scholarly journals First report of Eucalyptus wilt caused by Chalaropsis thielavioides 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 ◽  
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 ◽  
2021 ◽  
Author(s):  
Lei Li ◽  
Yishuo Huang ◽  
Yanxia Shi ◽  
A LI CHAI ◽  
Xuewen Xie ◽  
...  

Coriander (Coriandrum sativum L.) or Chinese parsley is a culinary herb with multiple medicinal effects that are widely used in cooking and traditional medicine. From September to November 2019, symptoms were observed in 2-month-old coriander plants from coriander fields in Lanzhou and Wenzhou, China. The disease developed rapidly under cold and wet climatic conditions, and the infection rate was almost 80% in open coriander fields. Typical symptoms on leaves included small, water-soaked blotches and irregular brown spots surrounding haloes; as the disease progressed, the spots coalesced into necrotic areas. Symptomatic leaf tissue was surface sterilized, macerated in sterile distilled water, and cultured on nutrient agar plates at 28 °C for 48 h (Koike and Bull, 2006). After incubation, six bacterial colonies, which were individually isolated from collected samples from two different areas, were selected for further study. Colonies on NA plate were small, round, raised, white to cream-colored, and had smooth margins. All bacterial isolates were gram-negative, rod-shaped and nonfluorescent on King's B medium. The bacteria were positive for levan production, Tween 80 hydrolysis, and tobacco hypersensitivity but negative for oxidase, potato slice rot test, arginine dihydrolase, ice nucleation activity, indole production and H2S production. The suspension of representative isolate for inoculating of plants was obtained from single colony on King's B medium for 2-3 days at 28 °C. DNA was extracted from bacterial suspensions of YS2003200102 cultured in 20 ml of King’s B medium broth at 28 °C for 1 day. Extraction was performed with a TIANamp Bacterial DNA Kit (TIANGEN, China) according to the manufacturer’s recommendations. The pathogen was confirmed by amplification and sequencing of the glyceraldehyde-3-phosphate dehydrogenase A (gapA) gene, the citrate synthase (gltA) gene, the DNA gyrase B (gyrB) gene and the RNA polymerase sigma factor 70 (rpoD) gene using gapA-For/gapA-Rev, gltA-For/gltA-Rev, gyrB-For/gryB-Rev, rpoD-For/rpoD-Rev primers, respectively (Popović et al., 2019). The sequences of the PCR products were deposited in GenBank with accession numbers MZ681931 (gapA), MZ681932 (gltA), MZ681933 (gyrB), and MZ681934 (rpoD). Phylogenetic analysis of multiple genes (Xu and Miller, 2013) was conducted with the maximum likelihood method using MEGA7. The sequences of our isolates and ten published sequences of P. syringae pv. coriandricola were clustered into one clade with a 100% confidence level. To confirm the pathogenicity of isolate YS2003200102, 2-month-old healthy coriander plants were inoculated by spraying the leaves with a bacterial suspension (108 CFU ml−1) at 28 °C incubation temperature and 70% relative humidity condition, and sterile distilled water was applied as a negative control treatment (Cazorla et al. 2005). Three replicates were conducted for every isolate, and each replicate included 6 coriander plants. After twelve days, only the inoculated leaves with bacterial suspension showed bacterial leaf spot resembling those observed on naturally infected coriander leaves. Cultures re-isolated from symptomatic leaves showed the same morphological characteristics and molecular traits as those initially isolated from infected leaves in the field. This bacterium was previously reported causing leaf spot of coriander in India and Spain (Gupta et al. 2013; Cazorla et al. 2005). To our knowledge, this is the first report of P. syringae pv. coriandricola causing leaf spot disease on coriander in China. Studies are needed on strategies to manage P. syringae pv. coriandricola in crops, because its prevalence may cause yield loss on coriander in China.


Plant Disease ◽  
2020 ◽  
Author(s):  
Wen Li ◽  
Yue-qiu He ◽  
Tao Fu ◽  
Li Lin ◽  
Feng Liu ◽  
...  

Zinnia elegans (syn. Zinnia violacea), known as common zinnia, is one of the most spectacular ornamental plants in the family Asteraceae. Zinnia plants are widely cultivated in China for their impressive range in flower colours and profuse bloom over a long period. In April 2019, Zinnia plants grown in Ningbo Botanical Garden (29°56′57″N, 121°36′20″E) were found to have many circular necrotic lesions. In the early infection stage, the lesions appeared as small circular specks which developed later into large spots (15 to 32 mm diameter). Typical symptoms appeared to be grayish white centers with a chlorotic edges and disease incidence reached approximately 80% of plants in the affected field. Moreover, the growth of Zinnia plants was seriously affected by the disease. To identify the causative pathogen associated with the disease, 10 symptomatic leaves were collected from ten different Zinnia plants. Leaf tissues were cut from the lesion margins, surface sterilized with 75% ethanol for 30 seconds and rinsed three times in sterile distilled water. The leaf tissues were then dipped into 10% sodium hypochlorite for 2-3 minutes, washed three times in distilled water and dried on a sterile filter paper. After drying, the surface-sterilized leaf discs were transferred to potato dextrose agar (PDA) plates and incubated at 28°C for 2 to 3 days under the 12 h photoperiod. A total of ten pure fungal isolates were obtained and all the isolates displayed the same colony structure. Afterwards, three pure strains were randomly selected (F1, F3 and F5) for further study. The fungal colonies showed gray to brownish aerial mycelia with pink-colored masses of conidia. Conidia were one-celled, hyaline, cylindrical to subcylindrical, spindle-shaped with obtuse ends, measuring from 15.6 to 17.3 × 4.6 to 5.1 μm with both ends rounded. These morphological characteristics were consistent with the description of Colletotrichum gloeosporioides complex (Weir et al. 2012). The identity of a representative isolate, F3, was confirmed by a multilocus approach. Genomic DAN of isolate F3 was extracted and partial sequences of actin (ACT), chitin synthase (CHS), glyceraldehyde-3-phosphate dehydrogenase (GAPDH), ribosomal internal transcribed spacer (ITS), manganese-superoxide dismutase (SOD2) , glutamine synthatase (GS), beta-tubulin (TUB2) and calmodulin (CAL) were amplified and sequenced as previously described (Weir et al. 2012). These nucleotide sequences were deposited in GenBank (accession MN972436 to MN972440, and MT266559 to MT266561; all sequences in FASTA format are shown (Supplementary S1). BLAST analysis of ITS, ACT, CHS, GAPDH and GS sequences from the F3 isolate revealed similarity to C. gloeosporioides voucher strain ZH01 with 100%, 100%,99%, 99% and 99% identity, respectively. SOD, TUB2 and CAL sequences showed similarity to C. siamense with 100%, 100% and 100% identity, respectively. The phylogenetic trees were constructed by Maximum Likelihood method (ML) using JTT model implemented in the MEGA 7. Results inferred from the concatenated sequences (ACT, CHS, GAPDH, ITS, SOD, GS, TUB2 and CAL) placed the isolate F3 within the C. siamense cluster (Supplementary S2). To confirm pathogenicity of the fungus, Koch’s postulates were conducted by spraying 20 Zinnia plants (60-day-old) with a 1 × 106 conidia/ml suspension. Plants were maintained in the growth chamber at 25°C and 85% relative humidity. After 10 to 15 days, symptoms were observed on all inoculated leaves and resembled those observed in the field, whereas the control plants remained asymptomatic. Here, C. siamense was isolated only from the infected Zinnia leaves and identified by morphological and gene sequencing analyses. C. siamense has been reported in many crops in China (Yang et al. 2019; Chen et al. 2019; Wang et al. 2019). However, to our knowledge, this is the first report of anthracnose caused by C. siamense on Zinnia elegans in China. References Chen, X., Wang, T., Guo, H., Zhu, P. K., and Xu, L. 2019. First report of anthracnose of Camellia sasanqua caused by Colletotrichum siamense in China. Plant Dis. 103:1423-1423. Wang, Y., Qin, H. Y., Liu, Y. X., Fan, S. T., Sun, D., Yang, Y. M., Li, C. Y., and Ai, J. 2019. First report of anthracnose caused by Colletotrichum siamense on Actinidia arguta in China. Plant Dis. 103:372-373. Weir, B. S., Johnston, P. R., and Damm, U. 2012. The Colletotrichum gloeosporioides species complex. Stud. Mycol. 73: 115-180. Yang, S., Wang, H. X., Yi, Y. J., and Tan, L. L. 2019. First report that Colletotrichum siamense causes leaf spots on Camellia japonica in China. Plant Dis. 103:2127-2127.


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):  
Yue Tian ◽  
Yingying Zhang ◽  
Chaodong Qiu ◽  
Zhenyu Liu

Weigela florida (Bunge) A. DC. is a dense, rounded, deciduous shrub commonly planted in landscapes. It is also used in Chinese medicine to treat sore throat, erysipelas, cold, and fever (Zheng et al. 2019). In May 2019, leaf spots were observed on approximately 50% of W. florida plants grown in the Wisdom Plaza Park of Anhui Agricultural University in Hefei, Anhui Province, China. Leaf spots begun as small light brown and irregular lesions, enlarged, turned reddish brown, coalesced to form large blighted areas, and eventually covered the entire leaf surface. Five pieces of tissues were removed from the lesion margins of each diseased leaf (five leaves from five different plants), chopped into several 3-4 mm2 pieces, disinfected with 1.5% NaOCl for 2 min, rinsed 3 times with sterile distilled water for 1 min, plated onto Potato Dextrose Agar (PDA) medium containing 50 μg/ml of ampicillin and kanamycin, and incubated at 25°C with a 12-hour photoperiod for 5 days. One segment of the fungal growth from the growing edge of the colony was transferred onto a fresh PDA plate for purification and incubated under the same conditions for another 5 days. The colony morphology of one representative isolate (AAU0519) was characterized by a pale orange cushion in the center surrounded by irregular pink margin, diffusing red orange pigments into the PDA medium. Isolate AAU0519 was cultured on PDA medium for 7 days at 25°C in the dark to induce sporulation. The produced conidia were globose, subglobose to pyriform, golden brown to brown, and with a diameter of 7.7 - 23.8 μm. Both cultural and morphological characteristics suggested that isolate AAU0519 was an Epicoccum species, according to the description by Chen et al. 2017. Amplification and sequencing of the internal transcribed spacer (ITS), beta-tubulin, and 28S large subunit ribosomal RNA (LSU) gene fragments from the extracted genomic DNA of AAU0519 were performed using primer sets ITS1/ITS4 (White et al. 1990), Bt2a/Bt2b (Glass and Donaldson 1995), and LSU1Fd/LR5 (Crous et al. 2009; Vilgalys and Hester 1990), respectively. A phylogenetic tree was constructed by the maximum-likelihood method with 1,000 bootstrapping replications based on the concatenated ITS, beta-tubulin, and LSU sequences from isolate AAU0519 and representative strains of 22 species of the genus Epicoccum (Chen et al. 2017). Isolate AAU0519 clustered with ex-holotype CGMCC 3.18362 of Epicoccum layuense Qian Chen, Crous & L. Cai (Chen et al. 2017). All obtained sequences were deposited into GenBank under accession numbers MK983497 (ITS), MN328723 (beta-tubulin), and MN328724 (LSU). A pathogenicity test was conducted on leaves of five 3-year-old W. florida cultivar “Red Prince” planted in the field (five leaves for each treatment and control per plant) by spraying 30 ml of a spore suspension (106 spores/ml) of isolate AAU0519 as treatment or sterilized distilled water as control. Before the inoculation, the leaves were disinfected with 70% ethanol. After inoculation, the leaves were wrapped with a plastic bag to keep high relative humidity. The average air temperature was about 28°C during the period of pathogenicity test. The experiment was repeated once. Ten days after inoculation, the fungal-inoculated leaves developed light brown lesions resembling those of naturally infected leaves, control leaves did not develop any symptoms. E. layuense was recovered from leaf lesions and its identity was confirmed by morphological and sequence analyses as described above. To our knowledge, E. layuense has been previously reported as a pathogen of Perilla sp. (Chen et al. 2017), oat (Avena sativa) (Chen et al. 2019), and tea (Camellia sinensis) plants (Chen et al. 2020), but this is the first report of E. layuense causing leaf spot on W. florida in China. This pathogen could pose a threat to the ornamental value of W. florida plants. Thus, it is necessary to adopt effective management strategies against leaf spot on W. florida.


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 ◽  
2011 ◽  
Vol 95 (5) ◽  
pp. 616-616 ◽  
Author(s):  
J. Kim ◽  
O. Choi ◽  
J.-H. Kwon

Sweet persimmon (Diospyros kaki L.), a fruit tree in the Ebenaceae, is cultivated widely in Korea and Japan, the leading producers worldwide (2). Sweet persimmon fruit with flyspeck symptoms were collected from orchards in the Jinju area of Korea in November 2010. The fruit had fungal clusters of black, round to ovoid, sclerotium-like fungal bodies with no visible evidence of a mycelial mat. Orchard inspections revealed that disease incidence ranged from 10 to 20% in the surveyed area (approximately 10 ha) in 2010. Flyspeck symptoms were observed on immature and mature fruit. Sweet persimmon fruit peels with flyspeck symptoms were removed, dried, and individual speck lesions transferred to potato dextrose agar (PDA) and cultured at 22°C in the dark. Fungal isolates were obtained from flyspeck colonies on 10 sweet persimmon fruit harvested from each of three orchards. Fungal isolates that grew from the lesions were identified based on a previous description (1). To confirm identity of the causal fungus, the complete internal transcribed spacer (ITS) rDNA sequence of a representative isolate was amplified and sequenced using primers ITS1 and ITS4 (4). The resulting 552-bp sequence was deposited in GenBank (Accession No. HQ698923). Comparison with ITS rDNA sequences showed 100% similarity with a sequence of Zygophiala wisconsinensis Batzer & Crous (GenBank Accession No. AY598855), which infects apple. To fulfill Koch's postulates, mature, intact sweet persimmon fruit were surface sterilized with 70% ethanol and dried. Three fungal isolates from this study were grown on PDA for 1 month. A colonized agar disc (5 mm in diameter) of each isolate was cut from the advancing margin of a colony with a sterilized cork borer, transferred to a 1.5-ml Eppendorf tube, and ground into a suspension of mycelial fragments and conidia in a blender with 1 ml of sterile, distilled water. The inoculum of each isolate was applied by swabbing a sweet persimmon fruit with the suspension. Three sweet persimmon fruit were inoculated per isolate. Three fruit were inoculated similarly with sterile, distilled water as the control treatment. After 1 month of incubation in a moist chamber at 22°C, the same fungal fruiting symptoms were reproduced as observed in the orchards, and the fungus was reisolated from these symptoms, but not from the control fruit, which were asymptomatic. On the basis of morphological characteristics of the fungal colonies, ITS sequence, and pathogenicity to persimmon fruit, the fungus was identified as Z. wisconsinensis (1). Flyspeck is readily isolated from sweet persimmon fruit in Korea and other sweet persimmon growing regions (3). The exposure of fruit to unusual weather conditions in Korea in recent years, including drought, and low-temperature and low-light situations in late spring, which are favorable for flyspeck, might be associated with an increase in occurrence of flyspeck on sweet persimmon fruit in Korea. To our knowledge, this is the first report of Z. wisconsinensis causing flyspeck on sweet persimmon in Korea. References: (1) J. C. Batzer et al. Mycologia 100:246, 2008. (2) FAOSTAT Database. Retrieved from http://faostat.fao.org/ , 2008. (3) H. Nasu and H. Kunoh. Plant Dis. 71:361, 1987. (4) T. J. White et al. Page 315 in: PCR Protocols: A Guide to Methods and Applications. M. A. Innis et al., eds. Academic Press, Inc., New York, 1990.


Plant Disease ◽  
2021 ◽  
Author(s):  
Jiahao Lai ◽  
Guihong Xiong ◽  
Bing Liu ◽  
Weigang Kuang ◽  
Shuilin Song

Blueberry (Vaccinium virgatum), an economically important small fruit crop, is characterized by its highly nutritive compounds and high content and wide diversity of bioactive compounds (Miller et al. 2019). In September 2020, an unknown leaf blight disease was observed on Rabbiteye blueberry at the Agricultural Science and Technology Park of Jiangxi Agricultural University in Nanchang, China (28°45'51"N, 115°50'52"E). Disease surveys were conducted at that time, the results showed that disease incidence was 90% from a sampled population of 100 plants in the field, and this disease had not been found at other cultivation fields in Nanchang. Leaf blight disease on blueberry caused the leaves to shrivel and curl, or even fall off, which hindered floral bud development and subsequent yield potential. Symptoms of the disease initially appeared as irregular brown spots (1 to 7 mm in diameter) on the leaves, subsequently coalescing to form large irregular taupe lesions (4 to 15 mm in diameter) which became curly. As the disease progressed, irregular grey-brown and blighted lesion ran throughout the leaf lamina from leaf tip to entire leaf sheath and finally caused dieback and even shoot blight. To identify the causal agent, 15 small pieces (5 mm2) of symptomatic leaves were excised from the junction of diseased and healthy tissue, surface-sterilized in 75% ethanol solution for 30 sec and 0.1% mercuric chloride solution for 2 min, rinsed three times with sterile distilled water, and then incubated on potato dextrose agar (PDA) at 28°C for 5-7 days in darkness. Five fungal isolates showing similar morphological characteristics were obtained as pure cultures by single-spore isolation. All fungal colonies on PDA were white with sparse creeping hyphae. Pycnidia were spherical, light brown, and produced numerous conidia. Conidia were 10.60 to 20.12 × 1.98 to 3.11 µm (average 15.27 × 2.52 µm, n = 100), fusiform, sickle-shaped, light brown, without septa. Based on morphological characteristics, the fungal isolates were suspected to be Coniella castaneicola (Cui 2015). To further confirm the identity of this putative pathogen, two representative isolates LGZ2 and LGZ3 were selected for molecular identification. The internal transcribed spacer region (ITS) and large subunit (LSU) were amplified and sequenced using primers ITS1/ITS4 (Peever et al. 2004) and LROR/LR7 (Castlebury and Rossman 2002). The sequences of ITS region (GenBank accession nos. MW672530 and MW856809) showed 100% identity with accessions numbers KF564280 (576/576 bp), MW208111 (544/544 bp), MW208112 (544/544 bp) of C. castaneicola. LSU gene sequences (GenBank accession nos. MW856810 to 11) was 99.85% (1324/1326 bp, 1329/1331 bp) identical to the sequences of C. castaneicola (KY473971, KR232683 to 84). Pathogenicity was tested on three blueberry varieties (‘Rabbiteye’, ‘Double Peak’ and ‘Pink Lemonade’), and four healthy young leaves of a potted blueberry of each variety with and without injury were inoculated with 20 μl suspension of prepared spores (106 conidia/mL) derived from 7-day-old cultures of LGZ2, respectively. In addition, four leaves of each variety with and without injury were sprayed with sterile distilled water as a control, respectively. The experiment was repeated three times, and all plants were incubated in a growth chamber (a 12h light and 12h dark period, 25°C, RH greater than 80%). After 4 days, all the inoculated leaves started showing disease symptoms (large irregular grey-brown lesions) as those observed in the field and there was no difference in severity recorded between the blueberry varieties, whereas the control leaves showed no symptoms. The fungus was reisolated from the inoculated leaves and confirmed as C. castaneicola by morphological and molecular identification, fulfilling Koch’s postulates. To our knowledge, this is the first report of C. castaneicola causing leaf blight on blueberries in China. The discovery of this new disease and the identification of the pathogen will provide useful information for developing effective control strategies, reducing economic losses in blueberry production, and promoting the development of the blueberry industry.


Plant Disease ◽  
2012 ◽  
Vol 96 (10) ◽  
pp. 1580-1580
Author(s):  
J. H. Park ◽  
K. S. Han ◽  
J. Y. Kim ◽  
H. D. Shin

Sweet basil, Ocimum basilicum L., is a fragrant herb belonging to the family Lamiaceae. Originated in India 5,000 years ago, sweet basil plays a significant role in diverse cuisines across the world, especially in Asian and Italian cooking. In October 2008, hundreds of plants showing symptoms of leaf spot with nearly 100% incidence were found in polyethylene tunnels at an organic farm in Icheon, Korea. Leaf spots were circular to subcircular, water-soaked, dark brown with grayish center, and reached 10 mm or more in diameter. Diseased leaves defoliated prematurely. The damage purportedly due to this disease has reappeared every year with confirmation of the causal agent made again in 2011. A cercosporoid fungus was consistently associated with disease symptoms. Stromata were brown, consisting of brown cells, and 10 to 40 μm in width. Conidiophores were fasciculate (n = 2 to 10), olivaceous brown, paler upwards, straight to mildly curved, not geniculate in shorter ones or one to two times geniculate in longer ones, 40 to 200 μm long, occasionally reaching up to 350 μm long, 3.5 to 6 μm wide, and two- to six-septate. Conidia were hyaline, acicular to cylindric, straight in shorter ones, flexuous to curved in longer ones, truncate to obconically truncate at the base, three- to 16-septate, and 50 to 300 × 3.5 to 4.5 μm. Morphological characteristics of the fungus were consistent with the previous reports of Cercospora guatemalensis A.S. Mull. & Chupp (1,3). Voucher specimens were housed at Korea University herbarium (KUS). An isolate from KUS-F23757 was deposited in the Korean Agricultural Culture Collection (Accession No. KACC43980). Fungal DNA was extracted with DNeasy Plant Mini DNA Extraction Kits (Qiagen Inc., Valencia, CA). The complete internal transcribed spacer (ITS) region of rDNA was amplified with the primers ITS1/ITS4 and sequenced. The resulting sequence of 548 bp was deposited in GenBank (Accession No. JQ995781). This showed >99% similarity with sequences of many Cercospora species, indicating their close phylogenetic relationship. Isolate of KACC43980 was used in the pathogenicity tests. Hyphal suspensions were prepared by grinding 3-week-old colonies grown on PDA with distilled water using a mortar and pestle. Five plants were inoculated with hyphal suspensions and five plants were sprayed with sterile distilled water. The plants were covered with plastic bags to maintain a relative humidity of 100% for 24 h and then transferred to a 25 ± 2°C greenhouse with a 12-h photoperiod. Typical symptoms of necrotic spots appeared on the inoculated leaves 6 days after inoculation, and were identical to the ones observed in the field. C. guatemalensis was reisolated from symptomatic leaf tissues, confirming Koch's postulates. No symptoms were observed on control plants. Previously, the disease was reported in Malawi, India, China, and Japan (2,3), but not in Korea. To our knowledge, this is the first report of C. guatemalensis on sweet basil in Korea. Since farming of sweet basil has recently started on a commercial scale in Korea, the disease poses a serious threat to safe production of this herb, especially in organic farming. References: (1) C. Chupp. A Monograph of the Fungus Genus Cercospora. Ithaca, NY, 1953. (2) D. F. Farr and A. Y. Rossman. Fungal Databases. Systematic Mycology & Microbiology Laboratory, ARS, USDA. Retrieved from http://nt.ars-grin.gov/fungaldatabases/ , May 5, 2012. (3) J. Nishikawa et al. J. Gen. Plant Pathol. 68:46, 2002.


Sign in / Sign up

Export Citation Format

Share Document