Your browser doesn't support javascript.
loading
Mostrar: 20 | 50 | 100
Resultados 1 - 20 de 559
Filtrar
1.
BMC Plant Biol ; 24(1): 641, 2024 Jul 06.
Artigo em Inglês | MEDLINE | ID: mdl-38971719

RESUMO

BACKGROUND: Early blight and brown leaf spot are often cited as the most problematic pathogens of tomato in many agricultural regions. Their causal agents are Alternaria spp., a genus of Ascomycota containing numerous necrotrophic pathogens. Breeding programs have yielded quantitatively resistant commercial cultivars, but fungicide application remains necessary to mitigate the yield losses. A major hindrance to resistance breeding is the complexity of the genetic determinants of resistance and susceptibility. In the absence of sufficiently resistant germplasm, we sequenced the transcriptomes of Heinz 1706 tomatoes treated with strongly virulent and weakly virulent isolates of Alternaria spp. 3 h post infection. We expanded existing functional gene annotations in tomato and using network statistics, we analyzed the transcriptional modules associated with defense and susceptibility. RESULTS: The induced responses are very distinct. The weakly virulent isolate induced a defense response of calcium-signaling, hormone responses, and transcription factors. These defense-associated processes were found in a single transcriptional module alongside secondary metabolite biosynthesis genes, and other defense responses. Co-expression and gene regulatory networks independently predicted several D clade ethylene response factors to be early regulators of the defense transcriptional module, as well as other transcription factors both known and novel in pathogen defense, including several JA-associated genes. In contrast, the strongly virulent isolate elicited a much weaker response, and a separate transcriptional module bereft of hormone signaling. CONCLUSIONS: Our findings have predicted major defense regulators and several targets for downstream functional analyses. Combined with our improved gene functional annotation, they suggest that defense is achieved through induction of Alternaria-specific immune pathways, and susceptibility is mediated by modulating hormone responses. The implication of multiple specific clade D ethylene response factors and upregulation of JA-associated genes suggests that host defense in this pathosystem involves ethylene response factors to modulate jasmonic acid signaling.


Assuntos
Alternaria , Resistência à Doença , Redes Reguladoras de Genes , Doenças das Plantas , Solanum lycopersicum , Doenças das Plantas/microbiologia , Doenças das Plantas/genética , Doenças das Plantas/imunologia , Solanum lycopersicum/microbiologia , Solanum lycopersicum/genética , Solanum lycopersicum/imunologia , Alternaria/fisiologia , Alternaria/patogenicidade , Resistência à Doença/genética , Regulação da Expressão Gênica de Plantas , Transcriptoma , Reguladores de Crescimento de Plantas/metabolismo , Etilenos/metabolismo
2.
EFSA J ; 22(7): e8890, 2024 Jul.
Artigo em Inglês | MEDLINE | ID: mdl-38984216

RESUMO

The European Commission requested the EFSA Panel on Plant Health to conduct a pest categorisation of Coniella castaneicola (Ellis & Everh) Sutton, following commodity risk assessments of Acer campestre, A. palmatum, A. platanoides, A. pseudoplatanus, Quercus petraea and Q. robur plants from the UK, in which C. castaneicola was identified as a pest of possible concern to the EU. When first described, Coniella castaneicola was a clearly defined fungus of the family Schizoparmaceae, but due to lack of a curated type-derived DNA sequence, current identification based only on DNA sequence is uncertain and taxa previously reported to be this fungus based on molecular identification must be confirmed. The uncertainty on the reported identification of this species translates into uncertainty on all the sections of this categorisation. The fungus has been reported on several plant species associated with leaf spots, leaf blights and fruit rots, and as an endophyte in asymptomatic plants. The species is reported from North and South America, Africa, Asia, non-EU Europe and Oceania. Coniella castaneicola is not known to occur in the EU. However, there is a key uncertainty on its presence and geographical distribution worldwide and in the EU due to its endophytic nature, the lack of systematic surveys and possible misidentifications. Coniella castaneicola is not included in Commission Implementing Regulation (EU) 2019/2072 and there are no interceptions in the EU. Plants for planting, fresh fruits and soil and other growing media associated with infected plant debris are the main pathways for its entry into the EU. Host availability and climate suitability in parts of the EU are favourable for the establishment and spread of the fungus. Based on the scarce information available, the introduction and spread of C. castaneicola in the EU is not expected to cause substantial impacts, with a key uncertainty. Phytosanitary measures are available to prevent its introduction and spread in the EU. Because of lack of documented impacts, Coniella castaneicola does not satisfy all the criteria that are within the remit of EFSA to assess for this species to be regarded as potential Union quarantine pest.

3.
Plant Dis ; 2024 Jul 15.
Artigo em Inglês | MEDLINE | ID: mdl-39010655

RESUMO

Melon (Cucumis melo L.) is an economically important crop in Brazil, with an annual production of 699.281 tons (FAO 2024). Fungal diseases are one of the biggest problems in melon production, and melon growers in northeastern Brazil have reported over 80% of plants showing anthracnose symptoms in the fields during rainy seasons. Plants were wilted, displaying brown necrotic lesions and water-soaked spots with yellowish edges on the leaves and vines. Melon fruits displayed necrotic lesions on the outside. From June 2022 to June 2023, melon leaves (varieties Yellow, Galia, and Cantaloupe) from anthracnose-symptomatic plants were collected in four melon farms located in the municipalities of Afonso Bezerra, Mossoró, Tibau, and Upanema in the state of Rio Grande do Norte. Small fragments of symptomatic leaves were disinfected in 70% ethanol (30 sec) and 2.5 % sodium hypochlorite (1 min), rinsed in sterile distilled water, and plated on PDA Petri dishes with tetracycline (0.05g/liter). Plates were maintained in a bio-oxygen demand incubator (BOD) for 3 days at 28 ± 2 °C, under a 12 hr photoperiod. Eleven representative fungal colonies resembling Colletotrichum spp. were selected and monosporically grown on PDA for seven days for morphology, pathogenicity, and molecular analyses.ight colonies showed pinkish-dark brown with acervuli in the center and cottony mycelium, and showing black edges in some isolates, resembling C. plurivorum (Zhang et al. 2023). Conidia from those colonies were hyaline, cylindrical with obtuse ends, and 17.76 x 7.06 µm, n= 50. Three colonies developed pinkish-gray mycelia with numerous black microsclerotia, and the conidia were hyaline, falcate, and 27.38 x 4.10 µm, n= 50, resembling C. truncatum (Yu et al. 2023). The total DNA of the eleven isolates was extracted, and the internal transcribed space (ITS), glyceraldehyde 3-phosphate dehydrogenase (GAPDH), actin (ACT), ß-tubulin (TUB), and chitin synthase 1 (CHS-1) regions were partially amplified by PCR. Amplicons were sequenced and deposited to Genbank (Table eXtra1). A phylogenetic tree was built with the Maximum likelihood method with the concatenated sequences of the five partial gene sequences on Software MEGA (Version 11.0.10) (Tamura et al. 2021). The isolates CML5, CML8, CML9, CML10, CML11, CML14, CML15, and CML25 were grouped with Colletotrichum plurivorum CBS 125474 (orchidearum complex), and the isolates CML26, CML27 and CML28 with Colletotrichum truncatum CBS 15:35 (truncatum complex) with 87 % e 97 % of Bootstrap support, respectively. C. plurivorum was detected in four farms visited (we selected two representative isolates per farm), while C. truncatum isolates were all from the farm in Afonso Bezerra municipality. A pathogenicity test was performed following the method of Baishuan et al. (2023), micro-injuries were made in leaves of melon seedlings 'Goldex Yellow' and inoculated with a spore suspension of colonies with seven days of growth (106 spore/mL) of each isolate and sprayed to the point of dripping. Sterile water was used as mock. After nine days, anthracnose symptoms similar to those observed in the field were seen in all inoculated leaves, while no symptom was observed in the leaves of the mock plants. The pathogens were reisolated and their identification was confirmed by morphology and sequencing. Five seedlings were inoculated per isolate and mock, the assay was repeated, and the same results were observed. The species C. plurivorum has already been reported to cause disease in Cucumbers in Brazil (Silva et al. 2023) and C. plurivorum and C. truncatum in Citrullus lanatus in China (Guo et al. 2022). To the best of our knowledge, this is the first report of C. plurivorum and C. truncatum causing anthracnose in melon plants in Brazil.

4.
Plant Dis ; 2024 Jul 09.
Artigo em Inglês | MEDLINE | ID: mdl-38982670

RESUMO

Amorphophallus albus P. Y. Liu & J. F. Chen is a typical cash crop widely planted in southwest China (Gao et al., 2022). In early August of 2021, a peculiar leaf spot disease was first detected on A. albus in Ankang Academy of Agricultural Sciences manufacturing base (32°69'N, 109°02'E), Shaanxi, China. Small irregular yellow-brown spots (1 to 2 mm) were observed on the surface of A. albus leaf. Following infection of the leaf, it expanded (3 to 5 mm) and became necrotic. Nine planting bases were investigated, and approximately 75% of plants were symptomatic during the rapid expansion period of bulb growth in Hanyin, Langao and Hanbin counties, Ankang City, Shaanxi, China. Higher disease incidence was observed at temperatures above 30℃ and humidity above 80%. Twenty-seven symptomatic tissues of infected leaves were first surface sterilized by immersion in 75% ethanol for 1 minute, followed by rinsing three times in sterile distilled water. The tissues were then cut into 4-5 mm pieces, plated on 1.5% potato dextrose agar (PDA), and incubated at 28±2°C. The hyphal tip from the growing edge of colonies cultured for three days at 28±2℃ was transferred to PDA to obtain pure cultures. Fungal colonies were white, then grey to black with an unevenly distributed, fast-growing aerial mycelium covering the petri dish within five days at 28±2℃. The colony turned dark brown when maintained in the dark at 28±2℃ after seven days, then grayish brown upon sporulation after 15 days (Fig.1f-g). Conidia were brown or black, smooth, spherical to sub-spherical, single-celled (8-12 µm × 10-13µm, average 9-11.5 µm in diameter, n=5µm). The nutritional hyphae exhibited septa, and a portion of the aerial hyphae formed a long, rough conidium, giving rise to a nearly spherical apical sac (Fig.1h). The surface gave rise to several small peduncles bearing clusters of surfaced spherical conidia (Fig.1i). Surfaced spherical conidia were generated on the surface of the small peduncle (Fig.1j). These morphological features were consistent with Nigrospora oryzae (Li et al., 2017). Genomic DNA was extracted from mycelia of the pathogen using an Ezup column fungal genomic DNA extraction kit (Sangon Biotech, Shanghai, China). To confirm the identity of the pathogen, the genomic fragments for the internal transcribed spacer (ITS), LSU (28S) and BenA gene of the isolate were amplified by PCR (Wang et al., 2017) and sent for sequencing. The resultant sequence (GeneBank ID of gene ITS, LSU, BenA are OR723825, OR775345, OR277316, respectively) were compared with the voucher specimens. BLAST results showed >99% identity with those of N.oryzae (GeneBank ID of N.oryzae strain LC2707 ITS, LSU, BenA are KX985954, KY806242, KY019481, respectively). A neighbor joining phylogenetic tree with the concatenated sequences of these genes showed that A-pb169 had the closest match with N. oryzae (Fig. 2). For pathogenicity testing, fifty plants in a period of rapid expansion of bulb growth were selected. Four leaves per plant were inoculated by sprayed till runoff with a conidial suspension of the pathogen (50 µL, 1×106 conidia/ml sterile water), and incubated at 30±2℃ and 80 ± 5% humidity. Control plants received sterile water. On the third day after inoculation, a yellow-brown spot appeared on leave surfaces, the spot gradually expanded; the infection rate was 90 to 95%. Fifteen days after inoculation, infected leaves showed symptoms like those observed in the field, whereas 100 control leaves sprayed with sterile water remained symptomless (Fig.1 a-e). The pathogen was reisolated from infected leaves and confirmed as N. oryzae by morphology and molecular identification. To our knowledge, this is the first report of leaf spot disease of A. albus caused by N. oryzae in China. Since its one of the major cash crops of the southeastern China, further work is necessary to determine its spread and economic impact as well as developing sustainable disease management options.

5.
BMC Microbiol ; 24(1): 255, 2024 Jul 09.
Artigo em Inglês | MEDLINE | ID: mdl-38982358

RESUMO

BACKGROUND: Alternaria alternata is the primary pathogen of potato leaf spot disease, resulting in significant potato yield losses globally. Endophytic microorganism-based biological control, especially using microorganisms from host plants, has emerged as a promising and eco-friendly approach for managing plant diseases. Therefore, this study aimed to isolate, identify and characterize the endophytic fungi from healthy potato leaves which had great antifungal activity to the potato leaf spot pathogen of A. alternata in vitro and in vivo. RESULTS: An endophytic fungal strain SD1-4 was isolated from healthy potato leaves and was identified as Talaromyces muroii through morphological and sequencing analysis. The strain SD1-4 exhibited potent antifungal activity against the potato leaf spot pathogen A. alternata Lill, with a hyphal inhibition rate of 69.19%. Microscopic and scanning electron microscope observations revealed that the strain SD1-4 grew parallel to, coiled around, shrunk and deformed the mycelia of A. alternata Lill. Additionally, the enzyme activities of chitinase and ß-1, 3-glucanase significantly increased in the hyphae of A. alternata Lill when co-cultured with the strain SD1-4, indicating severe impairment of the cell wall function of A. alternata Lill. Furthermore, the mycelial growth and conidial germination of A. alternata Lill were significantly suppressed by the aseptic filtrate of the strain SD1-4, with inhibition rates of 79.00% and 80.67%, respectively. Decrease of leaf spot disease index from 78.36 to 37.03 was also observed in potato plants treated with the strain SD1-4, along with the significantly increased plant growth characters including plant height, root length, fresh weight, dry weight, chlorophyll content and photosynthetic rate of potato seedlings. CONCLUSION: The endophyte fungus of T. muroii SD1-4 isolated from healthy potato leaves in the present study showed high biocontrol potential against potato leaf spot disease caused by A. alternata via direct parasitism or antifungal metabolites, and had positive roles in promoting potato plant growth.


Assuntos
Alternaria , Endófitos , Doenças das Plantas , Folhas de Planta , Solanum tuberosum , Talaromyces , Alternaria/crescimento & desenvolvimento , Alternaria/fisiologia , Doenças das Plantas/microbiologia , Doenças das Plantas/prevenção & controle , Solanum tuberosum/microbiologia , Talaromyces/genética , Talaromyces/crescimento & desenvolvimento , Endófitos/fisiologia , Endófitos/isolamento & purificação , Endófitos/genética , Folhas de Planta/microbiologia , Hifas/crescimento & desenvolvimento , Antibiose , Quitinases/metabolismo , Agentes de Controle Biológico , Controle Biológico de Vetores/métodos
6.
Plant Dis ; 2024 Jul 10.
Artigo em Inglês | MEDLINE | ID: mdl-38985508

RESUMO

In July 2023, a new leaf spot disease emerged on tobacco leaves in Meitan County, Guizhou Province, China (27°20'18" - 28°12'30"N, 107°15'36" - 107°41'08"E, average altitude 972 meters). Initially, the symptoms showed raised yellow-brown spots; subsequently, the lesions expanded and became broken and perforated, leading to a significant loss of economic value, the prevalence rate exceeded 30%. For isolation, two tissue fragments (0.2 × 0.2 cm) of symptomatic leaves were sterilized in 75% ethanol for 30 s, 3% NaClO for 2 min, and were washed 3 times in sterilized distilled water, and were subsequently inoculated on potato dextrose agar (PDA), and incubated at 28°C for 9 days in the dark. The two strains CW16 and CW28 were isolated using the single hyphae method (Nouri et al. 2023). Both strains formed pale to yellow white colonies on PDA. Conidia had three constricted transverse septa and 1 to 2 longitudinal septa in the central cells, with thick and hyaline conidiophores and mostly globose, pale brown conidia with slightly constricted septa, their average size were measured as 13.4-22.4×8.358-13.347 µm (n = 50). Genomic DNA was extracted from the isolated strains CW16 and CW28. The internal transcribed spacer regions 1 and 2 as well as 5.8S nuclear ribosomal RNA (ITS), large subunit nrRNA (LSU), and partial DNA-directed RNA polymerase II second largest subunit (RPB2) genes were amplified using primers (Cui et al. 2023). The sequences had been deposited in GenBank under accession numbers ITS: PP024201, PP024205; LSU: PP024207, PP024209; RPB2: PP060480, PP060481. The sequences analysis revealed a high similarity of 99.74 to 100% between strains CW16 and CW28 with P. palmicola isolate KM42 (ITS OQ875842, LSU OQ875844, RPB2 OQ883943) in GenBank. Using BLAST for homology matching, two isolates (CW16, CW28) and with the sequences of the ten type isolates from GenBank, phylogenetic analysis was conducted using the Maximum Likelihood method in MEGA (11.0) software based on ITS, LSU and RPB2 sequences, which showed that strains CW16, CW28 clustered in the same score as the Pseudopithomyces palmicola, confirming the morphological and molecular characteristics identification. The pathogenicity tests were conducted on healthy tobacco plants with 4-5 leaves (Fig. S1B), the isolated strains, CW16 and CW28, were used to inoculate the healthy tobacco leaves, while blank PDA was used as a control. All plants were maintained in a greenhouse at 28°C with a relative humidity of 90%. After 9 days, necrotic spots were observed on all tobacco leaves inoculated with CW16 and CW28 fungal plugs, while the blank PDA-inoculated tobacco leaves showed no symptoms. Based on morphological and molecular characteristics, the same pathogen P. palmicola was identified from the inoculated leaves, fulfilling Koch's postulates. This study represents the first reported of tobacco leaf spot caused by P. palmicola in China and provides a theoretical basis for future prevention and control measures.

7.
Plant Dis ; 2024 Jul 17.
Artigo em Inglês | MEDLINE | ID: mdl-39021153

RESUMO

Polygonatum kingianum Coll. et Hemsl., a Polygonatum species in the Asparagaceae family, plays an important role in Chinese herbal medicine (Zhao et al. 2018). P. kingianum is widely planted in the Southwestern China. In September 2023, we observed a leaf spot of P. kingianum with disease incidence of 100%, and disease index reached 60 in commercial plantings in Kunming, Yunnan province, China (24.3610°N, 102.3740°E). In the initial stage of infection, symptoms manifested as a small circular brown spot. As the spots gradually expanded, they formed oval to irregular shaped lesions with grayish-white or dark-brown borders. Progressively the entire leaf withered and died. For identification of the causal agent of the leaf spot, leaf sections (5×5 mm2) were cut from the margin of the lesion and soaked in 75% ethanol for 10 s, 1% sodium hypochlorite for 3 min, washed with sterile distilled water, dried on sterilized tissue paper and placed on potato dextrose agar (PDA). The Petri dishes were then incubated at 28℃ for 3 days with a 12-h photoperiod. A predominant fungus was isolated from 95% of the samples. Three monosporic isolates were screened using a single-spore isolation method. After 4 days of incubation the colonies were white, after 7 days turned yellow-white. Conidia were black-brown, oblong or fusiform, with 3-7 transverse septa and 0-3 longitudinal septa, with dimensions of 19.5 to 49.5 × 8.7 to 17.6 µm (n = 30). Total genomic DNA of these three isolates was extracted from mycelia by the cetyltrimethylammonium bromide (CTAB) protocol. The nucleotide sequences of the elongation factor 1-alpha (EF1α), nuclear ribosomal internal transcribed spacer (ITS), 28S nuclear ribosomal large subunit rRNA gene (LSU), 18S nuclear ribosomal small subunit rRNA gene (SSU), and the second largest subunit of nuclear DNA-directed RNA polymerase II (RPB2) gene regions were amplified using the primer pairs EF1-728F/EF1-986R (Carbone and Kohn 1999), ITS1/ITS4 (White et al. 1990), LR0R/LR5 (Schoch et al. 2012), NS1/NS4 (Schoch et al. 2012), and fRPB2-5F/fRPB2-7Cr (Liu et al. 1999), respectively. Amplicons were cloned in a pMDTM19-T vector (code no. 6013, Takara, Kusatsu, Japan) and bidirectionally sequenced. All three isolates had identical nucleotide sequences. Sequences from one isolate (PkF03) were deposited in GenBank. BLASTn analyses showed that sequences of EF1α (GenBank accession no. PP695240), ITS (PP694046), LSU (PP683406), SSU (PP683407), and RPB2 (PP695241) of isolate PkF03 were 99.6 (KP125134), 100 (KP124358), 100 (KP124510), 99.9 (KP124980), and 100% (KP124826), respectively, identical with Alternaria alternata (Fr.) Keissl. strain CBS 118815. Based on the nucleotide sequences of EF1α, ITS, LSU, SSU, and RPB2, a maximum likelihood phylogenetic tree was constructed using MEGAX with Tamura-Nei model. Isolate PkF03 was grouped in the same clade as A. alternata. According to the morphology and sequence analyses isolate PkF03 was identified as A. alternata (Woudenberg et al. 2013). To determine pathogenicity of isolate PkF03, a spore suspension (106 spores/mL) was sprayed on 1-year-old healthy leaves of P. kingianum. The control leaves were sprayed with sterile water. All plants were incubated at 28℃, 70% relative humidity, and a 12-h photoperiod. The pathogenicity tests were repeated three times with six plants in each treatment. Fifteen days post-inoculation, the inoculated leaves showed brown-yellow lesions, whereas the control leaves remained symptomless. A. alternata was reisolated from infected leaves. To our knowledge, this is the first report of A. alternata causing leaf spot on P. kingianum in Kunming, China. The results provide a scientific basis for prevention and control of the disease.

8.
Plant Dis ; 2024 Jun 11.
Artigo em Inglês | MEDLINE | ID: mdl-38861469

RESUMO

Mung bean (Vigna radiata (L.) R. Wilczek) is a legume with high nutritional and economic value that is cultivated widely across Asia (Kang et al. 2014). In March 2022, a leaf spot disease in mung bean was observed at the Gangneung-Wonju National University Experimental farm (Gangneung, South Korea, 37.77°N, 128.86°E). The affected plants had irregular brown-gray leaf spots, and the bottom of the leaves showed concentric brown-gray rings that eventually progressed to necrotic lesions. Regardless of the cultivar, approximately 30% of the plants in the field were infected. To isolate the pathogen, the affected leaves were surface-sterilized by washing with 70% ethanol for 1 min, followed by washing with 2% NaClO for 2 min, then rinsing with sterile distilled water. We placed 3-mm sized diseased lesions on potato-dextrose agar (PDA), then incubated them at 27 ± 1 °C in the dark for 7 days and we obtained 1 isolate (CC1). The fungus on PDA had white aerial mycelia that became gray toward the center. Single spore cultures were obtained from fungal isolate. Isolated conidia were single celled, hyaline, cylindrical, and measured between 20.75 to 22.07 µm × 5.85 to 6.92 µm (n = 20), similar to other reports of C. camelliae(Wang et al. 2016). Mycelium from the single spore isolate was used for DNA extraction using Exgene™ Plant SV / (GeneAll®, Cat.No. 117-152), and we amplified the ITS region with primers ITS1 + ITS2 and ITS3 + ITS4, a portion of the actin gene with primers ACT-512F + 738R, and a portion of the beta-tubulin gene with primers BT2aF + BT2bR (Carbone et al. 1999; Glass et al. 1995; White et al. 1990). The amplified regions were sequenced by by Macrogen® (Seoul, South Korea). Sequences were deposited under GenBank accession numbers OR523262 (ITS), OR582483 (Actin), and OR566953 (beta-tubulin). MegaBLAST analysis of the ITS1, ITS2, ACT, and TUB regions showed 99% (216/217 bp) similarity with C. camelliae strain HNCS-26 (MK041440.1), 99% (303/305 bp) similarity with C. camelliae strain G3 (ON025203.1), 99% (242/244 bp) similarity with C. camelliae strain FWT41 (MN525820.1), and 99% (456/460 bp) with C. camelliae strain LF152 (KJ955239.1), respectively. To fulfill Koch's postulates, we conducted a pathogenicity teston the mung bean cultivar VC1973A (Seonhwanokdu) grown for four weeks at 25 °C with a 16-h day/8-h night cycle, simulating the field conditions when the symptoms were observed. We tested the pathogenicity on six plants , three control and three infected plants. Using three leaf replicates per plant resulting in total of nine leaves per group. Leaves were first injured using a sterile needle then either sterile 5 mm PDA plugs or plugs with C. camelliae were placed on the leaf for control and infected conditions, respectively. Irregular gray leaf spots were observed on the abaxial and adaxial of the infected leaf after 2 weeks, like the symptoms observed in the field. This was observed only on infected leaves and nowhere else on the plant and the control plants had no infection. We re-isolated the pathogen from diseased leaves and identified it as C. camelliae using the same molecular markers described previously, completing Koch's postulate. To the best of our knowledge, this is the first report of leaf spot caused by C. camelliae in mung bean plants in Korea, previously the fungi was reported to infect tea plants in Korea (Hassan et al. 2023). More studies are required to investigate potentially resistant mung bean varieties to minimize future damage caused by this fungus.

9.
Plant Dis ; 2024 Jun 09.
Artigo em Inglês | MEDLINE | ID: mdl-38853332

RESUMO

Nanhaia speciosa, commonly known as Niudali, is a medicinal woody vine belonging to the Leguminosae family. Valued for its culinary and medicinal properties, it is extensively cultivated, covering approximately 5,973 hm2 in the Guangxi Zhuang Autonomous Region of China. The edible tubers of this plant are reported to possess antibacterial and antioxidant effects (Luo et al., 2023; Shu et al., 2020). In July 2021, a Niudali plantation in Yulin, Guangxi, China (22°64'N; 110°29'E) exhibited leaf spot symptoms, with an incidence rate exceeding 40% across a 46,690 m2 area. Initially, small circular, pale yellow spots appeared on the leaves, which subsequently evolved into dark brown lesions surrounded by yellow halos, ultimately leading to foliage wilting. Leaves exhibiting typical symptoms were collected for pathogen investigation. The leaves were thoroughly washed with sterile water and small tissue fragments (5×5 mm) were excised from the lesion periphery. These fragments were surface-sterilized with 75% ethanol and 1% NaClO, rinsed three times with sterile water, and subsequently cultured on potato dextrose agar (PDA) at 28 °C in darkness for 7 days. Through single-spore isolation, seven isolates with similar morphological traits were obtained. After 7 days of incubation on PDA at 28 °C in dark, the colonies exhibited a white to grey coloration on the upper surface with abundant aerial hyphae, while the underside appeared dark black. The conidia, cylindrical or obclavate in shape, were straight, pale brown, and measured 30.1-128.9 µm × 4.8-15.0 µm (n=50). The morphological characteristics matched those of Corynespora sp.(Wang et al. 2021). For molecular identification, the isolate N5-2 underwent DNA sequence analysis using genomic DNA and primers ITS1/ITS4 and EF1-688F/EF1-1251R. The sequences (ITS: OP550425; TEF1-α: OQ117118) were deposited in GenBank, exhibiting 98% identity to C. cassiicola (OP981637) for TEF1-α and 99% homology to C. cassiicola (OP957070) for ITS. Based on the concatenated ITS and TEF1-α, a maximum likelihood phylogenetic analyses using MEGA7.0 clustered the isolate with C. cassiicola. Consequently, the fungus was identified as C. cassiicola based on its morphological and molecular features. In the pathogenicity test on 1-year-old Nanhaia speciosa seedlings, leaves were gently scratched and inoculated with mycelial plugs (5 mm). Control seedlings received PDA plugs. Five leaves per plant and five plants per treatment were selected for assessment. All seedling were maintained in a greenhouse (12/12h light/dark cycle, 25 ± 2°C, 90% humidity). After a 7-day incubation period, all leaves subjected to fungal inoculation exhibited symptoms consistent with those observed in the field, while control plants remained symptom-free. The fungus was successfully reisolated from the infected leaves in three successive trials, fulfilling Koch's postulates. While C. cassiicola is well-documented for inducing leaf spots on various plant species, including Jasminum nudiflorum, Strobilanthes cusia, Acanthus ilicifolius, Syringa species (Hu et al., 2023; Liu et al., 2023; Xie et al., 2021; Wang et al., 2021), this study represents the first report of C. cassiicola causing leaf spots on Nanhaia speciosa in China. The identification of this pathogen in Nanhaia speciosa has significant implications for future epidemiological investigations and serves as a valuable reference for controlling leaf spot disease in Nanhaia speciosa.

10.
J Plant Physiol ; 299: 154277, 2024 Aug.
Artigo em Inglês | MEDLINE | ID: mdl-38843655

RESUMO

Glomerella leaf spot (GLS), caused by Colletotrichum fructicola (Cf), has been one of the main fungal diseases afflicting apple-producing areas across the world for many years, and it has led to substantial reductions in apple output and quality. HD-Zip transcription factors have been identified in several species, and they are involved in the immune response of plants to various types of biotic stress. In this study, inoculation of MdHB-7 overexpressing (MdHB-7-OE) and interference (MdHB-7-RNAi) transgenic plants with Cf revealed that MdHB-7, which encodes an HD-Zip transcription factor, adversely affects GLS resistance. The SA content and the expression of SA pathway-related genes were lower in MdHB-7-OE plants than in 'GL-3' plants; the content of ABA and the expression of ABA biosynthesis genes were higher in MdHB-7-OE plants than in 'GL-3' plants. Further analysis indicated that the content of phenolics and chitinase and ß-1, 3 glucanase activities were lower and H2O2 accumulation was higher in MdHB-7-OE plants than in 'GL-3' plants. The opposite patterns were observed in MdHB-7-RNAi apple plants. Overall, our results indicate that MdHB-7 plays a negative role in regulating defense against GLS in apple, which is likely achieved by altering the content of SA, ABA, polyphenols, the activities of defense-related enzymes, and the content of H2O2.


Assuntos
Colletotrichum , Resistência à Doença , Malus , Doenças das Plantas , Proteínas de Plantas , Fatores de Transcrição , Malus/genética , Malus/microbiologia , Malus/metabolismo , Malus/imunologia , Colletotrichum/fisiologia , Proteínas de Plantas/genética , Proteínas de Plantas/metabolismo , Resistência à Doença/genética , Doenças das Plantas/microbiologia , Doenças das Plantas/genética , Doenças das Plantas/imunologia , Fatores de Transcrição/genética , Fatores de Transcrição/metabolismo , Regulação da Expressão Gênica de Plantas , Plantas Geneticamente Modificadas/genética , Folhas de Planta/microbiologia , Folhas de Planta/metabolismo , Folhas de Planta/genética
11.
Plant Dis ; 2024 Jun 17.
Artigo em Inglês | MEDLINE | ID: mdl-38885029

RESUMO

Avocado (Persea americana), which is native to Latin America, is mostly planted in southwest China. In November 2021, leaf spot symptoms were observed in a nursery in Chongzuo (22.2019°N, 106.4723°E), Guangxi, China. Approximately 90% of avocado seedlings in the nursery were affected. Symptomatic plant fully expanded leaves showed small brown spots that ranged from 1 to 3 mm, with a yellow halo around (Fig.1). Lesions gradually expanded and became nearly round and dark brown. Finally, leaves withered or curled. For pathogen isolation, 15 symptomatic leaves were randomly sampled from different plants of the nursery, five leaves were selected and four samples size 4×4mm were taken from each leaf and were plated on potato glucose agar. Identical fungus colonies were observed in 80% of the samples, and no bacteria were isolated. Single conidial isolation was performed. After 4 days, the colony diameter reached 74.6 mm, colonies appeared gray, and developed aerial hyphae. Conidiophores were mostly solitary with a few clustered erect or slightly curved, knee shaped, and 3.89 to 5.24 µm wide. Conidia were 39.33 -96.88 × 9.96 - 15.59 µm, slightly curved, rarely straight, light brown to yellowish brown, fusoid or navicular, and truncated at the base with 4 to 10 septa. Based on morphological and cultural characteristics, the fungus was identified as Bipolaris sp. (Manamgoda et al. 2014). An isolate named MP211122 was grown on Sachs' ager at 27℃ under 12-h light/dark for 1 week and consistently with Adhikari et al. (2021) no sexual from was observed. To confirm the tentative identification, genomic DNA was extracted, ITS and GAPDH gene were amplified and sequenced using primers ITS1/ITS4 and GPD/GPD2, respectively (Tan et al. 2022). The ITS sequence (GenBank ON248469) shared 100% identity with B. setariae (MN215632.1), and the GAPDH sequence (ON642344) shared 99.82% identity with B. setariae (MF490833.1, MK144540.1) and B. yamadae (MK026428.1). A maximum likelihood phylogenetic analysis based on GAPDH and ITS sequences using MEGA 7.0 revealed that the isolate clustered with B. setariae with 100% bootstrap support(Fig. 2). Healthy 11-month old potted avocado seedlings from disease-free nursery were selected , the conidial suspension (1 × 105 conidia/mL) of MP211122 isolate was prepared by harvesting conidia from a 10-day-old culture on water agar. Conidia were sprayed onto young leaves of six potted plants. Three additional seedlings sprayed with sterile distilled water served as controls. All plants were covered with plastic bags for 3 days to maintain high humidity and then maintained in a greenhouse at 30℃ with a 12-h/12-h light/dark cycle. After 5 days, typical symptoms of small brown spots were observed on all inoculated leaves (Fig.3). All leaves on control plants remained asymptomatic. The reisolated fungus was morphologically identical to the original isolate used for inoculation, fulfilling Koch's postulates. This is the first report of B. setariae as a pathogen causing leaf spot on avocado in China. This information will facilitate further studies, monitoring and control of the disease as accurate identification of the causal agent is a primary requisite for designing management strategies.

12.
Plant Dis ; 2024 Jun 20.
Artigo em Inglês | MEDLINE | ID: mdl-38902877

RESUMO

Agave sisalana, as an excellent fiber producing plant, is mainly planted in Guangxi Province, China. In November 2023, a foliar disease occured on A. sisalana at Liangjiang Town (108.3593 W, 23.4723 N), Wuming District, Nanning in GuangXi, China. Approximately 50 to 60% of the plants (n=200) had obvious leaf spots on more than 70% of the leaves. On the leaves of sisal, circular or irregularly shaped yellow brown spots can be seen, sunken, with no halo on the edges. As time goes on, the lesion gradually expands to the entire blade of the sword (Figure 1A, 1B). To identify the disease etiology, ten agave leaves were collected from GuangXi. Symptomatic midribs were cut into 3×3 mm pieces, surface sterilized with 75 % ethanol for 20 s, rinsed with sterilized distilled water three times, air dried on sterile filter paper, plated on photo dextrose agar (PDA) medium, and incubated at 28 ℃ in the dark. Five isolates (JM01, JM02, JM03, JM05, JM06) with similar morphology were obtained. Colonies on PDA medium were white to grayish-white with atrial mycelia growing initially upward and then forming clusters (Figure 1E). After five days, mycelia turned grayish black. Immature conidia were initially hyaline, aseptate, and ellipsoid. Mature conidia were dark brown, one septate, longitudinal striate, and 22.1 to 26.3×10.2 to 14.9 µm (Figure 1F). Morphologically , the isolates were identified as Lasiodiplodia theobromae (Alves et al. 2008). For molecular identification, genome DNA of five representative isolate was extracted using the Fungi Genomic DNA Purification kit. The internal transcribed spacer (ITS) region of rDNA and translation elongation factor 1-alpha (TEF-1α) and ß-tublin (TUB) gene were amplified with primer pairs ITS1/ITS4 (White et al. 1990), EF1-728F/EF1-986R (Carbone and Kohn 1999), and Bt2a/Bt2b (Glass and Donaldson 1995), respectively, and sequenced. The ITS (PP209594), TEF-1α (PP234629), and TUB (PP234628) sequences of representative isolate JM01 were deposited in GeneBank. BLAST searches showed >99% nucleotide identity to sequences of L. theobromae (ITS, 99.26% to NR111174; TEF-1α, 99.69% to MM840490; TUB, 98.92% to MN172230). Phylogenetic analysis using maximum likelihood based on the combined ITS, TEF-1α, and TUB sequences of the isolates and reference sequences of Lasiodiplodias spp. from GenBank indicated the isolates obtained in this study formed a clade strongly supported based on bootstrap values to the ex-type isolate CBS164.96 sequences of L.theobromae (Figure 2). To test pathogenicity, JM01 was tested by inoculation leaves of one year old agave plants, the epidermis at the inoculation site, 10, 15 and 20 cm below to the crown, was wiped with a 75% alcohol cotton ball, washed three times with sterile water, and punctured (5 mm diameter) with a sterile inoculation needle. A 5 mm block of each isolate cultured on PDA for 3 days was attached to the inoculation site. Controls were inoculated with sterile PDA. The inoculation area was covered with plastic wrap. All plants were kept in a controlled greenhouse at 27℃, 80% relative humidity, and natural daylight, and watered weekly. Each treatment was repeated three times. Remove the block one day later. Three days after inoculation, all inoculated had typical symptoms,but control were healthy (Figure 1C, 1D). Fungal isolates were only recovered from symptomatic stems and were morphologically identical to L. theobromae, completing Koch's postulates. L. Theobromae has been reported as the cause of leaf rot on A. angustifolia in Mexico (Reyes-García et al. 2023). To our knowledge, this is the first report of L. theobromae causing leaf spot on A. sisalana in GuangXi, China. L. theobromae is primarily a plant pathogen that causes rotting and dieback in fruits and plants in tropical and subtropical regions (Puttanna 1967). This study is useful to focus on management strategies for leaf rot disease by L. theobromae of A. sisalana.

13.
Plant Dis ; 2024 Jun 20.
Artigo em Inglês | MEDLINE | ID: mdl-38902881

RESUMO

Saposhnicovia divaricata (Trucz.) Schischk. is one of the traditional medicinal herbs in northeast China, and its roots are used for medicinal purposes. In 2020, a fungus isolated from S. divaricata seeds was observed to cause root rot of seedlings, leaf spot and stem spot of adult plants in Shuangyashan, Heilongjiang, China. Based on morphological and molecular data, isolates of all fungi were identified as Alternaria alternata. To our knowledge, this is the first report of A. alternata isolated from S. divaricata seeds in China. The carrying rate of S. divaricata seeds from 20 different collection sites reached 100% in 70% of the sites in Hulunbeier area, Inner Mongolia, China. The A. alternata isolate could infect the roots of cucumber, sorghum, mung bean and maize seedlings and cause root rot. Considering the control of seed-associated fungal diseases, prochloraz 45% EW had the best control effect of 92.6%, followed by flusilazole 400 g L-1 EC (88.9%) and azoxystrobin·propiconazole 18.7% SE (70.7%) of 15 fungicides. Further field control efficacy showed that 45% prochloraz EW had an 80% control efficacy on the disease at a dose of 0.225 g L-1. It is recommended that soaking seeds and spraying are the best treatments for controlling seed-associated fungi and leaf spot on S. divaricata caused by A. alternata. Therefore, above methods can effectively prevent the occurrence of fungal diseases of S. divaricata and provide a method to reduce reinfestation in the field.

14.
Plant Dis ; 2024 Jun 25.
Artigo em Inglês | MEDLINE | ID: mdl-38916907

RESUMO

The Chinese quince (Chaenomeles sinensis (Thouin) Koehne), belongs to the Rosaceae family, is widely distributed throughout Asia, including Republic of Korea. It is used as a traditional treatment for asthma, common cold, and dry pharynx. Numerous recent pharmacological studies on antiinfluenza, antioxidant, and antidiabetic properties have confirmed the medicinal properties of the Chinese quince fruit (Chun et al., 2012). In March 2022, leaf spots on Chinese quince, resulting in defoliation, were observed in Andong, Gyeongsangbuk Province, Korea (Fig. 1A). The disease symptoms are dark brown spots on leaves. Later, the chlorophyll is lost, causing the entire leaf to become wilted and fell off (Fig. 1B). To identify the pathogen, symptomatic leaves were brought to the laboratory, cut into small pieces, and surface-disinfected in 70% ethanol for 15 s and rinsed with sterile distilled water (SDW). The specimens were then treated with 1% NaOCl for 15 s, followed by rinsing with SDW. Thus, surface-disinfected tissues were placed onto potato dextrose agar (PDA) plates and incubated at 25°C for 7 d. A total of four isolates were obtained from the infected leaves. The colonies were transferred onto freshly prepared PDA plates by the single spore method for further purification. GYUN-10746 isolate was selected as the representative strain among the four isolates and deposited in the Korean Agricultural Culture Collection (KACC 410367). They initially produced white mycelia, which turned dark brown or pale brown at the center and beige at the periphery after 7 d (Fig. 1C and D). Conidiophores were pyriform, sometimes ovoid, or ellipsoidal and brown, measuring 30.8 ± 0.49 × 12.9 ± 0.26 µm (length × width) (n=100) (Fig. 1E). The morphological characteristics were consistent with those of Alternaria alternata (Woudenberg et al. 2015). For molecular identification, DNA was amplified using the following primers: ITS1/ITS4 (White et al. 1990), EF1-728F/EF1-986R (Carbone et al. 1999), Gpd-R/Gpd-F (Berbee et al. 1999), Alt a1-F/Alt a1-R (Hong et al. 2005) and rpb2F/rpb2R (Liu et al. 1999) by PCR. DNA sequences from all 4 isolates (GYUN-10746, GYUN-11193, GYUN-11194 and GYUN-11195) were identical. The ITS (OP594615), TEF1-α (OR327062), GAPDH (OR372157), Alt a 1 (OR327061), and RPB2 (OR352741) sequences from the representative isolate GYUN-10746 were 100% identical to those of previously identified A. alternate isolates. A phylogenetic tree was constructed using sequences of ITS, TEF1-α, GAPDH, Alt a l, and RPB2 to illustrate their relationship with A. alternata and related Alternaria species (Fig. 2). For the pathogenicity test, healthy Chinese quince branch containing leaves were inoculated with 7-day-old mycelial plugs of A. alternata, while leaves on a branch inoculated with PDA plugs alone served as a control group. Thus inoculated branches were incubated at 25°C for 7 d. Disease symptoms were developed on leaves of the branches inoculated with mycelial plugs of the fungal pathogen (Fig. 1F), while no symptoms developed on control group. The resulting leaf spots resembled those on the original infected plants. To confirm Koch's postulates, the pathogen was re-isolated from inoculated leaves with identical morphological and molecular characteristics. To the best of our knowledge, this is the first report of leaf spot caused by A. alternata in C. sinensis in Korea. The identification of the pathogen may provide pertinent information for the development of disease controlling strategies.

15.
Plant Dis ; 2024 May 27.
Artigo em Inglês | MEDLINE | ID: mdl-38803070

RESUMO

Polygonatum kingianum is a Chinese herbal medicine that belongs to the genus Polygonatum of the family Liliaceae. In June 2023, Polygonatum kingianum Coll. et Hemsl. in nurseries in Qujing, Yunnan Province, China, showed irregular brown spots on the leaves, whole leaf necrosis, and plant death in serious cases, with an incidence of 10-20% (Fig. S1). To identify the pathogens of P. kingianum, six diseased samples were collected from nurseries with 0.6 acre. These diseased sample leaves were soaked in 0.1% HgCl2 for 1 min and 75% ethanol for 2 min and then rinsed thrice with sterile water. Treated leaves were cut into small pieces (5×5 mm) and cultured on potato dextrose agar (PDA) for five days at 28°C. Total thirteen fungal strains were isolated from PDA medium. The nuclear ribosomal internal transcribed spacer of ribosomal DNA (ITS rDNA) region of these 13 strains was amplified by polymerase chain reaction (PCR) using universal primers ITSI/ITS4 (White et al. 1990). Sequencing and BLAST of the ITS region on NCBI showed that 11 out of 13 fungal strains belonged to the genus Alternaria, with an identity ≥99%. We selected one of the Alternaria strains, HJ-A1, for further study. The HJ-A1 colony appeared grayish brown white-to-gray with a flocculent texture on the front side and a dark gray underside on the PDA medium (Fig. S1). The conidiophores appeared brown, either single or branched, and produced numerous short conidial chains. The conidia were obclavate to obpyriform or ellipsoid in shape and contained 1-4 transverse septa and 0-2 oblique septa. The conidial diameter was 27.30µm in length and 12.27µm in width. (Fig. S1). To further determine the species of HJA1, the genomic DNA of HJ-A1 was extracted using the Lysis Buffer for PCR (AG, Hunan, China). Four Alternaria genomic DNA regions including the ITS, translation elongation factor 1-α gene (TEF1-α), glyceraldehyde-3-phosphate dehydrogenase (GAPDH), and Alternaria major allergen gene (Alt a1) were amplified by PCR using the primers as previously reported (Woudenberg et al. 2013, Hong et al. 2005). Sequence analysis revealed that the ITS (484bp) of HJ-A1 (NCBI No. PP082633), TEF1-α (267bp) of HJ-A1 (NCBI No. PP419893), GAPDH (582bp) of HJ-A1 (NCBI No. PP419892), and Alt a1 (522bp) of HJ-A1 (NCBI No. PP228046) shared the highest identity with A. alternata respectively (99≥%). A maximum likelihood phylogenetic tree was constructed with the combined sequence data sets of ITS, GAPDH, TEF, and Alt a1 using MEGA 7. The results showed that HJ-A1 strain clustered with A. alternate (Fig. S2). The pathogenicity of HJ-A1 was tested according to Koch's postulates by inoculating HJ-A1 conidia suspension (2×105 conidia/mL) into leaves of 1-year-old P. kingianum, with sterile water as a control. Each treatment group included 3 plants with 3 replicates. The tested plants were planted in a phytotron at 28℃ and 90% humidity. Three days after inoculation, symptoms similar to those under natural conditions were observed in the HJ-A1-inoculated plants, whereas no symptoms were observed in the control plants (Fig. S1). The same fungal strains were re-isolated from inoculated leaves and identified by morphologically and sequence of ITS. Previous studies showed that Alternaria alternata funji cause many plant diseases, such as fig fruit rot (Latinovic N et al. 2014),daylily leaf spot (Huang D et al. 2022), fruit blight on sesame (Cheng H et al. 2021),leaf spot of Cynanchum atratum Bunge (Sun H et al. 2021) and so on. To our knowledge, this is the first report of A. alternata causing P. kingianum leaf spot in China. The discovery of this pathogen will help to guide the protection and control of P. kingianum disease.

16.
Plant Dis ; 2024 May 27.
Artigo em Inglês | MEDLINE | ID: mdl-38803066

RESUMO

Rosa roxburghii Tratt is a plant from the Rosaceae family whose fruits are rich in vitamins, dietary fiber, flavonoids, phenolic acids, and other active components (Jiang, et al. 2024). In July 2023, about R. roxburghii 500 plants were investigated in a field of 6000 m2 in Guiding County (107°14'E, 26°45'N), Guizhou province, China, and the results showed a leaf spot incidence of s 20 to 30%. . The affected leaves had irregular, black lesions with a clear blackish brown boundary and faint black conidiomata in a brown center. Fifteen symptomatic leaves were collected from 10 plants washed with sterile distilled water, and 5 × 5 mm pieces of the infected tissues were cut. After surface sterilization for30 s with 75% ethanol, 2 min with 3% NaOCl, three washes in sterilized distilled water, the leaf pieces were dried and placed on potato dextrose agar (PDA) and incubated at 25℃ for 5 days. Three isolates (H3-Y-1-1, H3-Y-1-2, H3-Y-1-3) with identical morphology were obtained, and the isolate H3-Y-1-1was selected for further study. The colonies on PDA exhibited irregular growth patterns, with white felty aerial mycelium on the upper surface, and white mycelium on the lower surface. Conidiomata were irregularly distributed over the agar surface. The isolate H3-Y-1-1 produced darkly pigmented pycnidia on PDA after 30 days and oozed milky mucilaginous drops. The fungus produced two types of conidia, α and ß. Regular α conidia were 4.74 - 5.96 × 1.52 - 2.24 µm (n = 50), hyaline, elongated, biguttulate and non-septate. Beta conidia were 20.13 - 25.74 × 0.86 - 1.29 µm (n = 50), aseptate, hyaline, smooth, spindle shaped, slightly curved to bent. The morphological features were consistent with the description of Diaporthe eres (Pereira, et al. 2022). The pathogen was confirmed to be D. eres by amplification and sequencing of the internal transcribed spacer region (ITS), the partial ß-tubulin (TUB), the partial translation elongation factor 1-alpha (TEF) genes using primers ITS1/ITS4, Bt-2a/Bt-2b, EF1-728F/EF1-986R, respectively. Sequences from PCR amplification were deposited in GenBank with accession numbers PP411998 (ITS), PP502153 (TUB), PP502156 (TEF). BLAST searches of the sequences revealed (96%) (500/523nt), 97% (479/494 nt) and 99% (334/338 nt) homology with those of D. eres CBS 138594 from GenBank (OM698848, OM752196 and OM752197), respectively. Phylogenetic analysis using maximum-likelihood and Bayesian methods placed the isolate H3-Y-1-1 in a well-supported cluster with D. eres CBS 101742. The pathogen was thus identified as D. eres based on the morphological characterization and molecular analyses (Feng, et al. 2013; Tao, et al. 2020). To assess its pathogenicity, healthy R. roxburghii potted plants were inoculated with H3-Y-1-1 spore suspensions. Symptomatic leaves mirroring field symptoms were observed after XX days of incubations at XX°C, while control plants exhibited no symptoms. Diaporthe eres was consistently reisolated from the infected leaves showing brown irregular or round lesions at the initial stage of the disease, expanding and become more irregular over time ultimately causing leaf curling and plant death. To our knowledge, this is the first report of leaf spot on R. roxburghii caused by D. eres in China. The disease may become a serious threat to fruit of R. roxburghii production in China. Therefore, detection of this pathogen is very important to ensure timely disease management.

17.
Arch Microbiol ; 206(6): 284, 2024 May 30.
Artigo em Inglês | MEDLINE | ID: mdl-38814366

RESUMO

The tea plant, Camellia sinensis [L.] O. Kuntze, is a vital global agricultural commodity, yet faces challenges from fungal infections, which affects its production. To reduce the loss in the tea production, the fungal infections must be removed which is managed with fungicides, which are harmful to the environment. Leaf necrosis, which decreases tea quality and quantity, was investigated across Assam, revealing Lasiodiplodia theobromae as the causative agent. Pathogenicity tests, alongside morphological and molecular analyses, confirmed its role in leaf necrosis. Genome and gene analysis of L. theobromae showed multiple genes related to its pathogenicity. The study also assessed the impact of chemical pesticides on this pathogen. Additionally, the findings in this study highlight the significance of re-assessing management approaches in considering the fungal infection in tea.


Assuntos
Ascomicetos , Camellia sinensis , Doenças das Plantas , Folhas de Planta , Camellia sinensis/microbiologia , Ascomicetos/genética , Ascomicetos/isolamento & purificação , Doenças das Plantas/microbiologia , Índia , Folhas de Planta/microbiologia , Fungicidas Industriais/farmacologia
18.
Plant Dis ; 2024 May 02.
Artigo em Inglês | MEDLINE | ID: mdl-38698523

RESUMO

Photinia × fraseri is a well-known ornamental shrub in southern China. In December 2021, we observed leaf spots that were circular to irregular, gray with dark red margins and violet brown with brownish violet edges on the leaves of Photinia × fraseri shrubs in the scenic area of Shenlongtan (28°46'10″N, 115°42'93″E), Jiangxi Province, China. Almost 15% of the leaves in the 1300 m2 Photinia × fraseri planting area were symptomatic. Thirty symptomatic leaves were randomly collected from different plants, and sectioned into 5-mm2 pieces, which were surface-sterilized using 1% NaOCl for 30 s. After rinsing thrice in sterile distilled water and drying, the pieces were transferred onto potato dextrose agar (PDA) and incubated at 28 ℃ for 5-7 days. A total of sixteen morphologically similar isolates were obtained. After incubation on PDA for 20 days, the fungi had irregular edges, were white to pale brown, and had spare aerial mycelium on the surface with irregularly distributed black, gregarious conidiomata. Conidia were fusoid, subcylindrical, straight to slightly curved, 4-septated, slightly constricted at the septa, and 23 to 36 × 6 to 10 µm (mean: 27.6 × 7.7 µm). The morphological characteristics were consistent with the features of Pseudopestalotiopsis species (Maharachchikumbura et al. 2014). The genomic DNA of two representative isolates (JFRL032 and JFRL033) was extracted for further identification. The internal transcribed spacer (ITS) region, translation elongation factor 1-ɑ (tef1-ɑ) and ß-tubulin (tub2) genes were amplified and sequenced using primers ITS5/ITS4, EF1-526F/EF1-1567R, and Bt2A/Bt2B, respectively (Maharachchikumbura et al. 2012). The sequences of the two representative isolates were 100% identical to each other. These nucleotide sequences were deposited in GenBank with accession numbers, ON342794 and ON342795 (ITS); ON375851 and ON375852 (tef1-ɑ); ON375853 and ON375854 (tub2). BLASTn searches of the obtained sequences revealed 99%-100% to ITS (MG816316, 478/478 nucleotides), tef1-ɑ (MG816336, 924/926 nucleotides), tub2 (MG816326, 441/442 nucleotides) sequences of the ex-type strain of Pseudopestalotiopsis ixorae (NTUCC17-001.1). Phylogenetic analysis was conducted using the concatenation of multiple sequences (ITS, tef1-ɑ and tub2) with the Maximum likelihood statistics in PhyloSuite v1.2.2 (Zhang et al.2020). The phylogenetic tree showed the two isolates clustered with P. ixorae in a clade with 100% bootstrap support. The isolates were identified as P. ixorae based on morphological and molecular data. To confirm pathogenicity, eight healthy leaves of 3-year-old Photinia × fraseri were surface sterilized, scratched with a pair of sterilized tweezers, and ten µl of conidial suspension (106 conidia/ml) was sprayed on the injured leaves and the control was sprayed with sterile distilled water. Then, All plants were potted in a climate chamber at 25℃ and 85% relative humidity. After 3 days, leaf spot symptoms similar to those described above were observed on inoculated leaves, while the non-inoculated leaves remained symptomless. The pathogen was reisolated from the inoculated leaves to fulfill Koch's postulates and confirmed as P. ixorae by morphological and molecular analysis. It has been reported that P. ixorae can infect the Ixora plant (Tsai et al., 2018). To the best of our knowledge, this is the first report of P. ixorae causing leaf spot on Photinia × fraseri in China. The study provides valuable information for identifying and controlling the leaf spot on Photinia × fraseri.

19.
Plant Dis ; 2024 May 07.
Artigo em Inglês | MEDLINE | ID: mdl-38715157

RESUMO

Daphniphyllum macropodum Miq., an evergreen arbor, is widely cultivated in southern China for its ornamental and medicinal value (Su et al. 2013). In October 2019, a severe leaf spot was observed on D. macropodum in Jinggangshan National Nature Reserve in Ji'an city, Jiangxi, China (114°06'23″E, 26°32'25″N). The plants were about 15 years old, and the disease incidence was estimated to be 15% (4/26 plants). The disease primarily appeared as small black spots on the leaves. At the late stage, the spots enlarged and coalesced into regular or irregular gray necrotic lesions with dark margins. We collected five samples per plant and total 20 samples were collected to isolated the pathogen strains. The margin of the diseased tissues was cut into 5 mm × 5 mm pieces; surface disinfected with 70% ethanol and 2% NaOCl for 30 s and 60 s, respectively; and rinsed thrice with sterile water. Tissues were placed on potato dextrose agar (PDA) and incubated at 25°C in the dark. Pure cultures were obtained by single-spore isolation method, and the representative isolates, JRM3, JRM6, and JRM8 were used for morphological studies and phylogenetic analyses. The colonies of three isolates grown on PDA were white, cottony, and flocculent, contained undulate edges with dense aerial mycelium on the surface at 25 °C. Conidiomata was black conidial masses on PDA. Conidia were 5-celled, clavate to fusiform, smooth, 19.3 to 24.4 long × 6.1 to 8.6 µm wide (n = 50). The 3 median cells were dark brown to olivaceous, the central cell was darker than the other 2 cells, and the basal and apical cells were hyaline. All conidia developed one basal appendage (3.4 to 8.3 µm long; n = 50), and 2 to 3 apical appendages (18 to 32 µm long; n = 50), filiform. The morphological characteristics of the isolates are comparable with those of the genus Neopestalotiopsis (Maharachchikumbura et al. 2014). The internal transcribed spacer (ITS) regions, ß-tubulin 2 (TUB2) and translation elongation factor 1-alpha (TEF1-α) were amplified from genomic DNA for the three isolates using primers ITS1/ITS4, T1/Bt-2b, EF1-728F/EF-2 (Maharachchikumbura et al. 2014), respectively. The sequences of the isolates were submitted to GenBank (ITS, OQ372202 to OQ372204; TUB2, OQ390129 to OQ390131; TEF1-α, OQ390126 to OQ390128). A maximum likelihood and Bayesian posterior probability analyses using IQtree v. 1.6.8 and Mr. Bayes v. 3.2.6 with the concatenated sequences placed JRM3, JRM6, and JRM8 in the clade of N. clavispora. Based on the multi-locus phylogeny and morphology, three isolates were identified as N. clavispora. To confirm pathogenicity, eight healthy 10-year-old D. macropodum plants growing in the field were chosen, and 4 leaves per plant were wounded with a sterile needle and inoculated with 10 µL conidial suspension per leaf (106 conidia/ml). Eight plants inoculated with sterile water were used as control. All the inoculated leaves were covered with plastic bags to maintian a humidity environment for 2 days. The leaves inoculated with conidial suspension showed similar symptoms to those observed in the field, whereas control leaves were asymptomatic for 10 days. The same fungus were re-isolated from the lesions, whereas no fungus was isolated from control leaves. N. clavispora was determined as the pathogen of a variety of plant diseases, including Kadsura coccinea (Xie et al. 2018), Dendrobium officinale (Cao et al. 2022), Macadamia integrifolia (Santos et al. 2019). However, this is the first report of N. clavispora infecting D. macropodum in China. This work provided crucial information for epidemiologic studies and appropriate control strategies for this newly emerging disease.

20.
Sensors (Basel) ; 24(9)2024 May 01.
Artigo em Inglês | MEDLINE | ID: mdl-38733002

RESUMO

In order to efficiently identify early tea diseases, an improved YOLOv8 lesion detection method is proposed to address the challenges posed by the complex background of tea diseases, difficulty in detecting small lesions, and low recognition rate of similar phenotypic symptoms. This method focuses on detecting tea leaf blight, tea white spot, tea sooty leaf disease, and tea ring spot as the research objects. This paper presents an enhancement to the YOLOv8 network framework by introducing the Receptive Field Concentration-Based Attention Module (RFCBAM) into the backbone network to replace C2f, thereby improving feature extraction capabilities. Additionally, a mixed pooling module (Mixed Pooling SPPF, MixSPPF) is proposed to enhance information blending between features at different levels. In the neck network, the RepGFPN module replaces the C2f module to further enhance feature extraction. The Dynamic Head module is embedded in the detection head part, applying multiple attention mechanisms to improve multi-scale spatial location and multi-task perception capabilities. The inner-IoU loss function is used to replace the original CIoU, improving learning ability for small lesion samples. Furthermore, the AKConv block replaces the traditional convolution Conv block to allow for the arbitrary sampling of targets of various sizes, reducing model parameters and enhancing disease detection. the experimental results using a self-built dataset demonstrate that the enhanced YOLOv8-RMDA exhibits superior detection capabilities in detecting small target disease areas, achieving an average accuracy of 93.04% in identifying early tea lesions. When compared to Faster R-CNN, MobileNetV2, and SSD, the average precision rates of YOLOv5, YOLOv7, and YOLOv8 have shown improvements of 20.41%, 17.92%, 12.18%, 12.18%, 10.85%, 7.32%, and 5.97%, respectively. Additionally, the recall rate (R) has increased by 15.25% compared to the lowest-performing Faster R-CNN model and by 8.15% compared to the top-performing YOLOv8 model. With an FPS of 132, YOLOv8-RMDA meets the requirements for real-time detection, enabling the swift and accurate identification of early tea diseases. This advancement presents a valuable approach for enhancing the ecological tea industry in Yunnan, ensuring its healthy development.


Assuntos
Redes Neurais de Computação , Doenças das Plantas , Chá , Doenças das Plantas/microbiologia , Algoritmos , Camellia sinensis , Folhas de Planta/química
SELEÇÃO DE REFERÊNCIAS
DETALHE DA PESQUISA