenta from China (GenBank accessions KT963152, MH793673) by BLAST analysis. Neighbor-joining phylogenetic analyse were performed using MEGA 6.06 based on ITS sequences (Fig. 1). Analyses showed that all isolates located on the same clade with all C. fimbriata with a high bootstrap support. Therefore, the fungus was identified as C. fimbriata based on morphology and molecular evidences. Pathogenicity of C. fimbriata isolated from this study was tested by inoculation of three one-year-old pot-grown (3L) seedlings of rubber tree. The soil of three seedlings was inoculated by drenching with 30 ml spore suspension (2.0 × 106 spores / ml). Three control plants were inoculated with 30 ml of sterile distilled water. The experiment was repeated three times. The plants were kept in a controlled greenhouse at 25°C and watered weekly. After the inoculation for one month, all the isolates produced typical wilt symptoms, while control plants showed no symptoms. The original fungus was successfully re-isolated from inoculated trees and identified as C. fimbriata according to the methods described above. The pathogenicity assay showed that C. fimbriata was pathogenic to rubber trees. C. fimbriata was first reported on rubber tree in Brazil (Albuquerque et al. 1972; Silveira et al. 1985). To the best of our knowledge, this is the first report of C. fimbriata causing wilt of rubber tree in China. This finding contributes to understanding the diversity of this pathogen, and it appears to be a significant threat to rubber trees in its ecosystem.


Plant Disease ◽  
2011 ◽  
Vol 95 (7) ◽  
pp. 874-874 ◽  
Author(s):  
Y. M. Shen ◽  
C. H. Chao ◽  
H. L. Liu

Gynura bicolor (Roxb. ex Willd.) DC., known as Okinawa spinach or hong-feng-cai, is a commonly consumed vegetable in Asian countries. In May 2010, plants with blight and wilt symptoms were observed in commercial vegetable farms in Changhua, Taiwan. Light brown-to-black blight lesions developed from the top of the stems to the petioles and extended to the base of the leaves. Severely infected plants declined and eventually died. Disease incidence was approximately 20%. Samples of symptomatic tissues were surface sterilized in 0.6% NaOCl and plated on water agar. A Phytophthora sp. was consistently isolated and further plated on 10% unclarified V8 juice agar, with daily radial growths of 7.6, 8.6, 5.7, and 2.4 mm at 25, 30, 35, and 37°C, respectively. Four replicates were measured for each temperature. No hyphal growth was observed at 39°C. Intercalary hyphal swellings and proliferating sporangia were produced in culture plates flooded with sterile distilled water. Sporangia were nonpapillate, obpyriform to ellipsoid, base tapered or rounded, and 43.3 (27.5 to 59.3) × 27.6 (18.5 to 36.3) μm. Clamydospores and oospores were not observed. Oospores were present in dual cultures with an isolate of P. nicotianae (p731) (1) A2 mating type, indicating that the isolate was heterothallic. A portion of the internal transcribed spacer sequence was deposited in GenBank (Accession No. HQ717146). The sequence was 99% identical to that of P. drechsleri SCRP232 (ATCC46724) (3), a type isolate of the species. The pathogen was identified as P. drechsleri Tucker based on temperature growth, morphological characteristics, and ITS sequence homology (3). To evaluate pathogenicity, the isolated P. drechsleri was inoculated on greenhouse-potted G. bicolor plants. Inoculum was obtained by grinding two dishes of the pathogen cultured on potato dextrose agar (PDA) with sterile distilled water in a blender. After filtering through a gauze layer, the filtrate was aliquoted to 240 ml. The inoculum (approximately 180 sporangia/ml) was sprayed on 24 plants of G. bicolor. An equal number of plants treated with sterile PDA processed in the same way served as controls. After 1 week, incubation at an average temperature of 29°C, blight and wilt symptoms similar to those observed in the fields appeared on 12 inoculated plants. The pathogen was reisolated from the lesions of diseased stems and leaves, fulfilling Koch's postulates. The controls remained symptomless. The pathogenicity test was repeated once with similar results. G. bicolor in Taiwan has been recorded to be infected by P. cryptogea (1,2), a species that resembles P. drechsleri. The recorded isolates of P. cryptogea did not have a maximal growth temperature at or above 35°C (1,2), a distinctive characteristic to discriminate between the two species (3). To our knowledge, this is the first report of P. drechsleri being associated with stem and foliar blight of G. bicolor. References: (1) P. J. Ann. Plant Pathol. Bull. 5:146, 1996. (2) H. H. Ho et al. The Genus Phytophthora in Taiwan. Institute of Botany, Academia Sinica, Taipei, 1995. (3) R. Mostowfizadeh-Ghalamfarsa et al. Fungal Biol. 114:325, 2010.


Plant Disease ◽  
2015 ◽  
Vol 99 (2) ◽  
pp. 287-287 ◽  
Author(s):  
G. Z. Wang ◽  
M. P. Guo ◽  
Y. B. Bian

Coprinus comatus is one of the most commercially important mushrooms in China. Its fruiting body possesses rich nutritional and medicinal value. In November 2013, unusual symptoms were observed on C. comatus on a mushroom farm in Wuhan, Hubei, China. At first, fruiting bodies were covered by white and cobweb-like mycelia. Later, the cap and stipe turned brown or dark before rotting and cracking. The pathogen was isolated from infected tissue of C. comatus. Colonies of the pathogen on potato dextrose agar (PDA) medium first appeared yellowish, followed by an obvious ochraceous or pinkish color. Aerial mycelia grew along the plate wall, cottony, 1 to 4 mm high. Conidiophores were borne on the tops of hyphae, had two to four branches, and were cylindrical, long clavate, or fusiform. Conidia were borne on the tops of the branches of conidiophores, had one to two separates, and were long and clavate. The spores ranged from 15.3 to 22.1 μm long and were 5.1 to 8.3 μm wide, which was consistent with the characteristics of Cladobotryum protrusum (1). The species was identified by ribosomal internal transcribed spacer sequencing. The ribosomal ITS1-5.8S-ITS2 region was amplified from the isolated strain using primers ITS1 and ITS4. A BLAST search in GenBank revealed the highest similarity (99%) to C. protrusum (GenBank Accession Nos. FN859408.1 and FN859413.1). The pathogen was grown on PDA at 25°C for 3 days, and the inoculation suspension was prepared by flooding the agar surface with sterilized double-distilled water for spore suspension (1 × 105 conidia/ml). In one treatment, the suspension was sprayed on casing soil (106 conidia/m2) and mixed thoroughly with it, then cased with treated soil for 2 to 3 cm thickness on the surface of compost in cultivation pots (35 × 25× 12 cm), with sterile distilled water as a control (2). Eight biological replicates were included in this treatment. In the second treatment, mycelia plugs (0.3 × 0.3 cm) without spore production were added to 20 fruiting bodies. Mushrooms treated with blank agar plugs (0.3 × 0.3 cm) were used as a control. The plugs were covered with sterilized cotton balls to avoid loss of moisture. Tested cultivation pots were maintained at 18°C and 85 to 95% relative humidity. In the samples where casing soil was sprayed with conidia suspension, white mildew developed on the pileus, and a young fruiting body grew out from the casing soil. Eventually, the surface of the mushroom was overwhelmed by the mycelia of the pathogen and the pileus turned brown or black. For the other group inoculated with mycelia plugs, only the stipe and pileus inoculated with mycelia turned brown or dark; it rotted and cracked 2 to 3 days later. The symptoms were similar to those observed on the C. comatus cultivation farm. Pathogens re-isolated from pathogenic fruiting bodies were confirmed to be C. protrusum based on morphological characteristics and ITS sequence. To our knowledge, this is the first report of the occurrence of C. protrusum on the edible mushroom C. comatus (3). Based on the pathogenicity test results, C. protrusum has the ability to severely infect the fruiting body of C. comatus. References: (1) K. Põldmaa. Stud. Mycol. 68:1, 2011. (2) F. J. Gea et al. Plant Dis. 96:1067, 2012. (3) W. H. Dong et al. Plant Dis. 97:1507, 2013.


Plant Disease ◽  
2012 ◽  
Vol 96 (9) ◽  
pp. 1374-1374 ◽  
Author(s):  
M. K. Kim ◽  
Y. H. Lee ◽  
K. M. Cho ◽  
J. Y. Lee

Pleurotus eryngii is one of the most commercially important mushrooms in Korea. In May 2009, unusual symptoms were observed in P. eryngii grown in mushroom farms in Changnyeong and Hapcheon, in Gyeong-nam Province, Korea. One of the main symptoms was cobweb-like growth of fungal mycelia over the mushroom surface. Colonies on the surface rapidly overwhelmed the mushrooms, which turned pale brown or yellow. Mushrooms eventually turned dark brown and became rotten. Colonies of the isolates on potato dextrose agar (PDA) were yellowish, and a reddish or orange color was evident in the agar. The colonies grew 20 to 30 mm per day on PDA. Large spores with a single septum were produced on vertically branched conidiophores bearing two to four, mostly three to four, sporogenous cells, ranging from 17.2 to 20.5 μm long and 8.0 to 10.2 μm thick. The shape of the conidia was ellipsoid and obovoid. These morphological characteristics are consistent with descriptions of Cladobotryum mycophilum, a causal agent of cobweb disease in Agaricus bisporus (1,4). To identify the isolated fungal pathogen, the ITS region was amplified with ITS1 and ITS4 primers and sequenced. The sequence data from the isolate was deposited in GenBank (Accession No. JF693809). A BLAST search showed that the isolated strain belonged to a species of Cladobotryum. The highest similarity (99.5%) was to the ITS sequence of C. mycophilum (teleomorph Hypomyces odoratus) (GenBank Accession Nos. JF505112 and Y17096) (3,4). The strain that was tested for pathogenicity was grown on PDA at 25°C for 72 h. The inoculum was prepared by flooding the agar surface with 10 ml of sterilized double distilled water and scraping it with a spatula. The resulting spore suspension was filtered through three layers of cheesecloth. Conidial concentration was adjusted with a hemacytometer to 1 × 106 conidia ml–1. A conidia suspension was inoculated onto each of several stages of mushroom cultivation with a pipette. The control was spotted with double distilled water. In the case of infection during the inoculation and spawn running stages, the fungal mycelia colonized the media and hampered development of the mycelium of P. eryngii. In the regeneration and primordia formation stages of the host, the mycelium of the pathogen covered the surface of the plastic bottle containing the substrates and developed many spores. In the growing and harvesting stages, the surface of mushroom was overwhelmed by the mycelium of the fungal pathogen and turned pale or dark brown, accompanied by cracking of the stipe surface and finally rotting with a foul odor. These symptoms were similar to the observation from natural infection. The symptoms of the cobweb-like disease in A. bisporus (1,2) were observed within 5 to 7 days of inoculation with conidia suspensions of C. mycophilum. Fungi isolated from inoculated mushrooms were shown to be identical, based on phenotypic characteristic, to the inoculated strain used in these pathogenicity tests. No symptoms were observed on controls. To our knowledge, this is the first report on the occurrence of C. mycophilum on the edible mushroom P. eryngii in Korea. Based on the pathogenicity test results, the pathogen could attack P. eryngii in any cultivation stage, making it a potentially serious fungal pathogen in P. eryngii. References: (1) C. G. Back et al. J. Gen. Plant Pathol. 76:232, 2010. (2) R. H. Gaze. Mushroom J. 546:23, 1995. (3) F. J. Gea et al. Plant Dis. 95:1030, 2011. (4) H. M. Grogan and R. H. Gaze. Mycol. Res. 104:357, 2000.


Plant Disease ◽  
2021 ◽  
Author(s):  
Shuwu Zhang ◽  
Jinhuan Chen ◽  
Lijun Ma ◽  
Enchen Li ◽  
Baoli Ji ◽  
...  

Wilting of branches and leaves was observed on 4-5 year old apple trees of the varieties Delicious and Fuji in orchards located in Wushan, Gansu Province, China in April 2018. Subsequently, the stem vascular tissue and woody xylem became discolored and necrotic. The stem dieback expanded rapidly to the entire vasculature of the branches. Finally, the epidermis of the stem bases split and was covered with light pink mold. For the pathogen isolation, 25 symptomatic stems were collected from 25 symptomatic trees in 3 individual orchards. Fragments (approximately 0.5 cm in length × 0.5 cm in width) of symptomatic stems were surface sterilized and individually transferred to Petri dishes containing potato dextrose agar (PDA), and incubated for 4 days at 25°C. Five types of isolates with distinct morphological characteristics (PJ1 to PJ5) were obtained from the 25 symptomatic stems after the single spore inoculation and sub-culture. The isolation frequency of PJ1, PJ2, PJ3, PJ4 and PJ5 types was 11%, 8%, 100%, 4% and 13%, respectively, in the 25 symptomatic stems. A spore suspension of PJ1, PJ2, PJ3, PJ4 and PJ5 types was prepared by adding 5 ml of sterile distilled water in the 14-day old culture colonies and filtered through 0.22 mm Millipore membranes, and the final concentration was adjusted to 108 per ml for inoculation. Detached healthy apple stems (15 cm in length) were surface-disinfested and used to evaluate the pathogenicity of PJ1 (7 isolates), PJ2 (5 isolates), PJ3 (32 isolates), PJ4 (2 isolates) and PJ5 (9 isolates) by dipping the stems into sterilised tubs containing the spore suspension (108 per ml) of each isolate. Apple stems dipped in sterile distilled water served as the control. Each control and treatment were repeated 3 times. At day 15 and 35, the stems infected with the spore suspension of PJ3 isolates developed symptoms that were similar to those observed in the apple orchards. However, the other four types (PJ1, PJ2, PJ4 and PJ5) exhibited either no symptoms or different symptoms from those observed in the apple orchards. There were no symptoms on the control stems. After the colony of the pathogen (PJ3 type) was re-isolated from the infected stem bases 35 days inoculation. The PJ3 type isolates with same morphological characteristics as the original PJ3 type isolates were used for further examination and identification. After 4 days of incubation on PDA, the colonies of PJ3 type isolates developed velvety aerial mycelia with white or light pink color when they were viewed from the front/top side of the PDA and orange-red color when they were viewed from the reverse/bottom side. After 14 days of incubation, the color in the centre of the colonies changed to yellow green in the front view and carmine red in the reverse view of the plates. Three types of spores (microconidia, macroconidia and chlamydospores) were observed after incubation of PJ3 type isolates for 14 days. The size (width and length) of 30 conidia in each of PJ3 type isolates was measured and averaged. The microconidia were abundant on aerial mycelia and limoniform, oval or pyriform with 0-1 septa. Their size ranged from 1.94 μm to 8.05 μm in length and 1.48 μm to 3.62 μm in width. The macroconidia were falciform and curved in shape, mostly with 3-5 septa and a size ranging from 13.52 μm to 22.43 μm in length and 2.31 μm to 3.87 μm in width. The chlamydospores were spherical, intercalary and formed in chains on PDA plates. These morphological characteristics indicate that the PJ3 type isolates could be Fusarium tricinctum (Chen et al. 2019; Aktaruzzaman et al. 2018). To confirm the morphological identification, the sequences of internal transcribed spacer (ITS), transcriptional enhancer factor-1 (TEF-lα) and ribosomal RNA large subunit gene (LSU) of the representative isolate PJ3-3 selected from the PJ3 type isolates with same morphological characteristics were sequenced and used for molecular identification (Laurence et al. 2011; Abd-Elsalam et al. 2003; Miller et al. 1996). The sequences of ITS, TEF-lα and LSU of the PJ3-3 isolate were deposited in NCBI database with the accession numbers of MZ799356, MZ820045 and MZ820044, respectively. In BLAST analyses, the obtained sequences of the PJ3-3 isolate showed 99.47%, 100% and 99.01% identity to the corresponding region of F. tricinctum ITS (JX179207.1: 3-566 Fusarium tricinctum isolate Fyx 1), TEF-lα (MK032320.1 F. tricinctum isolate ZD3) and LSU (KC311496.1 Fusarium tricinctum isolate L12), respectively. The phylogenetic analysis clustered the PJ3-3 isolate sequences within the same clade with ITS, TEF-lα and LSU sequences of F. tricinctum isolates. Thus, the PJ3-3 isolate was identified as F. tricinctum based on the pathogenicity tests, morphological characteristics and molecular analyses. Previously, the symptoms of xylem browning and dieback were observed in the twigs of wild apple trees that were collected from wild apple forests, and the species F. avenaceum, F. solani, F. tricinctum, F. proliferatum, and F. sporotrichioides were isolated from diseased wild apple trees (Chen et al. 2019). Only F. avenaceum, F. solani, F. proliferatum, and F. sporotrichioides were reported as the pathogens causing the disease symptoms of xylem browning and dieback in wild apple trees in Xinjiang, China (Chen et al. 2019). In our present study, we found that F. tricinctum can cause stem vascular and woody xylem browning, wilting, and dieback in the apple tree varieties Delicious and Fuji. These are new symptoms discovered in our present research and different from the previous paper (Chen et al. 2019). Therefore, to our knowledge, this study is the first report of F. tricinctum causing a new disease on apple trees in China following Koch’s postulates. Our findings are important for the management of apple disease and protect apple trees in the future.


Plant Disease ◽  
2012 ◽  
Vol 96 (9) ◽  
pp. 1380-1380 ◽  
Author(s):  
A. Garibaldi ◽  
G. Gilardi ◽  
G. Ortu ◽  
M. L. Gullino

Lettuce (Lactuca sativa L.) is widely grown in Italy, with the production for the preparation of ready-to-eat salads becoming increasingly important. During the spring of 2011, a previously unknown leaf spot was observed on L. sativa plants, cv Rubia, grown in several plastic tunnels in Lumbardy (northern Italy), 20 to 25 days after sowing. Thirty to forty per cent of leaves of the plants growing in the part of the tunnel with the highest relative humidity were affected. Leaves of infected plants showed extensive, irregular, dark brown, necrotic lesions with a chlorotic halo. Lesions initially ranged from 0.5 to 3 mm, then eventually coalesced, reaching 2 to 3 cm, showing a well-defined, dark brown border. Affected leaves senesced and withered. The crown was not affected by the disease. Diseased tissue was excised, immersed in a solution containing 1% sodium hypochlorite for 60 s, rinsed in water, then cultured on potato dextrose agar (PDA), amended with 25 mg/l of streptomycin sulphate. After 5 days, a fungus developed, producing a greenish grey mycelium with a white border when incubated under 12 h/day of fluorescent light at 21 to 23°C. In order to favor the production of conidia, the fungus was transferred on malt extract agar (MA) and maintained under 12 h/day of fluorescent light at 22°C. After 15 days, black pycnidia, 175 to 225 μm, developed, with hyaline, elliptical, unicellular conidia, measuring 3.21 to 6.7 × 1.08 to 3.2 (average 5.5 × 1.9) μm. On the basis of these morphological characteristics, the fungal causal agent of the disease could be related to the genus Phoma (2). The internal transcribed spacer (ITS) region of rDNA of the isolate PHT30 was amplified using the primers ITS1/ITS4 and sequenced. BLAST analysis (1) of the 466-bp segment showed a 99% similarity with the sequence of Phoma tropica (GenBank Accession No. JF923820.1). The nucleotide sequence has been assigned the GenBank Accession No. JQ954396. Pathogenicity tests were performed by spraying healthy 20-day-old lettuce plants, cv Rubia, with a spore suspension (1 × 105 conidia/ml) prepared from 14-day-old colonies of the strain PHT30 grown on MA cultures. Plants inoculated with water alone served as controls. Ten plants per isolate were used. Plants were covered with plastic bags for 5 days after inoculation and maintained in a growth chamber at 20°C and 80% relative humidity. The first foliar lesions, similar to those occurring on the naturally infected plants, developed on leaves 12 days after inoculation. Control plants remained healthy. The pathogen was consistently reisolated from leaf lesions. The pathogenicity test was completed twice. To our knowledge, this is the first report of the presence of P. tropica on lettuce in Italy as well as worldwide. In the United States, the presence of P. exigua was reported in 2006 (3). The economic importance of the disease at present is limited, probably also because symptoms can be confused with those caused by Botrytis cinerea. However, P. tropica could become a more significant problem because of the importance of the crop. References: (1) S. F. Altschul et al. Nucleic Acids Res. 25:3389, 1997. (2) G. H. Boerema. Trans. Br. Mycol. Soc. 67:289, 1976. (3) S. Y. Koike. Plant Dis. 90:1268, 2006.


Plant Disease ◽  
2021 ◽  
Author(s):  
Donghun Kang ◽  
Jungyeon Kim ◽  
Youn Mi Lee ◽  
Balaraju Kotnala ◽  
Yongho Jeon

In September 2020, typical anthracnose symptoms were observed on cotton (Gossypium indicum Lam.) leaves growing in Hahoe village, Andong, Gyeongbuk Province, Korea. The leaves of the infected plants initially showed spots with halo-lesions which became enlarged and spread to the entire leaf surface area. The infected leaves later became yellowish and chlorotic (Fig. 1A). The disease incidence was at least 90% in the field. For pathogen isolation, fresh samples collected from symptomatic leaves were cut into small pieces (4 to 5 mm2), surface-sterilized in 1% sodium hypochlorite for 1 min, rinsed three times, and macerated in sterile distilled water (SDW). They were spread onto potato dextrose agar (PDA) plates and incubated at 25 °C for 5 days under a 12-h photoperiod. Five isolates were recovered from the infected leaves. Purified fungal colonies were initially white, later turned yellow on PDA medium. Conidia were yellow-colored, smooth-walled, aseptate, straight or slightly distorted, and cylindrical with one end slightly acute or with broadly rounded ends, and with size ranges from 15.3 to 17.5 µm (length) × 4.5 to 5.2 µm (width) (Fig. 1B). The morphological characteristics of the present isolates were consistent with those of Colletotrichum gloeosporioides (Weir et al. 2012). A single isolate, ANUK97, was selected for identification. The multilocus sequence analysis (MLSA) of the actin (ACT), calmodulin (CAL), glyceraldehyde-3-phosphate dehydrogenase (GAPDH), internal transcribed spacer (ITS) rDNA, and β-tubulin (Tub2) were amplified by PCR with the primer pairs of ACT-521F/ACT-783R, CL1C/CL2C, GDF/GDR, ITS1/ITS4, and T1/T2, respectively (White et al. 1990). The resulting sequences were deposited in GenBank under accession numbers MW580367 (ACT), MW580368 (CAL), MW580369 (GAPDH), MW580370 (ITS), and MW580371 (TUB2). A nucleotide BLAST search revealed that ACT, CAL, GAPDH, ITS, and TUB2 sequences be 99% similar to accession numbers MN307380.1, MH155176.1, MK796226.1, MW580370.1, and JX010377.1, respectively of C. theobromicola. Maximum likelihood (ML) phylogenetic analysis was conducted based on a combined dataset of ACT, CAL, GAPDH, ITS, and TUB2 sequences using MEGA-X 10.1.8. The isolate ANUK97 was clustered with a representative strain C. theobromicola CBS124945 100% bootstrap support (Fig. 2). For the pathogenicity test, two-month-old cotton seedlings (n = 10) were inoculated with conidial suspensions (10⁶ spore/mL) of C. theobromicola obtained from 7-day-old PDA cultures at 25 °C by spray method. Seedlings treated with sterile distilled water served as controls. Inoculated and control cotton plants were incubated in the greenhouse at 25 °C under a 12-h photoperiod. After 7 days, necrotic lesions were observed on the artificially inoculated cotton plants, while control plants did not develop any disease symptoms. The pathogen was re-isolated from infected cotton leaves, but not from control plants to fulfill Koch’s postulates. To our knowledge, this is the first report of anthracnose of cotton caused by Colletotrichum theobromicola in Korea.


Plant Disease ◽  
2021 ◽  
Author(s):  
Svetlana Živković ◽  
Danijela Ristić ◽  
Stefan Stošić

Tomato (Solanum lycopersicum, L.) is one of the most important vegetable crop in Serbia, with a total production of 111,639 t in 2019 (Statistical Office of the Republic of Serbia). In July 2020, six tomatoes (cv. Balkan) with symptoms of fruit rot were collected from market in Belgrade, Serbia. The incidence of disease was about 2%, and the symptomatic samples were stored for 10 days after harvest. The initial symptoms on fruits were small circular, slightly sunken and water-soaked spots with white mycelia, that progressively expanded into larger grey lesions following the occurrence of sporulation. Isolations were conducted from one spot/fruit. Small pieces (2 to 3 mm2) from the margins of lesions were surface sterilized for 1 min in 1% NaOCl, washed twice with sterile distilled water, and cultivated on potato dextrose agar (PDA) at 25°C. The isolation frequency of Penicillium-like colonies was 100%. In total, six monosporic isolates were obtained and two isolates (SZ-20-6 and SZ-20-7) were selected as representative for morphological and molecular identification, and pathogenicity test. Morphological characteristics of both isolates were observed after growth on malt extract agar (MEA) for 7 days at 25ºC. On MEA, mycelia were white and colonies turned greyish-green with abundant sporulation. On the reverse sides colonies were pale yellow. The mean colony diameter on MEA for isolate SZ-20-6 was 25 ± 1.2 mm and 26 ± 1.0 mm for isolate SZ-20-7. The colony texture was velvety, without exudates and pigmentation. The conidiophores of both isolates were terverticillate, unbranched; phialides were flask shaped with a short neck, and conidia were smooth, greenish and subglobose to ellipsoidal. The conidial diameter for isolate SZ-20-6 was 3 to 4 × 2.5 to 3 µm, and for isolate SZ-20-7 was 3.5 to 4 × 2.5 to 3.5 µm (n =50). Based on these characteristics, isolates were identified as Penicillium olsonii (Pitt 1979). To confirm the morphological identification, genomic DNA was extracted from isolates (SZ-20-6 and SZ-20-7), and the rDNA ITS region and partial β-tubulin gene (BenA) were amplified using the primers ITS1/ITS4 (White et al. 1990) and Bt2a/Bt2b (Glass and Donaldson 1995), respectively. All sequences showed 99 to 100% similarity to P. olsonii and were deposited in GenBank (ITS, MW130235 and MW130236; BenA, MW145147 and MW145148). In multilocus phylogenetic analysis (ITS+BenA), isolates from this study clustered together with other P. olsonii sequences with 100% bootstrap support. To complete Koch's postulates, asymptomatic fruits of tomato cv. Balkan (five fruits per isolate) were superficially sterilized with 70% ethanol, wounded with a sterile needle and inoculated with 10 μl of a spore suspension (1 × 106 spores/ml). Five control fruits were inoculated with 10 μl of sterile distilled water. The experiment was repeated twice. After 7 days of incubation in a moisture chamber at 25°C, typical grey lesions developed on inoculated fruits. The control fruits remained symptomless. The isolates recovered from symptomatic fruits showed the same morphological features as the original isolates. P. olsonii was previously reported on tomato fruit only in Canada (Chatterton et al. 2012) and Pakistan (Anjum et al. 2018). To our knowledge, this is the first report of P. olsonii causing postharvest fruit rot on tomato in Serbia, and in Europe, as well. Therefore, it is essential to monitor spreading of P. olsonii on tomato and other crops in storages, and develop efficient disease management strategies. References: Anjum, N. et al. 2018. Plant Dis. 102:451. Chatterton, S., et al. 2012. Can. J. Plant Pathol. 34:524. Glass, N. L. and Donaldson, G. C. 1995. Appl. Environ. Microbiol. 61:1323. Pitt, J. I. 1979. The Genus Penicillium and its Teleomorophic States Eupenicillium and Talaromyces. Academic Press, London, U.K. Statistical Office of the Republic of Serbia. https://www.stat.gov.rs/en-US/ White, T. J., et al. 1990. Page 315 in: PCR Protocols: A Guide to Methods and Applications. Academic Press, San Diego, CA. Funding: This research was financed by the Ministry of Education, Science and Technical Development of the Republic of Serbia, grant 451-03-68/2020-14/200010.


Sign in / Sign up

Export Citation Format

Share Document