scholarly journals First Report of Penicillium digitatum causing Postharvest rot of Citron Fruit in Kunming, China

Plant Disease ◽  
2021 ◽  
Author(s):  
Huan Ren ◽  
Gao Yang ◽  
Xue Li ◽  
Shijun Xing ◽  
Yating Gao ◽  
...  

Citron (Citrus medica L.) is a perennial evergreen woody tree of Rutaceae family and Genus of Citrus. The citron is cultivated for its economic, medicinal and ornamental values in the south of China. (Yang et al., 2015). The shapes range from spherical to ovate and the sizes range from 3 to 5 kg (Klein et al., 2016). In June 2021, some postharvest citron fruits (Citrus medica var. medica) were found to have decay with a green or greyish mycelium on part or whole citron in 2 farmer’s markets in Kunming city, Yunnan Province (N 25°02′; E 102°42′), southwest China. Initial symptoms appeared as white, brown, and irregular necrotic spots in the pericarp. The lesions enlarged gradually and developed into green, water-soaked areas which extend rapidly. Eventually, the diseased fruits were rotten, soften, and the green spore masses confined to the surface (Fig. 1A). The incidence of this disease in postharvest citron fruits ranges from 15 % to 35 %, which is extremely destructive to the fruit of Rutaceae family plants (Chen et al., 2019). Small pieces (5 mm2) of symptomatic citron fruits were surface disinfected in 75 % ethanol and 0.3 % NaClO for 30 s and 2 min respectively, rinsed with distilled water for three times, blotted dry, placed onto potato dextrose agar (PDA) medium aseptically and incubated in a growth chamber at 25 ± 1 ℃, after 7 days, different colonies grew on PDA plates that were isolated and purified on new PDA medium at 25 ± 1 ℃ for 7 days. Inoculating repeatedly until six single-strain (XY01 to XY06) were obtained, and these isolates were stored in 15 % glycerol at –80 ℃ in a refrigerator in the State Key Laboratory for Conservation and Utilization of Bio-Resources in Yunnan Agricultural University. The selected pathogens (XY01 to XY06) were inoculated on PDA medium, incubated at 25 ± 1 ℃. After 7 days, colonies of the isolate obverse are olive green, the white margin and greyish-green spores on the surface, and the reverse colorless to cream yellow or pale dull brown. Colonies texture was velutinous, with a special fragrance. The conidia structure was very fragile and break up easily into many cellular elements. Conidiophores were terverticillate, produced by subsurface or aerial hyphae, irregularly branched and composed of short stipes with few metulae and branches that terminate in whorls of three to six phialides, which are often solitary, cylindrical with a short neck. Conidia are hyaline to pale green, smooth-walled, without septate, partially ellipsoidal, or obovate (4.9 to11.9× 4.3 to 8.9 μm). Partial cylindrical (8.2 to 10.5× 2.7 to 5.3 μm), there are some small conidia, which were ellipsoidal or spherical (3.9 to 5.2× 2.7 to 5.2 μm). According to morphological characteristics, the fungus was identified as Penicillium digitatum (Pers.) Sacc. Isolate XY01 and XY02 were used for molecular identification and genomic DNA was extracted using the CTAB method (Aboul-Maaty & Oraby, 2019). The universal primers ITS1 and ITS4 were used to amplify and sequence the ITS1, 5.8S, and ITS2 rDNA region. Using NCBI’s BLASTn tools, the nucleotide sequences of XY01 and XY02 (Gen-Bank accessions MZ976843 and OK513274) show 100 % identity to MK450692 (P. digitatum strain CMV010G4). Pathogenicity tests have used the fruits (Citrus medica), which maturity was more than 80%. The pathogens (XY01, XY02) were cultured for 7 days on PDA medium, washed with sterilized water the resulting spore suspensions diluted to 1.0 × 106 spores/ml. Wounds (0.5 × 0.5 cm) were made on the surface of citron fruits by scraping with a sterile scalpel and then treated with 200 µl of spore suspension (Wild, 1994). Control citron fruits were treated with sterile water. citron fruits were incubated at 24-26 °C. Each treatment was performed in triplicate with 6 citron fruits. After 3 days, all fruits had developed lesions, in a water-stained, pale brown, and rapidly formed white hyphae, white mold layer was observed with a length of 1.5-2.5 cm and a width of 1-2 cm (Fig.1C), but control did induce infection. After 7 days, decay developed more quickly, the hyphae rapidly expanded on the surface of the pericarp, with vague and irregular edges, then a green mold layer was formed, the whole fruit was observed to rot and soften, When the citron was cut, the white flesh inside turned black and rotted (Fig.1B). P. digitatum was consistently reisolated from the inoculated plants but not from the controls. No symptoms developed on the control (Fig.1D). According to Koch’s postulates, the inoculated strains of XY01 and XY02 were the isolates causing citron decay disease. Based on symptoms, morphological characteristics, rDNA-ITS sequence analysis, and pathogenicity, this fungus was identified as P. digitatum. To our knowledge, this is the first report of the distribution of P. digitatum on Citron (Citrus medica) 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 ◽  
2014 ◽  
Vol 98 (5) ◽  
pp. 691-691 ◽  
Author(s):  
B. Z. Fu ◽  
Z. H. Zhang ◽  
L. H. Wang ◽  
G. Y. Li ◽  
J. Z. Zhang ◽  
...  

The Chinese dwarf banana (Ensete lasiocarpum) is one of the ornamental bananas that belongs to Musaceae family. The plant is native to the southwestern China, where it grows semi-wild in the mountains between 1,500 and 2,500 m above sea level. During July 2011, a leaf spot disease on this plant was observed in the campus and parks in Kunming, Yunnan Province. The incidence level was about 22%, mainly on the old leaves. The leaf symptoms were irregular spots with gray to off-white centers surrounded by dark brown margins, and usually also surrounded by chlorotic halos. Leaf tissues (3 × 5 mm), cut from the margins of lesions, were surface-disinfected (95% ethanol for 3 min, 0.1% HgCl2 for 2 min, rinsed three times with sterile water), plated on potato sucrose agar (PSA), and incubated at 26°C under natural lights. The same fungus was consistently isolated from the diseased leaves. Colonies of white-to-dark gray mycelia formed on PSA that were black on the underside. The colonies were further identified as Alternaria sp. based on the dark brown, obclavate to obpyriform catenulate conidia with longitudinal and transverse septa tapering to a prominent beak attached in chains on a simple and short conidiophore (2). Conidia were 5.26 to 30.26 μm long and 3.95 to 15.79 μm wide, averaging 10.21 (±3.17) × 20.02 (±5.75) μm (n = 50), with a beak length of 0 to 7.89 μm, and had 3 to 8 transverse and 0 to 3 longitudinal septa. PCR amplification was carried out by utilizing universal rDNA-ITS primer pair ITS4/ITS5 (1). The ITS region of isolate DY1 (GenBank Accession No. KF516556) was 572 bp in length. BLAST search revealed 99% identity with two Alternaria alternata isolates (JF440581.1 and GQ121322.2). Phylogenetic analysis (MEGA 5.1) using the neighbor-joining algorithm placed the isolate in a well-supported cluster with other A. alternata isolates. The pathogen was identified as A. alternate (Fr.:Fr.) Keissler based on the morphological characteristics and rDNA-ITS sequence analysis. To confirm pathogenicity, Koch's postulates were performed on detached leaves of E. lasiocarpum inoculated with mycelial plugs with ddH2O and agar plugs as a control. Leaf spots identical to those observed in the field developed in 9 days on the inoculated leaves but not on the control. The inoculation assay used three leaves, totaling 72 spots for control and 36 spots for inoculation. The experiments were repeated once. A. alternata was consistently re-isolated from the inoculated leaves. The symptom developed easier with wounds. To our knowledge, this is the first report of E. lasiocarpum leaf spot disease caused by A. alternata in China and the world. References: (1) T. J. White et al. Page 315 in: PCR Protocols: A Guide to Methods and Applications. Academic Press, San Diego, 1990. (2) T. Y. Zhang. Flora Fungorum Sinicorum, Vol. 16: Alternaria. Science Press, Beijing, China, 2003.


Plant Disease ◽  
2011 ◽  
Vol 95 (8) ◽  
pp. 1033-1033 ◽  
Author(s):  
L. L. Chern ◽  
L. S. Jie ◽  
C. T. Feng ◽  
W. C. Ho

Many Calathea species in the family Marantaceae are beautiful ornamental plants with variegated foliage. Among them, C. picturata ‘Argentea’, an evergreen perennial that has pale green leaves with dark green margins and a red underside, is a popular houseplant in Taiwan. In 2004, a new foliage disease that caused leaf blight of C. picturata ‘Argentea’ was first observed in a nursery in southern Taiwan. Initial symptoms were tiny, brown spots that appeared on the leaves of all ages, which quickly enlarged and coalesced. These necrotic lesions spread to cover the entire leaves in high temperature and moisture conditions and caused leaves to shrivel and eventually die. A dematiaceous hyphomycete with multicelled conidia was consistently isolated from the diseased leaves after being surfaced sterilized with 10% Clorox and placed on vegetable juice agar (10% V8 juice, 0.02% CaCO3, and 2% agar [VJA]). Pathogenicity of the isolate was tested by spraying ‘Argentea’ calathea leaves with a conidia suspension (1.6 × 105 conidia/ml) prepared from a culture grown on VJA at 28°C for 7 days. Plant leaves sprayed with distilled water were used as a control. Three pots of 15-cm high ‘Argentea’ calathea plants were inoculated with 10 ml of a conidia suspension and the experiment was conducted twice at 28°C and 90% relative humidity in a growth chamber. Tiny, brown spots started to show on all inoculated leaves 5 days after inoculation and the progression of symptom development was similar to that observed in nature. Control leaves remained asymptomatic. The same dematiaceous hyphomycete fungus was reisolated from 13 of 16 disease tissues taken from four symptomatic leaves. A colony of the calathea isolate was olive green when grown on potato dextrose agar (PDA) and conidia production was observed 7 days after incubation in darkness. The conidiophores were either branches from or the ends of normal mycelium, some of them geniculate with conidium produced at each bend measuring 142 to 602 (340) × 3 to 6 (4) μm on disease tissues and 51 to 150 (103) × 3 to 5 (4) μm on PDA. Conidia were multicelled with protruding hilum at the base, terminal cells thickened, olivaceous brown or golden brown in fusiform shape with blunt tips, 5 to 11 septate on disease tissues and 6 to 11 septate on PDA, measuring 46 to 166 (95) × 8 to 19 (13) μm on disease tissues and 58 to 145 (94) × 6 to 15 (11) μm on PDA, germinating by producing germ tubes semiaxially from each end. Morphological characteristics of the calathea isolate fit the description of the genus Exserohilum (2). Comparison of rDNA internal transcribed spacer (ITS) sequence of the calathea isolate with those in GenBank revealed that it shared 99.5% (549 of 552) similarity with a published sequence (GenBank Accession No. EU571210) (3) and Exserohilum rostratum was its closest species. ITS sequence analysis was done as previously described (1). Morphological and molecular data identified the pathogen as E. rostratum (Drechs.) Leonard & Suggs (= Bipolaris rostrata (Drechs.) Shoemaker). To our knowledge, this is the first report of leaf blight caused by E. rostratum on C. picturata in Taiwan. References: (1) L. L. Chern et al. Plant Dis. 94:1164, 2010. (2) K. J. Leonard. Mycologia 68:402, 1976. (3) R. Sappapan et al. J. Nat. Prod. 71:1657, 2008.


Plant Disease ◽  
2021 ◽  
Author(s):  
Charles Krasnow ◽  
Nancy Rechcigl ◽  
Jennifer Olson ◽  
Linus Schmitz ◽  
Steven N. Jeffers

Chrysanthemum (Chrysanthemum × morifolium) plants exhibiting stem and foliage blight were observed in a commercial nursery in eastern Oklahoma in June 2019. Disease symptoms were observed on ~10% of plants during a period of frequent rain and high temperatures (26-36°C). Dark brown lesions girdled the stems of symptomatic plants and leaves were wilted and necrotic. The crown and roots were asymptomatic and not discolored. A species of Phytophthora was consistently isolated from the stems of diseased plants on selective V8 agar (Lamour and Hausbeck 2000). The Phytophthora sp. produced ellipsoid to obpyriform sporangia that were non-papillate and persistent on V8 agar plugs submerged in distilled water for 8 h. Sporangia formed on long sporangiophores and measured 50.5 (45-60) × 29.8 (25-35) µm. Oospores and chlamydospores were not formed by individual isolates. Mycelium growth was present at 35°C. Isolates were tentatively identified as P. drechsleri using morphological characteristics and growth at 35°C (Erwin and Ribeiro 1996). DNA was extracted from mycelium of four isolates, and the internal transcribed spacer (ITS) region was amplified using universal primers ITS 4 and ITS 6. The PCR product was sequenced and a BLASTn search showed 100% sequence similarity to P. drechsleri (GenBank Accession Nos. KJ755118 and GU111625), a common species of Phytophthora that has been observed on ornamental and vegetable crops in the U.S. (Erwin and Ribeiro 1996). The gene sequences for each isolate were deposited in GenBank (accession Nos. MW315961, MW315962, MW315963, and MW315964). These four isolates were paired with known A1 and A2 isolates on super clarified V8 agar (Jeffers 2015), and all four were mating type A1. They also were sensitive to the fungicide mefenoxam at 100 ppm (Olson et al. 2013). To confirm pathogenicity, 4-week-old ‘Brandi Burgundy’ chrysanthemum plants were grown in 10-cm pots containing a peat potting medium. Plants (n = 7) were atomized with 1 ml of zoospore suspension containing 5 × 103 zoospores of each isolate. Control plants received sterile water. Plants were maintained at 100% RH for 24 h and then placed in a protected shade-structure where temperatures ranged from 19-32°C. All plants displayed symptoms of stem and foliage blight in 2-3 days. Symptoms that developed on infected plants were similar to those observed in the nursery. Several inoculated plants died, but stem blight, dieback, and foliar wilt were primarily observed. Disease severity averaged 50-60% on inoculated plants 15 days after inoculation. Control plants did not develop symptoms. The pathogen was consistently isolated from stems of symptomatic plants and verified as P. drechsleri based on morphology. The pathogenicity test was repeated with similar results. P. drechsleri has a broad host range (Erwin and Ribeiro 1996; Farr et al. 2021), including green beans (Phaseolus vulgaris), which are susceptible to seedling blight and pod rot in eastern Oklahoma. Previously, P. drechsleri has been reported on chrysanthemums in Argentina (Frezzi 1950), Pennsylvania (Molnar et al. 2020), and South Carolina (Camacho 2009). Chrysanthemums are widely grown in nurseries in the Midwest and other regions of the USA for local and national markets. This is the first report of P. drechsleri causing stem and foliage blight on chrysanthemum species in the United States. Identifying sources of primary inoculum may be necessary to limit economic loss from P. drechsleri.


Plant Disease ◽  
2014 ◽  
Vol 98 (6) ◽  
pp. 843-843 ◽  
Author(s):  
N.-H. Lu ◽  
Q.-Z. Huang ◽  
H. He ◽  
K.-W. Li ◽  
Y.-B. Zhang

Avicennia marina is a pioneer species of mangroves, a woody plant community that periodically emerges in the intertidal zone of estuarine regions in tropical and subtropical regions. In February 2013, a new disease that caused the stems of A. marina to blacken and die was found in Techeng Island of Zhanjiang, Guangdong Province, China. Initial symptoms of the disease were water-soaked brown spots on the biennial stems that coalesced so whole stems browned, twigs and branches withered, leaves defoliated, and finally trees died. This disease has the potential to threaten the ecology of the local A. marina community. From February to May 2013, 11 symptomatic trees were collected in three locations on the island and the pathogen was isolated as followed: tissues were surface disinfected with 75% ethanol solution (v/v) for 20 s, soaked in 0.1% mercuric chloride solution for 45 s, rinsed with sterilized water three times, dried, placed on potato dextrose agar (PDA), and incubated for 3 to 5 days at 28°C without light. Five isolates (KW1 to KW5) with different morphological characteristics were obtained, and pathogenic tests were done according Koch's postulates. Fresh wounds were made with a sterile needle on healthy biennial stems of A. marina, and mycelial plugs of each isolate were applied and covered with a piece of wet cotton to maintain moisture. All treated plants were incubated at room temperature. Similar symptoms of black stem were observed only on the stems inoculated the isolate KW5 after 35 days, while the control and all stems inoculated with the other isolates remained symptomless. An isolate similar to KW5 was re-isolated from the affected materials. The pathogenic test was repeated three times with the same conditions and it was confirmed that KW5 was the pathogen causing the black stem of A. marina. Hyphal tips of KW5 were transferred to PDA medium in petri dishes for morphological observation. After 48 to 72 h, white, orange, or brown flocculence patches of KW5 mycelium, 5.0 to 6.0 cm in diameter, grew. Tapering and spindle falciform macroconidia (11 to 17.3 μm long × 1.5 to 2.5 μm wide) with an obviously swelled central cell and narrow strips of apical cells and distinctive foot cells were visible under the optical microscope. The conidiogenous cells were intertwined with mycelia and the chlamydospores were globose and formed in clusters. These morphological characteristics of the isolate KW5 are characteristic of Fusarium equiseti (1). For molecular identification, the ITS of ribosomal DNA, β-tubulin, and EF-1α genes were amplified using the ITS4/ITS5 (5), T1/T2 (2), and EF1/EF2 (3) primer pairs. These sequences were deposited in GenBank (KF515650 for the ITS region; KF747330 for β-tubulin region, and KF747331 for EF-1α region) and showed 98 to 99% identity to F. equiseti strains (HQ332532 for ITS region, JX241676 for β-tubulin gene, and GQ505666 for EF-1α region). According to both morphological and sequences analysis, the pathogen of the black stem of A. marina was identified as F. equiseti. Similar symptoms on absorbing rootlets and trunks of A. marina had been reported in central coastal Queensland, but the pathogen was identified as Phytophthora sp. (4). Therefore, the disease reported in this paper differs from that reported in central coastal Queensland. To our knowledge, this is the first report of black stems of A. marina caused by F. equiseti in China. References: (1) J. F. Leslie and B. A. Summerell. The Fusarium Laboratory Manual, 1st ed. Wiley-Blackwell, Hoboken, NJ, 2006. (2) K. O'Donnell and E. Cigelnik. Mol. Phylogenet. Evol. 7:103, 1997. (3) K. O'Donnell et al. Proc. Natl. Acad. Sci. USA. 95:2044, 1998. (4) K. G. Pegg. Aust et al. Plant Pathol. 3:6, 1980. (5) A. W. Zhang et al. Plant Dis. 81:1143, 1997.


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 ◽  
2021 ◽  
Author(s):  
Qing Qu ◽  
Liu Shiwei ◽  
Ning Liu ◽  
Yunxia Liu ◽  
Jia Hui ◽  
...  

Abelmoschus manihot (Linn. ) Medicus (A. manihot) is an annual to perennial herb of the Malvaceae okra, mainly distributed in Guangdong, Guangxi, Fujian, Hunan, Hubei provinces. It can not only be used as an ornamental flower, but also has important economic and medicinal value. Last year, 10% A. manihot in 1,000 acres were observed with stalk rot in the Zhongshang Agricultural Industrial Park, 50 meters east of Provincial Highway 235 in Gaoyang County of Hebei province. Internal discoloration of the stem began brown to black discoloration of the vascular system and became hollow, with the mycelium growing on the surface. Stems from symptomatic plants (approximately 5 mm2) were dissected, washed free of soil, then soaked in 75% ethanol for 16 s to surface-sterilize, and 40 s in HgCl2, then rinsed three times in sterile water. After being dried with blotting paper, five pieces were placed on potato dextrose agar (PDA). After cultured 2 or 3 days, five isolates were purified and re-cultured on PDA in the dark at 25°C. The color of the colony was white. The hyphae were radial in PDA, and the aerial hyphae were flocculent, well-developed with luxuriant branches. The colonies were white and floccus, and the aerial hyphae were well developed, branched and without septum on corn meal agar (CMA). The sporangia were large or petal shaped, composed of irregular hyphae, terminal or intermediate , with the size of (31.6-88.4) μm ×(12.7- 14.6) μm. Vesicles were spherical, terminal or intermediate, ranging from 14.6 to 18.5μm. Oogonia were globose, terminal and smooth which stipe was straight. Antheridia were clavate or baggy and mostly intercalary, sometimes terminal. Oospores were aplerotic, 21.5 to 30.0 μm in diameter, 1.6 to 3.1 μm in wall thickness. The isolates morphological characteristics were consistent with P. aphanidermatum (van der Plaats-Niterink 1981, Wu et al. 2021 ). To identify the isolates, universal primers ITS1/ITS4 (White et al. 1993) were used for polymerase chain reaction–based molecular identification. The amplification region was sequenced by Sangon Biotech (Shanghai, China) and submitted to GenBank (MW819983). BLAST analysis showed that the sequence was 100% identical to Pythium aphanidermatum. Pathogenicity tests were conducted 3 times, with 4 treatments and 2 controls each time. The plants treated were 6 months old. Then the hyphae growing on PDA for 7 days were cut into four pieces. Next, they were inoculated into the soil of the A. manihot. Negative control was inoculated only with PDA for 7 days ( Zhang et al. 2000). The plants were then placed in a greenhouse under 28°C, 90% relative humidity. After inoculated 20 to 30 days, the infected plants showed stalk rot, the same symptoms as observed on the original plants. The control plants didn’t display symptoms. Pythium aphanidermatum was re-isolated from infected stems and showed the same characteristics as described above and was identical in appearance to the isolates used to inoculate the plants. To our knowledge, this is the first report of Pythium aphanidermatum infecting A. manihot stem and causing stalk rot in China. It may become a significant problem for A. manihot. Preliminary management practices are needed for reducing the cost and losses of production.


Plant Disease ◽  
2012 ◽  
Vol 96 (1) ◽  
pp. 147-147
Author(s):  
S. H. Lee ◽  
C. K. Lee ◽  
M. J. Park ◽  
H. D. Shin

Aralia elata (Miq.) Seem., known as Japanese angelica tree, is a deciduous shrub belonging to the Araliaceae, which is native to East Asia. The young shoots have long been used in various dishes in East Asia. Commercial cultivation of this shrub, especially in polytunnels, is expanding in Korea. Several diseases including Sclerotinia rot have been known to be present on this plant (1,2). In early September 2007, leaf spot symptoms were first observed on several trees in Hongcheon, Korea. Microscopic observations revealed that the leaf spots were associated with an Ascochyta sp. Further surveys of the Ascochyta leaf spot showed the occurrence of the disease in approximately 5 to 10% of the trees in the 3 ha of commercial fields surveyed in Chuncheon, Gapyeong, Inje, and Jinju, Korea. Initial symptoms on leaves were circular to irregular, brown to dark brown, becoming zonate, and finally fading to grayish brown in the center with a yellow halo. Representative samples were deposited in the herbarium of Korea University. Conidiomata on leaf lesions were pycnidial, amphigenous, but mostly epiphyllous, immersed or semi-immersed in host tissue, light brown to olive brown, and 60 to 200 μm in diameter. Ostioles were papillate, 20 to 35 μm wide, and surrounded by a ring of darker cells. Conidia were hyaline, smooth, cylindrical to clavate, straight to mildly curved, slightly constricted at the septa, medianly one-septate, sometimes aseptate, 8 to 16 × 2.5 to 3.5 μm, and contained small oil drops. These morphological characteristics were consistent with the previous reports of Ascochyta marginata J.J. Davis (3,4). A monoconidial isolate was cultured on potato dextrose agar (PDA) plates and accessioned in the Korea Agricultural Culture Collection (Accession KACC43082). The conidia were readily formed on PDA. Inoculum for the pathogenicity tests was prepared by harvesting conidia from 30-day-old cultures of KACC43082 and a conidial suspension (approximately 2 × 106 conidia/ml) was sprayed onto leaves of three healthy seedlings. Three noninoculated seedlings served as controls. Inoculated and noninoculated plants were covered with plastic bags for 48 h in a glasshouse. After 7 days, typical leaf spot symptoms started to develop on the leaves of the inoculated plants. The fungus, A. marginata, was reisolated from those lesions, confirming Koch's postulates. No symptoms were observed on control plants. Previously, the disease was reported in Japan (4) and China (3). To our knowledge, this is the first report of A. marginata on Japanese angelica trees in Korea. According to our field observations in Korea, the Ascochyta leaf spot mostly occurred on plants growing in a humid environment, especially during the rainy season. The seedlings as well as the trees growing in sunny, well-ventilated plots were nearly free from this disease. Therefore, the growing conditions seemed to be the most important factor for the development and severity of the disease. References: (1) C. K. Lee et al. Plant Pathol. J. 26:426, 2010. (2) S. H. Lee et al. Diseases of Japanese Angelica Tree and Their Control. Research Report 08-10. Korea Forest Research Institute. Seoul, Korea, 2008. (3) J. Sun et al. Acta Mycol. Sin. 14:107, 1995. (4) M. Yoshikawa and T. Yokoyama. Mycoscience 36:67, 1995.


Plant Disease ◽  
2020 ◽  
Author(s):  
Yue Lian Liu ◽  
Jian Rong Tang ◽  
Yu Han Zhou

Monstera deliciosa Liebm is an ornamental foliage plant (Zhen et al. 2020De Lojo and De Benedetto 2014). In July of 2019, anthracnose lesions were observed on leaves of M. deliciosa cv. Duokong with 20% disease incidence of 100 plants at Guangdong Ocean University campus (21.17N,110.18E), Guangdong Province, China. Initially affected leaves showed chlorotic spots, which coalesced into larger irregular or circular lesions. The centers of spots were gray with a brown border surrounded by a yellow halo (Supplementary figure 1). Twenty diseased leaves were collected for pathogen isolation. Margins of diseased tissue was cut into 2 × 2 mm pieces, surface-disinfected with 75% ethanol for 30 s and 2% sodium hypochlorite (NaOCl) for 60 s, rinsed three times with sterile water before isolation. Potato dextrose agar (PDA) was used to culture pathogens at 28℃ in dark. Successively, pure cultures were obtained by transferring hyphal tips to new PDA plates. Fourteen isolates were obtained from 20 leaves. Three single-spore isolates (PSC-1, PSC-2, and PSC-3) were obtained ,obtained, which were identical in morphology and molecular analysis (ITS). Therefore, the representative isolate PSC-1 was used for further study. The culture of isolate PSC-1 on PDA was initially white and later became cottony, light gray in 4 days, at 28 °C. Conidia were single celled, hyaline, cylindrical, clavate, and measured 13.2 to 18.3 µm × 3.3 to 6.5 µm (n = 30). Appressoria were elliptical or subglobose, dark brown, and ranged from 6.3 to 9.5 µm × 5.7 to 6.5 µm (n = 30). Morphological characteristics of isolate PSC-1 were consistent with the description of Colletotrichum siamense (Prihastuti et al. 2009; Sharma et al. 2013). DNA of the isolate PSC-1 was extracted for PCR sequencing using primers for the rDNA ITS (ITS1/ITS4), GAPDH (GDF1/GDR1), ACT (ACT-512F/ACT-783R), CAL (CL1C/CL2C), and TUB2 (βT2a/βT2b) (Weir et al. 2012). Analysis of the ITS (accession no. MN243535), GAPDH (MN243538), ACT (MN512640), CAL (MT163731), and TUB2 (MN512643) sequences revealed a 97-100% identity with the corresponding ITS (JX010161), GAPDH (JX010002), ACT (FJ907423), CAL (JX009714) and TUB2 (KP703502) sequences of C. siamense in GenBank. A phylogenetic tree was generated based on the concatenated sequences of ITS, GAPDH, ACT, CAL, and TUB2 which clustered the isolate PSC-1 with C. siamense the type strain ICMP 18578 (Supplementary figure 2). Based on morphological characteristics and phylogenetic analysis, the isolate PSC-1 associated with anthracnose of M. deliciosa was identified as C. siamense. Pathogenicity test was performed in a greenhouse at 24 to 30oC with 80% relative humidity. Ten healthy plants of cv. Duokong (3-month-old) were grown in pots with one plant in each pot. Five plants were inoculated by spraying a spore suspension (105 spores ml-1) of the isolate PSC-1 onto leaves until runoff, and five plants were sprayed with sterile water as controls. The test was conducted three times. Anthracnose lesions as earlier were observed on the leaves after two weeks, whereas control plants remained symptomless. The pathogen re-isolated from all inoculated leaves was identical to the isolate PSC-1 by morphology and ITS analysis, but not from control plants. C. gloeosporioides has been reported to cause anthracnose of M. deliciosa (Katakam, et al. 2017). To the best of our knowledge, this is the first report of C. siamense causing anthracnose on M. deliciosa in ChinaC. siamense causes anthracnose on a variety of plant hosts, but not including M. deliciosa (Yanan, et al. 2019). To the best of our knowledge, this is the first report of C. siamense causing anthracnose on M. deliciosa, which provides a basis for focusing on the management of the disease in future.


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.


Sign in / Sign up

Export Citation Format

Share Document