scholarly journals First Report of Anthracnose on Kadsura coccinea Caused by Colletotrichum siamense in China

Plant Disease ◽  
2021 ◽  
Author(s):  
Guihe Jiang ◽  
Aiming Jiang ◽  
Chunli Fan ◽  
Ji-Guang Wei ◽  
Liyun Ren ◽  
...  

Kadsura coccinea (Lem.) A. C. Smith, an evergreen liana, is widely cultivated in China for its economic importance in traditional medicine. Many phytochemical studies on the stems and roots of K. coccinea have shown numerous biological activities, such as anti-tumor, anti-HIV, and anti-oxidant (Yang et al. 2020). In June 2019, an anthracnose on K. coccinea was observed in a plantation in Longan (23°03´N, 107°54´E), Guangxi province. Disease incidence was up to 30% in a plantation. Its symptoms began as small brown spots that expanded into nearly circular spots (Fig. 1A). To isolate pathogen, diseased leaves were collected. The leaves were sterilized with 75% ethanol for 15 s followed by 2% sodium hypochlorite for 90 s, then rinsed three times in sterilized distilled water, cut into 5 × 5 mm pieces, and placed into potato dextrose agar (PDA) plates. The plates were incubated in an incubator at 25°C in dark for 2–3 days. Fungal colonies with similar morphology of 27 isolates were isolated from the 30 infected tissues. Six representative isolates (YB1 to YB6) were selected to further study their characterization. Fungal colonies were grayish-white, orange-yellow conidial masses could be observed in colonies (Fig. 1C). The mature conidia were colorless and transparent, elliptical, and single-celled, 13.0–21.0 × 4.0–8.0 μm (average 16.92 × 5.92 µm; n =100) (Fig. 1B). The DNA sequences of ribosomal internal transcribed spacer region (ITS), glyceraldehyde-3-phosphate (GAPDH), calmodulin (CAL), actin (ACT), chitin synthase (CHS-1) and β-tubulin (TUB2) were amplified by PCR using the primer pairs ITS1/ITS4, GDF/GDR, CL1C/CL2C, ACT-512F/ACT-783R, CHS-79F/CHS-354R, and T1/Bt2b (Wang et al. 2020), respectively. Sequences were submitted to GenBank (Accession nos. MZ040489 to MZ040494 for ITS, MZ069043 to MZ069048 for GAPDH, MZ069049 to MZ069054 for CAL, MZ069055 to MZ069060 for ACT, MZ069061 to MZ069066 for CHS-1, and MZ069067 to MZ069072 for TUB2). These sequences were 98%–100% identical to that of reference isolates JX010278, JX010019, JX009709, GQ856775, GQ856730, and JX010410 of Colletotrichum siamense CBS 125378 ex-type recorded in GenBank. Phylogenetic analysis of combined ITS, GAPDH, CAL, ACT, CHS-1, TUB2 genes with 16 sequences obtained from GenBank using maximum likelihood method showed that the six isolates clustered with two reference isolates of Colletotrichum siamense as a distinct clade (Fig. 2). Based on morphological characteristics and phylogenetic analysis, six isolates were identified as C. siamense. Pathogenicity tests were performed on young, fully expanded leaves of 1-year seedlings. Every leaf was punctured at 6 points on the right half and 6 points on the left half using a sterile needle. A 10 μl conidial suspension (1×106 conidia/ml) was inoculated on each wound on the left-half leaf and a 10 μl sterile water was inoculated on each wound on the right-half leaf (control). Each treatment was repeated three times. Inoculated leaves were wrapped in plastic bags for 2 days and after removing the bags, plants were maintained in a growth chamber at 28°C, 80% relative humidity, and a 12-h photoperiod. Anthracnose spots formed 2 to 3 days after inoculation, whereas the control leaves remained symptomless. Morphological characters matched the descriptions of C. siamense. The pathogen was previously reported to cause anthracnose on Aloe vera (Azad et al. 2020), postharvest anthracnose in mango (Liu et al. 2017), pod rot in cacao (Serrato-Diaz et al. 2020). To our knowledge, this is the first report of anthracnose on K. coccinea caused by C. siamense in China.

Plant Disease ◽  
2012 ◽  
Vol 96 (8) ◽  
pp. 1227-1227 ◽  
Author(s):  
I. Jadrane ◽  
M. Kornievsky ◽  
D. E. Desjardin ◽  
Z.-H. He ◽  
L. Cai ◽  
...  

In October 2010, a Colletotrichum species was isolated from white Phalaenopsis flowers growing in a greenhouse in San Francisco, CA. This Phalaenopsis is a common commercial orchid hybrid generated mostly likely from Phalaenopsis amabilis and P. aphrodite. The white petals showed anthracnose-like lesions where necrotic tissue is surrounded by a ring of green tissue. The green halo tissues around the necrotic tissue contain functional chloroplasts. One-centimeter disks were cut around the necrotic sites and surface-sterilized with 95% ethanol and 0.6% sodium hypochlorite. The disks were placed on potato dextrose agar (PDA) medium to establish cultures. Pure cultures were obtained by subculturing hyphal tips onto fresh PDA plates. The generated colonies had white aerial mycelia and orange conidial mass. The color of the reverse colony varies between colorless and pale orange. Microscopic observations identified the conidia as cylindrical, straight, and rounded at both ends. In addition, the conidia were approximately 15.0 to 18.0 μm long and 5.0 to 6.5 μm in diameter. These observed morphological features suggested that these isolates possessed the same characteristics as previously described for Colletotrichum karstii, a species considered as part of the C. boninense species complex (1). Four putative independent Colletotrichum isolates were recovered (DED9596, DED9597, DED9598, and DED9599). To confirm the Colletotrichum isolates as the causative pathogen, healthy white Phalaenopsis flowers (five total) in a whole plant were sprayed with a conidial suspension (approximately 1.2 × 106 conidia/ml) of the isolates and incubated at 20°C and 100% relative humidity with cycles of 16 h light and 8 h of darkness. Approximately 1 ml of conidial suspension solution was used for each flower. The plants were watered regularly and flowers were sprayed with sterile double-distilled water daily. As negative controls, five flowers in a whole plant were sprayed with water. Fifteen to twenty days after inoculation, lesions started to form on the petals sprayed with the putative Colletotrichum isolates. All controls remained healthy. The Colletotrichum-inoculated flowers remained alive and did not die as a result of the infection. This same experiment was repeated and the same results were obtained. DNA was extracted from the necrotic regions of the petals infected by the pure cultures of the four isolates and used to sequence the 18S rRNA ITS (internal transcribed spacer) region. All four isolates gave identical ITS sequences. Analysis of the obtained representative sequences (GenBank Accession No. JQ277352) suggested that the isolated pathogen as C. karstii. Using the published ITS data for the C. boninense species complex (1), a phylogenetic tree was generated via the maximum likelihood method. This created tree places the isolates in the same group as C. karstii. This type of C. karstii infection in Phalaenopsis orchid petals was not documented in the U.S. before, although it has been reported in China and Thailand (2). To our knowledge, this is the first report of infection and green island formation caused by C. karstii on orchid flower in the United States. References: (1) Damm et al. Studies in Mycology 73:1, 2012. (2) Yang et al. Cryptogamie Mycologie 32:229, 2011.


Plant Disease ◽  
2014 ◽  
Vol 98 (5) ◽  
pp. 691-691 ◽  
Author(s):  
Y. H. Jeon ◽  
W. Cheon

Worldwide, Japanese yew (Taxus cuspidata Sieb. & Zucc.) is a popular garden tree, with large trees also being used for timber. In July 2012, leaf blight was observed on 10% of Japanese yew seedling leaves planted in a 500-m2 field in Andong, Gyeongsangbuk-do Province, South Korea. Typical symptoms included small, brown lesions that were first visible on the leaf margin, which enlarged and coalesced into the leaf becoming brown and blighted. To isolate potential pathogens from infected leaves, small sections of leaf tissue (5 to 10 mm2) were excised from lesion margins. Eight fungi were isolated from eight symptomatic trees, respectively. These fungi were hyphal tipped twice and transferred to potato dextrose agar (PDA) plates for incubation at 25°C. After 7 days, the fungi produced circular mats of white aerial mycelia. After 12 days, black acervuli containing slimy spore masses formed over the mycelial mats. Two representative isolates were further characterized. Their conidia were straight or slightly curved, fusiform to clavate, five-celled with constrictions at the septa, and 17.4 to 28.5 × 5.8 to 7.1 μm. Two to four 19.8- to 30.7-μm-long hyaline filamentous appendages (mostly three appendages) were attached to each apical cell, whereas one 3.7- to 7.1-μm-long hyaline appendage was attached to each basal cell, matching the description for Pestalotiopsis microspora (2). The pathogenicity of the two isolates was tested using 2-year-old plants (T. cuspidata var. nana Rehder; three plants per isolate) in 30-cm-diameter pots filled with soil under greenhouse conditions. The plants were inoculated by spraying the leaves with an atomizer with a conidial suspension (105 conidia/ml; ~50 ml on each plant) cultured for 10 days on PDA. As a control, three plants were inoculated with sterilized water. The plants were covered with plastic bags for 72 h to maintain high relative humidity (24 to 28°C). At 20 days after inoculation, small dark lesions enlarged into brown blight similar to that observed on naturally infected leaves. P. microspora was isolated from all inoculated plants, but not the controls. The fungus was confirmed by molecular analysis of the 5.8S subunit and flanking internal transcribed spaces (ITS1 and ITS2) of rDNA amplified from DNA extracted from single-spore cultures, and amplified with the ITS1/ITS4 primers and sequenced as previously described (4). Sequences were compared with other DNA sequences in GenBank using a BLASTN search. The P. microspora isolates were 99% homologous to other P. microspora (DQ456865, EU279435, FJ459951, and FJ459950). The morphological characteristics, pathogenicity, and molecular data assimilated in this study corresponded with the fungus P. microspora (2). This fungus has been previously reported as the causal agent of scab disease of Psidium guajava in Hawaii, the decline of Torreya taxifolia in Florida, and the leaf blight of Reineckea carnea in China (1,3). Therefore, this study presents the first report of P. microspora as a pathogen on T. cuspidata in Korea. The degree of pathogenicity of P. microspora to the Korean garden evergreen T. cuspidata requires quantification to determine its potential economic damage and to establish effective management practices. References: (1) D. F. Farr and A. Y. Rossman, Fungal Databases, Syst. Mycol. Microbiol. Lab. Retrieved from http://nt.ars-grin.gov/fungaldatabases/ (2) L. M. Keith et al. Plant Dis. 90:16, 2006. (3) S. S. N. Maharachchikumbura. Fungal Diversity 50:167, 2011. (4) T. J. White et al. PCR Protocols. Academic Press, San Diego, CA, 1990.


Plant Disease ◽  
2003 ◽  
Vol 87 (2) ◽  
pp. 203-203
Author(s):  
S. T. Koike ◽  
S. A. Tjosvold ◽  
J. Z. Groenewald ◽  
P. W. Crous

Bells-of-Ireland (Moluccella laevis) (Lamiaceae) is an annual plant that is field planted in coastal California (Santa Cruz County) for commercial cutflower production. In 2001, a new leaf spot disease was found in these commercially grown cutflowers. The disease was most serious in the winter-grown crops in 2001 and 2002, with a few plantings having as much as 100% disease incidence. All other plantings that were surveyed during this time had at least 50% disease. Initial symptoms consisted of gray-green leaf spots. Spots were generally oval in shape, often delimited by the major leaf veins, and later turned tan. Lesions were apparent on both adaxial and abaxial sides of the leaves. A cercosporoid fungus having fasciculate conidiophores, which formed primarily on the abaxial leaf surface, was consistently associated with the spots. Based on morphology and its host, this fungus was initially considered to be Cercospora molucellae Bremer & Petr., which was previously reported on leaves of M. laevis in Turkey (1). However, sequence data obtained from the internal transcribed spacer region (ITS1, ITS2) and the 5.8S gene (STE-U 5110, 5111; GenBank Accession Nos. AY156918 and AY156919) indicated there were no base pair differences between the bells-of-Ireland isolates from California, our own reference isolates of C. apii, as well as GenBank sequences deposited as C. apii. Based on these data, the fungus was subsequently identified as C. apii sensu lato. Pathogenicity was confirmed by spraying a conidial suspension (1.0 × 105 conidia/ml) on leaves of potted bells-of-Ireland plants, incubating the plants in a dew chamber for 24 h, and maintaining them in a greenhouse (23 to 25°C). After 2 weeks, all inoculated plants developed leaf spots that were identical to those observed in the field. C. apii was again associated with all leaf spots. Control plants, which were treated with water, did not develop any symptoms. The test was repeated and the results were similar. To our knowledge this is the first report of C. apii as a pathogen of bells-of-Ireland in California. Reference: (1) C. Chupp. A Monograph of the Fungus Genus Cercospora. Cornell University Press, Ithaca, New York, 1954.


Plant Disease ◽  
2021 ◽  
Author(s):  
Zhou Zhang ◽  
Zheng Bing Zhang ◽  
Yuan Tai Huang ◽  
FeiXiang Wang ◽  
Wei Hua Hu ◽  
...  

Peach [Prunus persica (L.) Batsch] is an important deciduous fruit tree in the family Rosaceae and is a widely grown fruit in China (Verde et al., 2013). In July and August 2018, a fruit rot disease was observed in a few peach orchards in Zhuzhou city, the Hunan Province of China. Approximately 30% of the fruit in more than 400 trees was affected. Symptoms displayed were brown necrotic spots that expanded, coalesced, and lead to fruit being rotten. Symptomatic tissues excised from the margins of lesions were surface sterilized in 70% ethanol for 10 s, 0.1% HgCl2 for 2 min, rinsed with sterile distilled water three times, and incubated on potato dextrose agar (PDA) at 26°C in the dark. Fungal colonies with similar morphology developed, and eight fungal colonies were isolated for further identification. Colonies grown on PDA were grayish-white with white aerial mycelium. After an incubation period of approximately 3 weeks, pycnidia developed and produced α-conidia and β-conidia. The α-conidia were one-celled, hyaline, fusiform, and ranged in size from 6.0 to 8.4 × 2.1 to 3.1 μm, whereas the β-conidia were filiform, hamate, and 15.0 to 27.0 × 0.8 to 1.6 μm. For molecular identification, total genomic DNA was extracted from the mycelium of a representative isolate HT-1 and the internal transcribed spacer region (ITS), β-tubulin gene (TUB), translation elongation factor 1-α gene (TEF1), calmodulin (CAL), and histone H3 gene (HIS) were amplified and sequenced (Meng et al. 2018). The ITS, TUB, TEF1, CAL and HIS sequences (GenBank accession nos. MT740484, MT749776, MT749778, MT749777, and MT749779, respectively) were obtained and in analysis by BLAST against sequences in NCBI GenBank, showed 99.37 to 100% identity with D. hongkongensis or D. lithocarpus (the synonym of D. hongkongensis) (Gao et al., 2016) (GenBank accession nos. MG832540.1 for ITS, LT601561.1 for TUB, KJ490551.1 for HIS, KY433566.1 for TEF1, and MK442962.1 for CAL). Pathogenicity tests were performed on peach fruits by inoculation of mycelial plugs and conidial suspensions. In one set, 0.5 mm diameter mycelial discs, which were obtained from an actively growing representative isolate of the fungus on PDA, were placed individually on the surface of each fruit. Sterile agar plugs were used as controls. In another set, each of the fruits was inoculated by application of 1 ml conidial suspension (105 conidia/ml) by a spray bottle. Control assays were carried out with sterile distilled water. All treatments were maintained in humid chambers at 26°C with a 12-h photoperiod. The inoculation tests were conducted twice, with each one having three fruits as replications. Six days post-inoculation, symptoms of fruit rot were observed on inoculated fruits, whereas no symptoms developed on fruits treated with agar plugs and sterile water. The fungus was re-isolated and identified to be D. hongkongensis by morphological and molecular methods, thus fulfilling Koch’s Postulates. This fungus has been reported to cause fruit rot on kiwifruit (Li et al. 2016) and is also known to cause peach tree dieback in China (Dissanayake et al. 2017). However, to our knowledge, this is the first report of D. hongkongensis causing peach fruit rot disease in China. The identification of the pathogen will provide important information for growers to manage this disease.


Plant Disease ◽  
2014 ◽  
Vol 98 (9) ◽  
pp. 1281-1281 ◽  
Author(s):  
S. Mahadevakumar ◽  
Vandana Yadav ◽  
G. S. Tejaswini ◽  
S. N. Sandeep ◽  
G. R. Janardhana

Lemon (Citrus lemon (L.) Burm. f.) is an important fruit crop cultivated worldwide, and is grown practically in every state in India (3). During a survey conducted in 2013, a few small trees in a lemon orchard near Mysore city (Karnataka) (12°19.629′ N, 76°31.892′ E) were found affected by dieback disease. Approximately 10 to 20% of trees were affected as young shoots and branches showed progressive death from the apical region downward. Different samples were collected and diagnosed via morphological methods. The fungus was consistently isolated from the infected branches when they were surface sanitized with 1.5% NaOCl and plated on potato dextrose agar (PDA). Plates were incubated at 26 ± 2°C for 7 days at 12/12 h alternating light and dark period. Fungal colonies were whitish with pale brown stripes having an uneven margin and pycnidia were fully embedded in the culture plate. No sexual state was observed. Pycnidia were globose, dark, 158 to 320 μm in diameter, and scattered throughout the mycelial growth. Both alpha and beta conidia were present within pycnidia. Alpha conidia were single celled (5.3 to 8.7 × 2.28 to 3.96 μm) (n = 50), bigittulate, hyaline, with one end blunt and other truncated. Beta conidia (24.8 to 29.49 × 0.9 to 1.4 μm) (n = 50) were single celled, filiform, with one end rounded and the other acute and curved. Based on the morphological and cultural features, the fungal pathogen was identified as Phomopsis citri H.S. Fawc. Pathogenicity test was conducted on nine healthy 2-year-old lemon plants via foliar application of a conidial suspension (3 × 106); plants were covered with polythene bags for 6 days and maintained in the greenhouse. Sterile distilled water inoculated plants (in triplicate) served as controls and were symptomless. Development of dieback symptoms was observed after 25 days post inoculation and the fungal pathogen was re-isolated from the inoculated lemon trees. The internal transcribed spacer region (ITS) of the isolated fungal genomic DNA was amplified using universal-primer pair ITS1/ITS4 and sequenced to confirm the species-level diagnosis (4). The sequence data of the 558-bp amplicon was deposited in GenBank (Accession No. KJ477016.1) and nBLAST search showed 99% homology with Diaporthe citri (teleomorph) strain 199.39 (KC343051.1). P. citri is known for its association with melanose disease of citrus in India, the United States, and abroad. P. citri also causes stem end rot of citrus, which leads to yield loss and reduction in fruit quality (1,2). Dieback disease is of serious concern for lemon growers as it affects the overall productivity level of the tree. To the best of our knowledge, this is the first report of P. citri causing dieback of lemon in India. References: (1) I. H. Fischer et al. Sci. Agric. (Piracicaba). 66:210, 2009. (2) S. N. Mondal et al. Plant Dis. 91:387, 2007. (3) S. P. Raychaudhuri. Proc. Int. Soc. Citriculture 1:461, 1981. (4) T. J. White et al. Page 315 in: PCR Protocols: A Guide to Methods and Applications. Academic Press, San Diego, CA, 1990.


Plant Disease ◽  
2021 ◽  
Author(s):  
Jun Guo ◽  
Jin Chen ◽  
Zhao Hu ◽  
Jie Zhong ◽  
Jun Zi Zhu

Cardamine hupingshanensis is a selenium (Se) and cadmium (Cd) hyperaccumulator plant distributed in wetlands along the Wuling Mountains of China (Zhou et al. 2018). In March of 2020, a disease with symptoms similar to gray mold was observed on leaves of C. hupingshanensis in a nursery located in Changsha, Hunan Province, China. Almost 40% of the C. hupingshanensis (200 plants) were infected. Initially, small spots were scattered across the leaf surface or margin. As disease progressed, small spots enlarged to dark brown lesions, with green-gray, conidia containing mold layer under humid conditions. Small leaf pieces were cut from the lesion margins and were sterilized with 70% ethanol for 10 s, 2% NaOCl for 2 min, rinsed with sterilized distilled water for three times, and then placed on potato dextrose agar (PDA) medium at 22°C in the dark. Seven similar colonies were consistently isolated from seven samples and further purified by single-spore isolation. Strains cultured on PDA were initially white, forming gray-white aerial mycelia, then turned gray and produced sclerotia after incubation for 2 weeks, which were brown to blackish, irregular, 0.8 to 3.0 × 1.2 to 3.5 mm (n=50). Conidia were unicellular, globose or oval, colourless, 7.5 to 12.0 × 5.5 to 8.3 μm (n=50). Conidiophores arose singly or in group, straight or flexuous, septate, brownish to light brown, with enlarged basal cells, 12.5 to 22.1 × 120.7 to 310.3 μm. Based on their morphological characteristics in culture, the isolates were putatively identified as Botrytis cinerea (Ellis 1971). Genomic DNA of four representative isolates, HNSMJ-1 to HNSMJ-4, were extracted by CTAB method. The internal transcribed spacer region (ITS), glyceraldehyde-3-phosphate dehydrogenase gene (G3PDH), heat-shock protein 60 gene (HSP60), ATP-dependent RNA helicaseDBP7 gene (MS547) and DNA-dependent RNA polymerase subunit II gene (RPB2) were amplified and sequenced using the primers described previously (Aktaruzzaman et al. 2018) (MW820311, MW831620, MW831628, MW831623 and MW831629 for HNSMJ-1; MW314722, MW316616, MW316617, MW316618 and MW316619 for HNSMJ-2; MW820519, MW831621, MW831627, MW831624 and MW831631 for HNSMJ-3; MW820601, MW831622, MW831626, MW831625 and MW831630 for HNSMJ-4). BLAST searches showed 99.43 to 99.90% identity to the corresponding sequences of B. cinerea strains, such as HJ-5 (MF426032.1, MN448500.1, MK791187.1, MH727700.1 and KX867998.1). A combined phylogenetic tree using the ITS, G3PDH, HSP60 and RPB2 sequences was constructed by neighbor-joining method in MEGA 6. It revealed that HNSMJ-1 to HNSMJ-4 clustered in the B. cinerea clade. Pathogenicity tests were performed on healthy pot-grown C. hupingshanensis plants. Leaves were surface-sterilized and sprayed with conidial suspension (106 conidia/ mL), with sterile water served as controls. All plants were kept in growth chamber with 85% humidity at 25℃ following a 16 h day-8 h night cycle. The experiment was repeated twice, with each three replications. After 4 to 7 days, symptoms similar to those observed in the field developed on the inoculated leaves, whereas controls remained healthy. The pathogen was reisolated from symptomatic tissues and identified using molecular methods, confirming Koch’s postulates. B. cinerea has already been reported from China on C. lyrate (Zhang 2006), a different species of C. hupingshanensis. To the best of our knowledge, this is the first report of B. cinerea causing gray mold on C. hupingshanensis in China and worldwide. Based on the widespread damage in the nursery, appropriate control strategies should be adopted. This study provides a basis for studying the epidemic and management of the disease.


Plant Disease ◽  
2021 ◽  
Author(s):  
Ali Chai ◽  
Qian Zhao ◽  
Yanxia Shi ◽  
Xuewen Xie ◽  
Lei Li ◽  
...  

Okra [Abelmoschus esculentus (L.) Moench], which belongs to the family Malvaceae, is widely grown in the tropics, sub-tropics and warmer areas of the temperate zones for its immature seed pods which are consumed as a vegetable. In China, okra pods are consumed as not only vegetables but also as a traditional medicine to cure dental diseases and gastric ulcers. During September 2018 to June 2019, extensive spots on okra leaves were observed in several commercial fields (approximately 2.0 hectares), with disease incidence of approximately 25%~50% in the Yanqing District (115°98′E, 40°46′N) of Beijing, China. Symptoms of the disease initially appeared as small pale brown spots with yellow haloes. As the disease progressed, some spots gradually coalesced, forming larger irregular dark brown lesions. The centers of the lesions became grayish white. A total of 13 small fragments (3 to 5 mm) excised from the lesion margins were sterilized in 1% sodium hypochlorite (NaClO) for 1 min, followed by three washes with sterile distilled water, and then placed on potato dextrose agar (PDA) and incubated at 25°C in the dark for 5 days. In total, 21 cultures were obtained and purified by single-spore subcultures on PDA for morphological identification. The colonies on PDA were whitish to gray, with cottony aerial mycelium. Conidiophores were fasciculate, olivaceous brown, straight or geniculate, uniform in width, multiseptate, and ranged from 286/span> to 711 μm (avg. = 578 μm, n = 50). Conidia were hyaline, slightly curved or straight, needle shaped, truncate at the base, and terminal at the tip, 3–17-septate, and measuring 52 to 231 μm (avg. = 182 μm, n = 50). The morphological features were consistent with Cercospora cf. flagellaris Ellis & G. Martin (Groenewald et al. 2013). Pathogenicity tests were conducted on potted okra plants cv. ‘Jiayuan’. Twenty four healthy okra plants at the true leaf stage were sprayed with conidial suspensions (1 × 106 conidia/mL), incubated at a glass cabinet maintained at 25°C and 90% relative humidity (RH). To each leaf approximately 10 mL of conidial suspension was applied. Plants sprayed with water were used as controls. Seven days later, dark brown spot, which were identical to those observed in the fields, were observed on inoculated leaves, whereas the control plants remained healthy. C. cf. flagellaris was reisolated from symptomatic leaves, confirming Koch’s Postulates. Genomic DNA was extracted from fungal mycelium using the Plant Genomic DNA Kit (Tiangen Biotech Co. Ltd., Beijing, China). The nuclear ribosomal internal transcribed spacer region (ITS), and portions of the actin (ACT), histone H3 (HIS3), and translation elongation factor 1-α (TEF1) genes were amplified using primers ITS1/ITS4 (Groenewald et al. 2013), ACT-512F/ACT-783R (Carbone & Kohn 1999), CYLH3F/CYLH3R (Crous et al. 2006), and EF1-728F/EF1-986R (Carbone & Kohn 1999). The resulting 542 bp ITS, 226 bp ACT, 410 bp HIS3 and 306 bp TEF1 sequences of isolate QK14091813 were deposited in GeneBank (Accession nos. MT949700, MT949701, MT949702 and MT949703, respectively). The ITS, ACT, HIS3 and TEF1 sequences shared 99.42% to 100% identities to previously published sequences of C. cf. flagellaris (Accession nos. MN633275 for ITS, MF680960 for ACT, MK991295 for HIS3, and MK991292.1 for TEF1, respectively). Multi-locus phylogenetic analyses (ITS, ACT, HIS3, and TEF1) were performed by neighbor-joining method using MEGA 7.0. The resulting trees showed that C. cf. flagellaris isolate QK14091813 (this study) nested within the clade that includes other isolates of C. cf. flagellaris with a 99% confidence level. To our knowledge, this is the first report of C. cf. flagellaris causing leaf spot on okra (Farr and Rossman 2020). The pathogen has a worldwide distribution and an unusually broad host range, which can be of great significance, and the plant protection policy of priority to prevention and synthetical prevention should be followed.


Plant Disease ◽  
2022 ◽  
Author(s):  
Xiang Xie ◽  
Shiqiang Zhang ◽  
Qingjie Yu ◽  
Xinye Li ◽  
Yongsheng Liu ◽  
...  

Camellia oleifera, a major tree species for producing edible oil, is originated in China. Its oil is also called ‘‘eastern olive oil’’ with high economic value due to richness in a variety of healthy fatty acids (Lin et al. 218). However, leaves are susceptible to leaf spot disease (Zhu et al. 2014). In May 2021, we found circular to irregular reddish-brown lesions, 4-11 mm in diameter, near the leaf veins or leaf edges on 30%-50% leaves of 1/3 oil tea trees in a garden of Hefei City, Anhui Province, China (East longitude 117.27, North latitude 31.86) (Figure S1 A). To isolate the causal agents, symptomatic leaves were cut from the junction of diseased and healthy tissues (5X5 mm) and treated with 70 % alcohol for 30 secs and 1 % NaClO for 5 min, and subsequently inoculated onto PDA medium for culture. After 3 days, hyphal tips were transferred to PDA. Eventually, five isolates were obtained. Then the isolates were cultured on PDA at 25°C for 7 days and the mycelia appeared yellow with a white edge and secreted a large amount of orange-red material to the PDA (Figure S1 B and C). Twenty days later, the mycelium appeared reddish-brown, and sub-circular (3-10 mm) raised white or yellow mycelium was commonly seen on the Petri dish, and black particles were occasionally seen. Meanwhile, the colonies on the PDA produced abundant conidia. Microscopy revealed that conidia were globular to pyriform, dark, verrucose, and multicellular with 14.2 to 25.3 μm (=19.34 μm, n = 30) diameter (Figure S1 D). The morphological characteristics of mycelial and conidia from these isolates are similar to that of Epicoccum layuense (Chen et al.2020). To further determine the species classification of the isolates, DNA was extracted from 7-day-old mycelia cultures and the PCR-amplified fragments were sequenced for internal transcribed spacer (ITS), beta-tubulin and 28S large subunit ribosomal RNA (LSU) gene regions ITS1/ITS4, Bt2a/Bt2b and LR0R/LR5, followed by sequencing and molecular phylogenetic analysis of the sequences analysis (White et al. 1990; Glass and Donaldson 1995; Vilgalys and Hester 1990). Sequence analysis revealed that ITS, beta-tubulin, and LSU divided these isolates into two groups. The isolates AAU-NCY1 and AAU-NCY2, representing the first group (AAU-NCY1 and AAU-NCY5) and the second group (AAU-NCY2, AAU-NCY3 and AAU-NCY4), respectively, were used for further studies. Based on BLASTn analysis, the ITS sequences of AAU-NCY1 (MZ477250) and AAU-NCY2 (MZ477251) showed 100 and 99.6% identity with E. layuense accessions MN396393 and KY742108, respectively. And, the beta-tubulin sequences (MZ552310; MZ552311) showed 99.03 and 99.35% identity with E. layuense accessions MN397247 and MN397248, respectively. Consistently, their LSU (MZ477254; MZ477255) showed 99.88 and 99.77% identity with E. layuense accessions MN328724 and MN396395, respectively. Phylogenetic trees were built by maximum likelihood method (1,000 replicates) using MEGA v.6.0 based on the concatenated sequences of ITS, beta-tubulin and LSU (Figure S2). Phylogenetic tree analysis confirmed that AAU-NCY1 and AAU-NCY2 are closely clustered with E. layuense stains (Figure S2). To test the pathogenicity, conidial suspension of AAU-NCY2 (106 spores/mL) was prepared and sterile water was used as the control. Twelve healthy leaves (six for each treatment) on C. oleifera tree were punched with sterile needle (0.8-1mm), the sterile water or spore suspension was added dropwise at the pinhole respectively (Figure S1 E and F). The experiment was repeated three times. By ten-day post inoculation, the leaves infected by the conidia gradually developed reddish-brown necrotic spots that were similar to those observed in the garden, while the control leaves remained asymptomatic (Figure S1 G and H). DNA sequences derived from the strain re-isolated from the infected leaves was identical to that of the original strain. E. layuense has been reported to cause leaf spot on C. sinensis (Chen et al. 2020), and similar pathogenic phenotypes were reported on Weigela florida (Tian et al. 2021) and Prunus x yedoensis Matsumura in Korea ( Han et al. 2021). To our knowledge, this is the first report of E. layuense causing leaf spot on C. oleifera in Hefei, China.


Plant Disease ◽  
2021 ◽  
Author(s):  
Xiujing Hong ◽  
Shijia Chen ◽  
linchao Wang ◽  
Bo Liu ◽  
Yuruo Yang ◽  
...  

Akebia trifoliata, a recently domesticated horticultural crop, produces delicious fruits containing multiple nutritional metabolites and has been widely used as medicinal herb in China. In June 2020, symptoms of dried-shrink disease were first observed on fruits of A. trifoliata grown in Zhangjiajie, China (110.2°E, 29.4°N) with an incidence about 10%. The infected fruits were shrunken, colored in dark brown, and withered to death (Figure S1A, B). The symptomatic fruits tissues (6 × 6 mm) were excised from three individual plants, surface-disinfested in 1% NaOCl for 30s and 70% ethanol solution for 45s, washed, dried, and plated on potato dextrose agar (PDA) containing 50 mg/L streptomycin sulfate in the dark, and incubated at 25℃ for 3 days. Subsequently, hyphal tips were transferred to PDA to obtain pure cultures. After 7 days, five pure cultures were obtained, including two identical to previously reported Colletotrichum gloeosporioides causing leaf anthracnose in A. trifoliata (Pan et al. 2020) and three unknown isolates (ZJJ-C1-1, ZJJ-C1-2, and ZJJ-C1-3). The mycelia of ZJJ-C1-1, ZJJ-C1-2 and ZJJ-C1-3 were white, and formed colonies of approximate 70 mm (diameter) in size at 25℃ after 7 days on potato sucrose agar (PSA) plates (Figure S1C). After 25 days, conidia were formed, solitary, globose, black, shiny, smooth, and 16-21 μm in size (average diameter = 18.22 ± 1.00 μm, n = 20) (Figure S1D). These morphological characteristics were similar to those of N. sphaerica previously reported (Li et al. 2018). To identify species of ZJJ-C1-1, ZJJ-C1-2 and ZJJ-C1-3, the internal transcribed spacer (ITS) region, β-tubulin (TUB2), and the translation elongation factor 1-alpha (TEF1-α) were amplified using primer pairs including ITS1/ITS4 (Vilgalys and Hester 1990), Bt-2a/Bt-2b (Glass and Donaldson 1995), and EF1-728F/EF-2 (Zhou et al. 2015), respectively. Multiple sequence analyses showed no nucleotide difference was detected among genes tested except ITS that placed three isolates into two groups (Figure S2). BLAST analyses determined that ZJJ-C1-1, ZJJ-C1-2 and ZJJ-C1-3 had 99.73% to N. sphaerica strains LC2705 (KY019479), 100% to LC7294 (KY019397), and 99.79-100% to LC7294 (KX985932) or LC7294 (KX985932) based on sequences of TUB2 (MW252168, MW269660, MW269661), TEF-1α (MW252169, MW269662, MW269663), and ITS (MW250235, MW250236, MW192897), respectively. These indicated three isolates belong to the same species of N. sphaerica. Based on a combined dataset of ITS, TUB2 and TEF-1α sequences, a phylogenetic tree was constructed using Maximum likelihood method through IQ-TREE (Minh et al. 2020) and confirmed that three isolates were N. sphaerica (Figure S2). Further, pathogenicity tests were performed. Briefly, healthy unwounded fruits were surface-disinfected in 0.1% NaOCl for 30s, washed, dried and needling-wounded. Then, three fruits were inoculated with 10 μl of conidial suspension (1 × 106 conidia/ml) derived from three individual isolates, with another three fruits sprayed with 10 μl sterilized water as control. The treated fruits were incubated at 25℃ in 90% humidity. After 15 days, all the three fruits inoculated with conidia displayed typical dried-shrink symptoms as those observed in the farm field (Figure S1E). The decayed tissues with mycelium and spores could be observed on the skin or vertical split of the infected fruits after 15 days’ inoculation (Figure S1F-H). Comparably, in the three control fruits, there were no dried-shrink-related symptoms displayed. The experiment was repeated twice. The re-isolated pathogens were identical to N. sphaerica determined by sequencing the ITS, TUB2 and TEF-1α. Previous reports showed N. sphaerica could cause postharvest rot disease in kiwifruits (Li et al. 2018). To our knowledge, this is the first report of N. sphaerica causing fruits dried-shrink disease in A. trifoliata in China.


Plant Disease ◽  
2021 ◽  
Author(s):  
Haixia Ding ◽  
Wan Peng Dong ◽  
Wei Di Mo ◽  
Lijuan Peng ◽  
Zuo-Yi Liu

Chinese rose (Rosa chinensis Jacq.) is cultivated for edible flowers in southwestern China (Zhang et al. 2014). In March 2020, a leaf spot disease was observed on about 3-5% leaves of Chinese rose cultivar ‘Mohong’ in Guizhou Botanical Garden (26°37' 45'' N, 106°43' 10'' E), Guiyang, Guizhou province, China. The symptomatic plants displayed circular, dark brown lesions with black conidiomata in grey centers on leaves, and leaf samples were collected. After surface sterilization (0.5 min in 75% ethanol and 2 min in 3% NaOCl, washed 3 times with sterilized distilled water) (Fang 2007), small pieces of symptomatic leaf tissue (0.3 × 0.3 cm) were plated on potato dextrose agar (PDA) and incubated at 28oC for about 7 days. Two single-spore isolates, GZUMH01 and GZUMH02, were obtained, which were identical in morphology and molecular analysis. Therefore, the representative isolate GZUMH01 was used for further study. The pathogenicity of GZUMH01 was tested through a pot assay. Ten healthy plants were scratched with a sterilized needle on the leaves. Plants were inoculated by spraying a spore suspension (106 spores ml-1) onto leaves until runoff, and the control leaves sprayed with sterile water. The plants were maintained at 25°C with high relative humidity (90 to 95%) in a growth chamber. The pathogenicity test was carried out three times using the method described in Fang (2007). The symptoms developed on all inoculated leaves but not on the control leaves. The lesions were first visible 48 h after inoculation, and typical lesions similar to those observed on field plants after 7 days. The same fungus was re-isolated from the infected leaves but not from the non-inoculated leaves, fulfilling Koch’s postulates. Fungal colonies on PDA were villiform and greyish. The conidia were abundant, oval-ellipsoid, aseptate, 15.8 (13.7 to 18.8) × 5.7 (4.3 to 6.8) µm. The fungal colonies, hyphae, and conidia were consistent with the descriptions of Colletotrichum boninense Moriwaki, Toy. Sato & Tsukib. (Damm et al. 2012; Moriwaki et al. 2003). The pathogen was confirmed to be C. boninense by amplification and sequencing of the internal transcribed spacer region (ITS), the glyceraldehyde-3-phosphate dehydrogenase (GADPH), actin (ACT), and chitin synthase 1 (CHS-1) genes using primers ITS1/ITS4, GDF1/GDR1, ACT512F/ACT783R, and CHS-79F/CHS-345R, respectively (Damm et al. 2012; Moriwaki et al. 2003). The sequences of the PCR products were deposited in GenBank with accession numbers MT845879 (ITS), MT861006 (GADPH), MT861007 (ACT), and MT861008 (CHS-1). BLAST searches of the obtained sequences of the ITS, GADPH, ACT, and CHS-1 genes revealed 100% (554/554 nucleotides), 100% (245/245 nucleotides), 97.43% (265/272 nucleotides), and 99.64% (279/280 nucleotides) homology with those of C. boninense in GenBank (JQ005160, JQ005247, JQ005508, and JQ005334, respectively). Phylogenetic analysis (MEGA 6.0) using the maximum likelihood method placed the isolate GZUMH01 in a well-supported cluster with C. boninense. The pathogen was thus identified as C. boninense based on its morphological and molecular characteristics. To our knowledge, this is the first report of the anthracnose disease on R. chinensis caused by C. boninense in the world.


Sign in / Sign up

Export Citation Format

Share Document