scholarly journals First Report of Leaf Spot in Farfugium japonicum Caused by Alternaria tenuissima in Korea

Plant Disease ◽  
2013 ◽  
Vol 97 (10) ◽  
pp. 1382-1382 ◽  
Author(s):  
J. H. Lee ◽  
D. S. Kim ◽  
H. J. Cho ◽  
G.-H. Gang ◽  
Y.-S. Kwak

Farfugium japonicum (L.) Kitam (common name: Leopard plant) is known as a medical herb and belongs to family Asteraceae (1). In June 2012, a leaf spot disease was observed on the leaf surface of F. japonicum at a forest research plot Jinju, Gyeongnam province, Korea. More than 95% of F. japonicum plants were infected and leaf spot symptom appeared in the regions under our investigation. Light brownish symptoms initially developed and the spot size gradually increased and turned dark brown with an irregular shape as the disease progressed (spot size 1 to 10 mm in diameter). At the late stage of disease, spots became hollow and completely dehydrated. The infected leaves were easily crumbled, possibly due to dryness. To isolate the causal agent, the infected leaves were surface disinfected and pieces of leaves were placed on water agar (WA). Nine isolates were isolated from 10 pieces of the infected leaves. Fungi mycelia from the WA were transferred on potato dextrose agar (PDA) and incubated at 28°C for 7 days. The colonies were purple navy to black and conidia spores developed on the media. The morphological characteristics of spores were multi-septate, dark brown, pyriform, and 6.7 to 12.8 × 22.2 to 38.4 μm. The spores had 1 to 4 transverse and 0 to 3 longitudinal septa. The morphological characteristics of the isolates showed considerably similar to well-known Alternaria tenuissima (2). The leaf spot disease caused by A. cinerariae of F. japonicum was reported from Japan (3). Spores of A. cinerariae are golden brown to brown with 3 to 9 transverse and 0 to 6 longitudinal septa and are 87.5 × 28.7 μm (avg.) (3). To verify pathogenicity of the isolate, the pure cultured fungi on the PDA medium was taken (4 mm in diameter) and placed on healthy leaves of Leopard plant. The artificially inoculated leaves were placed on wet filter paper in Petri dishes and incubated at 25°C and 80% humidity. At 7 days after inoculation, similar disease symptoms developed on 8 out of 10 infected Leopard plant leaves. The pathogen was reisolated from artificially infected leaves. To identify in molecular biology level, genomic DNA was extracted and the ITS-rDNA region was amplified using universal primers ITS1 and ITS4. The amplified PCR product was purified and sequenced (528 bp) with ITS1 and ITS4 primers for both directions and then deposited in GenBank (Accession No. KC415611.1). The BLAST search showed that it matched previously reported A. tenuissima with 100% identity. To the best of our knowledge, this is the first report of Leopard plant leaf spot disease in Korea. References: (1) E. Y. Kim et al. J. Ethnopharmacol. 146:40, 2013. (2) E. G. Simmons. Page 1 in: Alternaria Biology, Plant Diseases and Metabolites. J. Chelchowski and A. Visconti, eds. Elsevier, Amsterdam, 1992. (3) T. Sakoda et al. Res. Bull. Pl. Prot. Japan 46:73, 2010.

Plant Disease ◽  
2021 ◽  
Author(s):  
Yanxiang Qi ◽  
Yanping Fu ◽  
Jun Peng ◽  
Fanyun Zeng ◽  
Yanwei Wang ◽  
...  

Banana (Musa acuminate L.) is an important tropical fruit in China. During 2019-2020, a new leaf spot disease was observed on banana (M. acuminate L. AAA Cavendish, cv. Formosana) at two orchards of Chengmai county (19°48ʹ41.79″ N, 109°58ʹ44.95″ E), Hainan province, China. In total, the disease incidence was about 5% of banana trees (6 000 trees). The leaf spots occurred sporadically and were mostly confined to the leaf margin, and the percentage of the leaf area covered by lesions was less than 1%. Symptoms on the leaves were initially reddish brown spots that gradually expanded to ovoid-shaped lesions and eventually become necrotic, dry, and gray with a yellow halo. The conidia obtained from leaf lesions were brown, erect or curved, fusiform or elliptical, 3 to 4 septa with dimensions of 13.75 to 31.39 µm × 5.91 to 13.35 µm (avg. 22.39 × 8.83 µm). The cells of both ends were small and hyaline while the middle cells were larger and darker (Zhang et al. 2010). Morphological characteristics of the conidia matched the description of Curvularia geniculata (Tracy & Earle) Boedijn. To acquire the pathogen, tissue pieces (15 mm2) of symptomatic leaves were surface disinfected in 70% ethanol (10 s) and 0.8% NaClO (2 min), rinsed in sterile water three times, and transferred to potato dextrose agar (PDA) for three days at 28°C. Grayish green fungal colonies appeared, and then turned fluffy with grey and white aerial mycelium with age. Two representative isolates (CATAS-CG01 and CATAS-CG92) of single-spore cultures were selected for molecular identification. Genomic DNA was extracted from the two isolates, the internal transcribed spacer (ITS), large subunit ribosomal DNA (LSU rDNA), glyceraldehyde-3-phosphate dehydrogenase (GAPDH), translation elongation factor 1-alpha (TEF1-α) and RNA polymerase II second largest subunit (RPB2) were amplified and sequenced with universal primers ITS1/ITS4, LROR/LR5, GPD1/GPD2, EF1-983F/EF1-2218R and 5F2/7cR, respectively (Huang et al. 2017; Raza et al. 2019). The sequences were deposited in GenBank (MW186196, MW186197, OK091651, OK721009 and OK491081 for CATAS-CG01; MZ734453, MZ734465, OK091652, OK721100 and OK642748 for CATAS-CG92, respectively). For phylogenetic analysis, MEGA7.0 (Kumar et al. 2016) was used to construct a Maximum Likelihood (ML) tree with 1 000 bootstrap replicates, based on a concatenation alignment of five gene sequences of the two isolates in this study as well as sequences of other Curvularia species obtained from GenBank. The cluster analysis revealed that isolates CATAS-CG01 and CATAS-CG92 were C. geniculata. Pathogenicity assays were conducted on 7-leaf-old banana seedlings. Two leaves from potted plants were stab inoculated by puncturing into 1-mm using a sterilized needle and placing 10 μl conidial suspension (2×106 conidia/ml) on the surface of wounded leaves and equal number of leaves were inoculated with sterile distilled water serving as control (three replicates). Inoculated plants were grown in the greenhouse (12 h/12 h light/dark, 28°C, 90% relative humidity). Necrotic lesions on inoculated leaves appeared seven days after inoculation, whereas control leaves remained healthy. The fungus was recovered from inoculated leaves, and its taxonomy was confirmed morphologically and molecularly, fulfilling Koch’s postulates. C. geniculata has been reported to cause leaf spot on banana in Jamaica (Meredith, 1963). To our knowledge, this is the first report of C. geniculata on banana in China.


Plant Disease ◽  
2013 ◽  
Vol 97 (1) ◽  
pp. 138-138 ◽  
Author(s):  
B. Z. Fu ◽  
M. Yang ◽  
G. Y. Li ◽  
J. R. Wu ◽  
J. Z. Zhang ◽  
...  

Chinese bean tree, Catalpa fargesii f. duciouxii (Dode) Gilmour, is an ornamental arbor plant. Its roots, leaves, and flowers have long been used for medicinal purposes in China. During July 2010, severe outbreaks of leaf spot disease on this plant occurred in Kunming, Yunnan Province. The disease incidence was greater than 90%. The symptoms on leaves began as dark brown lesions surrounded by chlorotic halos, and later became larger, round or irregular spots with gray to off-white centers surrounded by dark brown margins. Leaf tissues (3 × 3 mm), cut from the margins of lesions, were surface disinfected in 0.1% HgCl2 solution for 3 min, rinsed three times in sterile water, plated on potato dextrose agar (PDA), and incubated at 28°C. The same fungus was consistently isolated from the diseased leaves. Colonies of white-to-dark gray mycelia formed on PDA, and were slightly brown on the underside of the colony. The hyphae were achromatic, branching, septate, and 4.59 (±1.38) μm in diameter on average. Perithecia were brown to black, globose in shape, and 275.9 to 379.3 × 245.3 to 344.8 μm. Asci that formed after 3 to 4 weeks in culture were eight-spored, clavate to cylindrical. The ascospores were fusiform, slightly curved, unicellular and hyaline, and 13.05 to 24.03 × 10.68 to 16.02 μm. PCR amplification was carried out by utilizing universal rDNA-ITS primer pair ITS4/ITS5 (2). Sequencing of the PCR products of DQ1 (GenBank Accession No. JN165746) revealed 99% similarity (100% coverage) with Colletotrichum gloeosporioides isolates (GenBank Accession No. FJ456938.1, No. EU326190.1, No. DQ682572.1, and No. AY423474.1). Phylogenetic analyses (MEGA 4.1) using the neighbor-joining (NJ) algorithm placed the isolate in a well-supported cluster (>90% bootstrap value based on 1,000 replicates) with other C. gloeosporioides isolates. The pathogen was identified as C. gloeosporioides (Penz.) Penz. & Sacc. (teleomorph Glomerella cingulata (Stoneman) Spauld & H. Schrenk) based on the morphological characteristics and rDNA-ITS sequence analysis (1). To confirm pathogenicity, Koch's postulates were performed on detached leaves of C. fargesii f. duciouxii, inoculated with a solution of 1.0 × 106 conidia per ml. Symptoms similar to the original ones started to appear after 10 days, while untreated leaves remained healthy. The inoculation assay used three leaves for untreated and six leaves for treated. The experiments were repeated once. C. gloeosporioides was consistently reisolated from the diseased tissue. C. gloeosporioides is distributed worldwide causing anthracnose on a wide variety of plants (3). To the best of our knowledge, this is the first report of C. gloeosporioides causing leaf spots on C. fargesii f. duciouxii in China. References: (1) B. C. Sutton. Page 1 in: Colletotrichum: Biology, Pathology and Control. CAB International. Wallingford, UK, 1992. (2) T. J. White et al. Page 315 in: PCR Protocols: A Guide to Methods and Applications. Academic Press, San Diego, 1990. (3) J. Yan et al. Plant Dis. 95:880, 2011.


Plant Disease ◽  
2019 ◽  
Vol 103 (7) ◽  
pp. 1786 ◽  
Author(s):  
H. Sun ◽  
C. Song ◽  
I. Mubeen ◽  
J. Huang

Plant Disease ◽  
2015 ◽  
Vol 99 (2) ◽  
pp. 289-289 ◽  
Author(s):  
Y. Z. Zhu ◽  
W. J. Liao ◽  
D. X. Zou ◽  
Y. J. Wu ◽  
Y. Zhou

In May 2014, a severe leaf spot disease was observed on walnut tree (Juglans regia L.) in Hechi, Guangxi, China. Leaf spots were circular to semicircular in shape, water-soaked, later becoming grayish white in the center with a dark brown margin and bordered by a tan halo. Necrotic lesions were approximately 3 to 4 mm in diameter. Diseased leaves were collected from 10 trees in each of five commercial orchards. The diseased leaves were cut into 5 × 5 mm slices, dipped in 75% ethanol for 30 s, washed three times in sterilized water, sterilized with 0.1% (w/v) HgCl2 for 3 min, and then rinsed five times with sterile distilled water. These slices were placed on potato dextrose agar (PDA), followed by incubating at 28°C for about 3 to 4 days. Fungal isolates were obtained from these diseased tissues, transferred onto PDA plates, and incubated at 28°C. These isolates produced gray aerial mycelium and then became pinkish gray with age. Moreover, the reverse of the colony was pink. The growth rate was 8.21 to 8.41 mm per day (average = 8.29 ± 0.11, n = 3) at 28°C. The colonies produced pale orange conidial masses and were fusiform with acute ends, hyaline, sometimes guttulate, 4.02 to 5.25 × 13.71 to 15.72 μm (average = 4.56 ± 0.31 × 14.87 ± 1.14 μm, n = 25). The morphological characteristics and measurements of this fungal isolate matched the previous descriptions of Colletotrichum fioriniae (Marcelino & Gouli) R.G. Shivas & Y.P. Tan (2). Meanwhile, these characterizations were further confirmed by analysis of the partial sequence of five genes: the internal transcribed spacer (ITS) of the ribosomal DNA, beta-tubulin (β-tub) gene, glyceraldehyde 3-phosphate dehydrogenase (GAPDH) gene, chitin synthase 3(CHS-1) gene, and actin (ACT) gene, with universal primers ITS4/ITS5, T1/βt2b, GDF1/GDR1, CHS1-79F/CHS1-354R, and ACT-512F/ACT-783R, respectively (1). BLAST of these DNA sequences using the nucleotide database of GenBank showed a high identify (ITS, 99%; β-tub, 99%; GAPDH, 99%; CHS-1, 99%; and ACT, 100%) with the previously deposited sequences of C. fioriniae (ITS, KF278459.1, NR111747.1; β-tub, AB744079.1, AB690809.1; GAPDH, KF944355.1, KF944354.1; CHS-1, JQ948987.1, JQ949005.1; and ACT, JQ949625.1, JQ949626.1). Koch's postulates were fulfilled by inoculating six healthy 1-year-old walnut trees in July 2014 with maximum and minimum temperatures of 33 and 26°C. The 6-mm mycelial plug, which was cut from the margin of a 5-day-old colony of the fungus on PDA, was placed onto each pin-wounded leaf, ensuring good contact between the mycelium and the wound. Non-colonized PDA plugs were placed onto pin-wounds as negative controls. Following inoculation, both inoculated and control plants were covered with plastic bags. Leaf spots, similar to those on naturally infected plants, were observed on the leaves inoculated with C. fioriniae within 5 days. No symptoms were observed on the negative control leaves. Finally, C. fioriniae was re-isolated from symptomatic leaves; in contrast, no fungus was isolated from the control, which confirmed Koch's postulates. To our knowledge, this is the first report of leaf disease on walnut caused by C. fioriniae. References: (1) L. Cai et al. Fungal Divers. 39:183, 2009. (2) R. G. Shivas and Y. P. Tan. Fungal Divers. 39:111, 2009.


Plant Disease ◽  
2021 ◽  
Author(s):  
Walftor Dumin ◽  
Mi-Jeong Park ◽  
You-Kyoung Han ◽  
Yeong-Seok Bae ◽  
Jong-Han Park ◽  
...  

Garlic (Allium sativum L. cv.namdo) is one of the most popular vegetables grown in Korea due to its high demand from the food industry. However, garlic is susceptible to a wide range of pest infestations and diseases that cause a significant decrease in garlic production, locally and globally (Schwartz and Mohan 2008). In early 2019, the occurrence of leaf blight disease was found spreading in garlic cultivation areas around Jeonnam (34.9671107, 126.4531825) province, Korea. Disease occurrence was estimated to affect 20% of the garlic plants and resulted in up to a 3-5% decrease in its total production. At the early stage of infection, disease symptoms were manifested as small, white-greyish spots with the occurrence of apical necrosis on garlic leaves. This necrosis was observed to enlarge, producing a water-soaked lesion before turning into a black-violet due to the formation of conidia. As the disease progressed, the infected leaves wilted, and the whole garlic plants eventually died. To identify the causal agent, symptomatic tissues (brown dried water-soak lesion) were excised, surface sterilized with 1% NaOCl and placed on the Potato Dextrose Agar (PDA) followed by incubation at 25°C in the dark for 5 days. Among ten fungal isolates obtained, four were selected for further analyses. On PDA, fungal colonies were initially greyish white in colour but gradually turned to yellowish-brown after 15 days due to the formation of yellow pigments. Conidia were muriform, brown in colour, oblong (almost round) with an average size of 18 – 22 × 16 – 20 μm (n = 50) and possessed 6 - 8 transverse septa. Fungal mycelia were branched, septate, and with smooth-walled hyphae. Morphological characteristics described above were consistent with the morphology of Stemphylium eturmiunum as reported by Simmons (Simmons, 2001). For molecular identification, molecular markers i.e. internal transcribed spacer (ITS) and calmodulin (cmdA) genes from the selected isolates were amplified and sequenced (White et al., 1990; Carbone and Kohn 1999). Alignment analysis shows that ITS and cmdA genes sequence is 100% identical among the four selected isolates. Therefore, representative isolate i.e. NIHHS 19-142 (KCTC56750) was selected for further analysis. BLASTN analysis showed that ITS (MW800165) and cmdA (LC601938) sequences of the representative isolates were 100% identical (523/523 bp and 410/410 bp) to the reference genes in Stemphylium eturmiunum isolated from Allium sativum in India (KU850545, KU850835) respectively (Woudenberg et al. 2017). Phylogenetic analysis of the concatenated sequence of ITS and cmdA genes confirmed NIHHS 19-142 isolates is Stemphylium eturmiunum. Pathogenicity test was performed using fungal isolate representative, NIHHS 19-142. Conidia suspension (1 × 106 conidia/µL) of the fungal isolate was inoculated on intact garlic leaves (two leaves from ten different individual plants were inoculated) and bulbs (ten bulbs were used) respectively. Inoculation on intact leaves was performed at NIHHS trial farm whereas inoculated bulbs were kept in the closed container to maintain humidity above 90% and incubated in the incubator chamber at 25°C. Result show that the formation of water-soaked symptoms at the inoculated site was observed at 14 dpi on intact leaves whereas 11 dpi on bulbs. As a control, conidia suspension was replaced with sterile water and the result shows no symptoms were observed on the control leaves and bulbs respectively. Re-identification of fungal colonies from symptomatic leaf and bulb was attempted. Result showed that the morphological characteristics and molecular marker sequences of the three colonies selected were identical to the original isolates thus fulfilled Koch’s postulates. Early identification of Stemphylium eturmiunum as a causal agent to leaf spot disease is crucial information to employ effective disease management strategies or agrochemical applications to control disease outbreaks in the field. Although Stemphylium eturmiunum has been reported to cause leaf spot of garlic disease in China, France and India (Woudenberg et al. 2017), to our knowledge, this is the first report of causing leaf spot disease on garlic in Korea.


Plant Disease ◽  
2021 ◽  
Author(s):  
Dongli Liu ◽  
Jing Li ◽  
Saisai Zhang ◽  
Xiangjing Wang ◽  
Wensheng Xiang ◽  
...  

Orychophragmus violaceus (L.) O. E. Schulz, also called February orchid or Chinese violet cress, belongs to the Brassicaceae family and is widely cultivated as a green manure and garden plant in China. During the prolonged rainy period in August 2020, leaf spot disease of O. violaceus was observed in the garden of Northeast Agricultural University, Harbin, Heilongjiang province. One week after the rainy days, the disease became more serious and the disease incidence ultimately reached approximately 80%. The disease symptoms began as small brown spots on the leaves, and gradually expanded to irregular or circular spots. As the disease progressed, spots became withered with grayish-white centers and surrounded by dark brown margins. Later on, the centers collapsed into holes. For severely affected plants, the spots coalesced into large necrotic areas and resulted in premature defoliation. No conidiophores or hyphae were present, and disease symptoms were not observed on other tissues of O. violaceus. To isolate the pathogen, ten leaves with typical symptoms were collected from different individual plants. Small square leaf pieces (5×5 mm) were excised from the junction of diseased and healthy tissues, disinfected in 75% ethanol solution for 1 min, rinsed in sterile distilled water, and then transferred to Petri dishes (9 cm in diameter) containing potato dextrose agar (PDA). After 3 days of incubation at 25 oC in darkness, newly grown-out mycelia were transferred onto fresh PDA and purified by single-spore isolation. Nine fungal isolates (NEAU-1 ~ NEAU-9) showing similar morphological characteristics were obtained and no other fungi were isolated. The isolation frequency from the leaves was almost 90%. On PDA plates, all colonies were grey-white with loose and cottony aerial hyphae, and then turned olive-green and eventually brown with grey-white margins. The fungus formed pale brown conidiophores with sparsely branched chains on potato carrot agar (PCA) plates after incubation at 25 oC in darkness for 7 days. Conidia were ellipsoidal or ovoid, light brown, and ranged from 18.4 to 59.1 × 9.2 to 22.3 µm in size, with zero to two longitudinal septa and one to five transverse septa and with a cylindrical light brown beak (n = 50). Based on the cultural and morphological characteristics, the fungus was tentatively identified as Alternaria tenuissima (Simmons 2007). Genomic DNA was extracted from the mycelia of five selected isolates (NEAU-1 ~ NEAU-5). The internal transcribed spacer region (ITS) was amplified and sequenced using primers ITS1/ITS4 (White et al., 1990). Blast analysis demonstrated that these five isolates had the same ITS sequence, and the ITS sequence of representative strain NEAU-5 (GenBank accession No. MW139354) showed 100% identity with the type strains of Alternaria alternata CBS916.96 and Alternaria tenuissima CBS918.96. Furthermore, the translation elongation factor 1-α gene (TEF), RNA polymerase II second largest subunit (RPB2), and glyceraldehyde 3-phosphate dehydrogenase (GPD) of representative strain NEAU-5 were amplified and sequenced using primers EF1-728F/EF1-986R (Carbone and Kohn 1999), RPB2-5F2/RPB2-5R (Sung et al., 2007), and Gpd1/Gpd2 (Berbee et al., 1999), respectively. The sequences of RPB2, GPD, and TEF of strain NEAU-5 were submitted to GenBank with accession numbers of MW401634, MW165223, and MW165221, respectively. Phylogenetic trees based on ITS, RPB2, GPD, and TEF were constructed with the neighbour-joining and maximum-likelihood algorithms using MEGA software version 7.0. The results demonstrated that strain NEAU-5 formed a robust clade with A. tenuissima CBS918.96 (supported by 99% and 96% bootstrap values) in the neighbour-joining and maximum-likelihood trees. As mentioned above, strain NEAU-5 produced seldomly branched conidial chains on PCA plates. The pattern is consistent with that of A. tenuissima (Kunze) Wiltshire, but distinct from that of A. alternata which could produce abundant secondary ramification (Simmons 2007). Thus, strain NEAU-5 was identified as A. tenuissima based on its morphology and phylogeny. Pathogenicity tests were carried out by inoculating five unwounded leaves with a conidial suspension of strain NEAU-5 (approximately 106 conidia/ml) on five different healthy plants cultivated in garden, and an equal number of leaves on the same plants inoculated with sterilized ddH2O served as negative controls. Inoculated and control leaves were covered with clear plastic bags for 3 days. After 6 days, small brown and irregular or circular spots were observed on all leaves inoculated with conidial suspension, while no such symptoms were observed in the control. The tests were repeated three times. Furthermore, the pathogenicity tests were also performed using 2-month-old potted plants in a growth chamber (28 oC, 90% relative humidity, 12 h/12 h light/dark) with two repetitions. Five healthy plants were inoculated by spraying 20 ml of a conidial suspension of strain NEAU-5 (approximately 106 conidia/ml) onto unwounded leaves. Five other healthy plants were inoculated with sterilized ddH2O as controls. After 7 days, similar symptoms were observed on leaves inoculated with strain NEAU-5, whereas no symptoms were observed in the control. The pathogen was reisolated from the inoculated leaves and identified as A. tenuissima by morphological and molecular methods. In all pathogenicity tests, A. tenuissima could successfully infect unwounded leaves of O. violaceus, indicating a direct interaction between leaves and A. tenuissima. It is known that high humidity and fairly high temperatures can favor the epidemics of Alternaria leaf spot (Yang et al., 2018), and this may explain why severe leaf spot disease of O. violaceus was observed after prolonged rain. Previously, it has been reported that Alternaria brassicicola and Alternaria japonica could cause leaf blight and spot disease on O. violaceus in Hebei and Jiangsu Provinces, China, respectively (Guo et al., 2019; Sein et al., 2020). Although these pathogens could lead to similar disease symptoms on the leaves of O. violaceus, it is easy to distinguish them by the morphological characteristics of conidiophores and ITS gene sequences. To our knowledge, this is the first report of A. tenuissima causing leaf spot disease of O. violaceus in China.


Plant Disease ◽  
2021 ◽  
Author(s):  
Yang Zhou ◽  
Rou Ye ◽  
Qin Ying ◽  
Yang Zhang ◽  
Linping Zhang

Dalbergia hupeana is a kind of wood and medicinal tree widely distributed in southern China. Since 2019, a leaf spot disease was observed on the leaves of D. hupeana in Gangxia village, Luoting town in Jiangxi Province, China (28°52′53″N, 115°44′58″E). The disease incidence was estimated to be above 50%. The symptoms began as small spots that gradually expanded, developing a brown central and dark brown to black margin. The spots ranged from 4 to 6 mm in diameter. Leaf pieces (5 × 5 mm) from lesion margins were surface sterilized in 70% ethanol for 30 s followed by 2% NaOCl for 1 min and then rinsed three times with sterile water. Tissues were placed on potato dextrose agar (PDA) and incubated at 25°C. Pure cultures were obtained by monosporic isolation. Fifteen strains with similar morphological characterizations were isolated, and three representative isolates (JHT-1, JHT-2, and JHT-3) were chosen and used for further study. Colonies on PDA of three isolates were grayish-green with white edges and dark green on the reverse side. Conidia were transparent, cylindrical with rounded ends, and measured 3.6-5.3 µm × 9.5-15.2 µm (3.7 ± 0.2 × 13.6 ± 1.1 µm, n = 100). Appressoria were dark brown, globose or subcylindrical, and ranged from 6.2-9.2 µm× 5.1-6.8 µm (7.9 ± 0.4 × 5.9 ± 0.3 µm, n=100). The morphological characteristics of the three strains were consistent with the description of species in the Colletotrichum gloeosporioides complex (Weir et al. 2012). The internal transcribed spacer (ITS) regions, actin (ACT), calmodulin (CAL), chitin synthase (CHS-1) and glyceraldehyde-3-phosphate dehydrogenase (GAPDH) and beta-tubulin 2 (TUB2) were amplified from genomic DNA for the three isolates using primers ITS1/ITS4, ACT-512F/ACT-783R, CL1/CL2, CHS-79F/CHS-345R, GDF/GDR and T1/Bt2b (Weir et al. 2012), respectively. The sequences were deposited in GenBank (Accession Nos. MZ482016 - MZ482018 for ITS; MZ463636 - MZ463638 for ACT; MZ463648- MZ463650 for CAL; MZ463639 - MZ463641 for CHS-1; MZ463642 - MZ463644 for GAPDH; MZ463645 - MZ463647 for TUB2). A neighbor-joining phylogenetic tree was constructed with MEGA 7.0 using the concatenation of multiple sequences (ITS, ACT, GAPDH, TUB2, CHS-1, CAL) (Kumar et al. 2016). According to the phylogenetic tree, three isolates fall within the Colletotrichum fructicola clade (boot support 99%). Based on morphological characteristics and phylogenetic analysis, three isolates were identified as C. fructicola. The pathogenicity of three isolates was conducted on two-yr-old seedlings (30 cm tall) of D. hupeana. Healthy leaves were wounded with a sterile needle and then inoculated with 10 μL spore suspension (106 conidia per mL). Controls were treated with sterile water. All plants were covered with transparent plastic bags and incubated in a greenhouse at 28°C with a 12 h photoperiod (relative humidity > 80%). Within five days, the inoculated leaves developed lesions similar to those observed in the field, whereas controls were asymptomatic. The experiments repeated three times showed similar results. The infection rate was 100%. C. fructicola was re-isolated from the lesions, whereas no fungus was isolated from control leaves. C. fructicola can cause leaf diseases in a variety of hosts, including Aesculus chinensis (Sun et al. 2020), Peucedanum praeruptorum (Ma et al. 2020), and Mandevilla × amabilis (Sun et al. 2020). C. brevisporum and C. gigasporum were also reported to infect Dalbergia odorifera (Chen et al. 2021; Wan et al. 2018). However, This is the first report of C. fructicola associated with leaf spot disease on D. hupeana in China. These results will help to develop effective strategies for appropriately managing this newly emerging disease.


Plant Disease ◽  
2021 ◽  
Author(s):  
Wang Hanyi ◽  
Hou Xiuming ◽  
Xueming Huang ◽  
Meng Gao ◽  
Tingting Chen ◽  
...  

Bletilla striata (Thunb.) Rchb. f. (Orchidaceae family, known as Baiji in Chinese) is a perennial herb and has been traditionally used for hemostasis and detumescence in China. In April of 2020, a leaf spot disease on B. striata was observed in plant nurseries (∼0.2 h) in Guilin, Guangxi Province, China. Approximately 20% of the plants were symptomatic, of which 150 plants were randomly selected for investigation. Initial symptoms include the appearance of small, circular or irregular light brown spots, randomly scattered on the edges and surfaces of the leaves, which progressively expand into large, suborbicular or irregular-shaped dark brown, necrotic areas. At the severe stage, the lesions coalesced into large necrotic areas and ultimately resulted in leaf abscission. To isolate the pathogen, three representative plants exhibiting symptoms were collected from the nurseries. Leaf tissues (5 × 5 mm) were cut from the margin of necrotic lesions (n = 18), surface-disinfected in 1% sodium hypochlorite (NaOCl) solution for 2 min, then rinsed three times in sterile water before isolation. The tissues were plated on potato dextrose agar (PDA) medium, and incubated at 28°C (12-h photoperiod) for 3 days. Hyphal tips from recently germinated spores were transferred to PDA to obtain pure cultures. Nine fungal isolates with similar morphological characteristics were obtained. Three single-spore isolates, BJ23.1, BJ55.1, and BJ91.3, were subjected to further morphological and molecular characterisation. Colonies on PDA plates were villose, had a dense growth of aerial mycelia and appeared white (1A1) to yellowish white (3A2). Macroconidia were smooth, hyaline, straight to slightly curved, usually contained three or five septa, and measuring 23.3 to 42.1 × 3.0 to 6.2 μm (mean ± SD: 31.2 ± 5.1 × 4.2 ± 0.6 μm, n = 50). Microconidia were generally cylindrical, straight to slightly curved, aseptate, and measuring 7.2 to 18.8 × 2.5 to 4.3 μm (mean ± SD: 12.1 ± 2.8 × 3.3 ± 0.5 μm, n = 62). Morphological characteristics are similar to those of F. commune (Skovgaard et al. 2003). For molecular identification, the genomic DNA of the isolates BJ23.1, BJ55.1, and BJ91.3 were extracted using the CTAB method (Guo et al. 2000). The internal transcribed spacer (ITS) region of rDNA, partial translation elongation factor-1 alpha (TEF-1α), RNA polymerase second largest subunit (RPB2), and the mitochondrial small subunit rDNA (mtSSU) genes were amplified using primer pairs [ITS1/ITS4 (White et al. 1990), EF-1/EF-2 (O’Donnell et al. 1998), and 5f2/11ar (Liu et al. 1999, Reeb et al. 2004), MS1/MS2 (Li et al. 1994), respectively]. The obtained sequences were deposited in NCBI GenBank under the following accession numbers: ITS (MZ424697 to MZ424699), TEF-1α (MZ513467 to MZ513469), RPB2 (MZ513473 to MZ513475), and mtSSU (MZ513470 to MZ513472). BLAST® analysis of the deposited sequences showed 99 to 100% identity with those of F. commune present in GenBank (Accession numbers: DQ016205, MH582348, MH582181, AF077383). In addition, a phylogenetic analysis using concatenated sequences of ITS, TEF-1α, mtSSU genes showed that BJ23.1, BJ55.1, and BJ91.3 located on the same clade with strains of F. commune. Therefore, based on morphological and molecular characteristics, the isolates were identified as F. commune (Skovgaard et al. 2003, Stewart et al. 2006). Pathogenicity was tested using 1.5-year-old B. striata plants. Healthy leaves on plants were inoculated with 5 × 5 mm mycelial discs of strains BJ23.1, BJ55.1, and BJ91.3 from 3-day-old PDA cultures, each isolate was inoculated onto three plants; three other plants inoculated with sterile PDA discs served as controls. All plants were enclosed in transparent plastic bags and incubated in a greenhouse at 28°C for 14 days (12-h photoperiod). Three days post-inoculation, leaf spot symptoms appeared on the inoculated leaves. No symptoms were detected on control plants. Experiments were replicated three times with similar results. To fulfill Koch’s postulates, F. commune was consistently re-isolated from symptomatic tissue and confirmed by morphology and sequencing, whereas no fungus was isolated from the control plants. F. commune has been reported to cause diseases on some plants, including sugarcane (Wang et al. 2018), maize (Xi et al. 2019) and Wax Gourd (Zeng et al. 2020). To our knowledge, this is the first report of F. commune causing leaf spot disease on B. striata in China. Identification of this pathogen provides the information for further studies to develop management strategies to control the disease.


Plant Disease ◽  
2020 ◽  
Author(s):  
Xiao Fei Chen ◽  
Yanqiao He ◽  
Aomen He ◽  
Juxia He ◽  
Qiongqiong Li ◽  
...  

Amygdalus triloba (Rosaceae; previously Prunus triloba) is a deciduous, flowering shrub that is widely used in the greening and beautification of lawns, parks and courtyards in China. In late May 2019, a leaf spot disease of A. triloba was observed on approximately 35% of plants in the Xinjiang Alaer city (40˚33′20′′N, 81˚17′19′′E). The disease symptoms began as small, suborbicular, brown spots on the leaves. As the disease progressed, the spots enlarged and coalesced into large necrotic areas and resulted in premature defoliation. Leaf sections (5 x 5 mm) from infected leaves were surface - sterilized with 75% ethanol for 30 s and 0.1% HgCl2 for 1 min, rinsed three times in sterile distilled water and then incubated on potato dextrose agar (PDA). Fifteen fungal isolates showing similar morphological characteristics were obtained by single-spore isolation. On the PDA plates, all fungal colonies had a dark olive color with loose, cottony mycelium. On the potato carrot agar, the fungus formed unbranched spore chains, but occasionally formed one or two lateral branches. Conidiophores were short, hazel-colored, septae, arising singly, and measuring 15.1 to 61.8 × 1.8 to 4.2 µm (35.2 ± 1.4 × 2.3 ± 0.1 µm, n = 50). Mature conidia were ellipsoidal to ovoid with a short conical beak at the tip, light brown with zero to three longitudinal septa and one to five transverse septa, and measuring 19.3 to 30.8 × 7.2 to 12.5 µm (21.8 ± 0.3 × 9.5 ± 0.2 µm, n = 50). Based on the cultural and morphological traits, the pathogen was preliminary identified as Alternaria tenuissima (Simmons 2007). Genomic DNA was extracted from the representative isolate YALAR-1, and the internal transcribed spacer (ITS) region, the partial coding sequence of endopolygalacturonase (endoPG), the glyceradehyde -3- phosphate dehydrogenase (GAPDA), the partial region of the histone 3 (H3) genes were amplified using primers ITS1/ITS4 (White et al. 1990), PG2b/PG3a (Andrew et al. 2009), GDF1/GDR1 (Berbee et al. 1999) and H3-1a/H3-1b (Glass and Donaldson 1995), respectively. The amplicons were sequenced and deposited in GenBank [MT459807 (ITS), MT459808 (endoPG), MT459805 (GAPDA), MT459806 (H3)]. MegaBLAST analyses revealed that our ITS, endoPG, GAPDA, and H3 sequences were 99-100% identical to those of A. tenuissima isolates in GenBank [AF347032 (ITS), KP124026 (endoPG), AY278809 (GAPDA), KF997086 (H3)], confirming the identity of the pathogen as A. tenuissima. Pathogenicity tests were performed by inoculating the fungus onto healthy, mature leaves of A. triloba in the field. Twenty five leaves (five leaves/plant) were sprayed with spore suspensions (1 × 106 spores/ml) of each fungal pathogen, and the same number of leaves were sprayed with distilled water as controls. Inoculated and control leaves were covered with clear plastic bags for 3 days. The experiment was repeated three times. Twelve days after inoculation, the observed symptoms were similar to the original symptoms and the same fungal pathogen was reisolated from the inoculated leaves and identified as A. tenuissima based on morphological features and sequence analysis. The control leaves remained asymptomatic and no fungus was isolated from these leaves. Previously, a leaf spot of A. triloba caused by Alternaria brassicae was reported in Dalian, China (Xie et al. 2017). In order to control this disease effectively, further studies are needed on the biology and ecology of A. tenuissima and A. brassicae respectively. To our knowledge, this is the first report of A. tenuissima associated with leaf spot disease on A. triloba in China. In late September 2020, the diseased plant rate increased to 38% in Alaer city. If the disease control and prevention is neglected, the landscape of Alaer city will be affected seriously. So, in order to effectively control the spread of the disease, it is urgent now to study the sensitivity of pathogen to fungicide and carry out the field efficacy trials. References: Andrew, M., et al. 2009. Mycologia. 101:95. Berbee, M. L., et al. 1999. Mycologia. 91:964. Glass, N. L., and Donaldson, G. C. 1995. Appl. Environ. Microbiol. 61:1323. Simmons, E. G. 2007. Alternaria: An Identification Manual. CBS Fungal Biodiversity Centre, Utrecht, The Netherlands. White, T. J., et al. 1990. PCR Protocols: A Guide to Methods and Applications. Academic Press, San Diego, CA. Xie, Y., et al. 2017. Liaoning Agricultural Sciences. 6: 73.


Sign in / Sign up

Export Citation Format

Share Document