Your browser doesn't support javascript.
loading
Mostrar: 20 | 50 | 100
Resultados 1 - 20 de 854
Filtrar
1.
Plant Dis ; 2024 Aug 15.
Artigo em Inglês | MEDLINE | ID: mdl-39146005

RESUMO

Stem End Rot (SER) is a devastating post-harvest disease of mango fruits causing severe losses during storage. In 22 July 2023, 31 out of 50 intact mangoes (cv. Sensation) collected from five orchards in Huaping county (26°37'N 101°15') showed typical symptoms of SER after stored for 9 d in room temperature (24-28℃). Initially, small dark brown to black spots appeared around the fruit peduncle, which rapidly expanded through the pulp tissues. The symptomatic mangoes were surface disinfected by 3% NaClO for 30 s after soaking in 75% alcohol for 3 min, and cleaned by sterile water for 3 times. Tissues were cut from the edge of lesions, dried by sterile filter paper, transferred to PDA and cultured at 28 ℃ for 5 d (Tovar-Pedraza et al., 2020). The single-spore isolation method was used to obtain pure culture. Thirty eight isolates presented four distinct kind of morphology on PDA medium. Among them, 11 isolates with same morphology were significantly distinct from common pathogens of SER. The colonies were white and pale yellow on reverse side. Mycelia grew fast and reached the edge of 90 mm Petri dish after cultured for 5d. Pycnidia were black and scattered on the mycelial mats after 15-20 d. Conidia were fusoid, straight to slightly curved, four septa, and brown. Pigmented median cells doliiform, 14.97 - 18.62(16.11 ±0.89)×5.61- 7.28 (6.61±0.51) µm. Apical cell hyaline, subcylindrical; 1-3 tubular transparent apical appendages 12.27 - 16.68 (13.65±3.78)×1.14 - 1.99 (1.59±0.36) µm. Basal cell conical with a truncate base, hyaline, and 1-2 tubulose basal appendages with 2.85 - 7.97 (5.18±1.88)×0.99 - 1.85 (1.38±0.29) µm (n=50). These fungi were described as Pestalotiopsis kenyana. based on morphological characters (Maharachchikumbura et al., 2014) which were different from isolates characterized as other common SER pathogens (Botryosphaeria, Neofusicoccum). Based on morphology, HPSX-4 was selected for further identification. ITS region, tef1-α, ß-tub of HPSX-4 were amplified and sequenced (Xun et al., 2023). The sequences were deposited in GenBank (ITS:OR889126, tef1-α:OR913431, ß-tub: OR913432). The ITS, tef1-α, ß-tub sequence of HPSX-4 showed 100% (525/525),99.59% (241/242), and 100% (742/742) identity to the P. kenyana CBS442.67 sequences (ITS: NR147549,tef1-α: KM199502, ß-tub: KM199395), respectively. HPSX-4 clustered with P. kenyana CBS 442.67 (type strain) based on maximum likelihood method by MEGA 7.0.21(Minh et al., 2013). Pathogenicity test was performed on 12 healthy mangoes (cv. Golek) by placing mycelial plugs around the peduncle and the middle of the fruit by pin-prick method according to Feng et al.(2023). Sterile PDA were used as control (three mangoes). Every inoculated fruit was incubated at 28°C, 95% ± 3% humidity with three replicates for each treatment. The experiment was repeated three times. Typical symptoms of SER were observed. There were no symptoms in the control group. The strain was reisolated and identified as P. kenyana with the method mentioned above which fulfilled Koch's postulates. This is the first report of P. kenyana causing SER disease on Mangifera indica L.. This study expands our understanding of the pathogen range of mango SER which conducive to prevent and control the SER caused by P. kenyana.

2.
Plant Dis ; 2024 Aug 15.
Artigo em Inglês | MEDLINE | ID: mdl-39146004

RESUMO

Cultivation of yellow dragon fruit (Selenicereus megalanthus) in Peru has recently expanded (Verona-Ruiz et al. 2020). In August 2021, approximately 170 of 1,110 dragon fruit cuttings (15.3%) in the university's nursery (6°26'10'' S; 77°31'25'' W) showed basal rot symptoms. Initial symptoms included small brown spots on the base of stems, expanding towards the top that became soft and watery. All symptomatic plants eventually died, i.e., a severity of 100%. The disease was more prevalent on cuttings during the rooting phase than on well-established cuttings. We collected five symptomatic cuttings from throughout the nursery. Four sections of 1 × 1 cm2 of tissue adjacent to the diseased area were excised from each cutting, immersed for 1 min in 2% NaClO, rinsed twice with sterile distilled water, placed on potato dextrose agar (PDA) medium (four sections per Petri plate, five plates), and incubated at 25°C for 7 days. Morphologically similar mycelia grew from all sections, and five monosporic isolates were obtained, one per plate. Colonies grew fast, reaching 60 to 64 mm in 7 days, and produced violet-white cottony aerial mycelia with orange sporodochia on PDA, and abundant macro- and microconidia on synthetic nutrient-poor agar. Macroconidia were straight to slightly curved, typically with 2 to 3 septa, 16.6 to 23.3 × 1.7 to 3.7 µm (n = 30); microconidia were oval or kidney-shaped, and commonly hyaline, 6.7 to 16.4 × 2.5 to 4.7 µm (n = 40). Genomic DNA was extracted from isolate AFHP-100, then the ITS region and the TEF1 and RPB2 partial genes were amplified and sequenced (Accession numbers PP977433, OR437358, PP537149) following Gardes and Bruns (1993) and O'Donnell et al. (1998). We conducted a BLASTn search of ITS sequence against the NCBI "nr" database and local 'megablast' searches of TEF1 and RPB2 sequences against FUSARIUM-ID v.3.0 (Torres-Cruz et al. 2022). We found 100%, 98.19 to 99.84%, and 98.81 to 99.76% identities in ITS, TEF1, and RPB2 sequences, respectively, to the ex-epitype and other reference strains of Fusarium oxysporum (CBS 144134, NRRL26406, among others). A maximum likelihood phylogenetic analysis with a TEF1-RPB2 concatenated dataset with FUSARIUM-ID sequences also showed isolate AFHP-100 was F. oxysporum. A pathogenicity test was carried out by inoculating wounded healthy roots of three cuttings with submersion in a 5 × 106 conidia/ml suspension for 25 min. Then, the inoculated plants were planted in sterile soil. One cutting with wounded roots submerged in sterile water served as a control. In parallel, sterile soil was inoculated with 20 mL of the conidial suspension, and another three healthy cuttings were planted. A cutting planted in noninoculated soil also served as a control. Basal rot symptoms developed in all inoculated plants after 25 days. After re-isolation, the same fungus, corroborated based on micromorphology and TEF1 sequence (PP335689), was recovered, fulfilling Koch's postulates. The isolate was deposited in the KUELAP Herbarium (voucher KUELAP-3214), located and administered by the National University Toribio Rodriguez de Mendoza de Amazonas, in Chachapoyas, Peru. Fusarium oxysporum has been reported to cause basal stem rot in Bangladesh and Argentina (Mahmud et al. 2021; Wright et al. 2007), and stem blight in Malaysia (Mohd Hafifi et al. 2019) on dragon fruit. This is the first report of F. oxysporum causing basal rot in S. megalanthus in Peru. This fungus is among the most destructive plant pathogens, and the rapid expansion of the crop in Peru requires a comprehensive knowledge of the biotic factors influencing production. Therefore, this report is foundational to implementing proper control strategies.

3.
Plant Dis ; 2024 08 15.
Artigo em Inglês | MEDLINE | ID: mdl-39148366

RESUMO

Cenchrus americanus (L.) Morrone (Poaceae), is an important millet crop cultivated mainly in arid and semiarid regions and is a staple food grain for millions of people. During 2021 July surveys in the pearl millet fields in Mysore (12°30'55" N; 76°56'54" E), Karnataka, India, plants showed spathe blight and leaf spot disease with an overall incidence ranging from 5 - 8% in the 15 hectares surveyed. Infected leaves appeared brown, and lesions extended to the sheath. Some spathes were also found infected with similar symptoms. Diseased leaves and spathes were collected (n = 5 each) for pathogen identification. Samples were cut into small pieces (0.5 cm2), sterilized with sodium hypochlorite (2%, v/v), and blotted dried. The associated fungal pathogen was isolated on potato dextrose agar (PDA) medium amended with Streptomycin (40 mg/L) and incubated at 28 ºC for 1 week. Colonies were grey, fluffy, cottony with an irregular margin, undulate and dark brown in the back of the plate. Conidiophores were pale brown, erect, slightly curved, septate, unbranched, verruculose and measured 27.1 - 94 µm in length × 2.3 - 4.5 µm in width (n = 20). Conidiogenous cells were brown, subcylindrical, irregularly shaped, and conidia were straight, mainly elliptical, dark brown smooth, with two to three septa, with measurements of 11.1 - 26.4 µm by 5.7 - 14.3 µm (n = 50). Based on morphological characters, the pathogen was identified as Curvularia sp. Two representative isolates (UOMPM1 & UOMPM2) were molecularly identified. The total genomic DNA was extracted with a CTAB method, and ITS, GAPDH and tef-1α loci were amplified using primers ITS1/ITS4 (White et al., 1990), GPD1/GPD2 (Berbee et al., 1999) and EF1983F/EF-2218R (Schoch et al., 2009) respectively. ITS sequence had 100% similarity (706/706bp) with reference sequence C. spicifera (MH863648; HF934915 & HF934916); tef-1α sequence had 100% (933/933bp) identity with C. spicifera (KM062878, KJ939505), and the GAPDH sequence was 99.8% identical to that of Curvularia sp. (MG979055), and C. spicifera (MH809681). Combined dataset of concatenated sequence (ITS-GAPDH-tef-1α) was used in a phylogenetic analysis and revealed that the isolates were in a common clade with the isolate of Curvularia spicifera (CBS 274.52) thus, confirming the identity of the isolated pathogen as C. spicifera. The sequences obtained in the present study were deposited in the GenBank (ITS: OQ253406, OQ253407; LSU: OQ253429, OQ253430; GAPDH: OQ263372, OQ263373 & TEF: OQ263374, OQ263375). Pathogenicity test was carried out by inoculating (foliar /whole plant spray) 60 healthy pearl millet plants (45-days old), grown in field plot with spore suspension (105 conidia/ml). Control plants (n=20) were treated with sterile water. The experiments were conducted in triplicates and repeated twice. Development of disease symptoms was recorded on 41 plants, and all control plants remained healthy. The identity was confirmed after re-isolation as C. spicifera based on cultural and molecular sequence analysis. To our knowledge, this is the first report of C. spicifera causing a leaf spot and spathe blight disease of pearl millet in India. This disease seriously affects grain production, and effective disease management strategies need to be investigated.

4.
Plant Dis ; 2024 Aug 08.
Artigo em Inglês | MEDLINE | ID: mdl-39115955

RESUMO

Vasconcellea x heilbornii, known as babaco, is a hybrid native to Ecuador grown in small orchards in sub-tropical highland regions. Over the last decade, several viruses have been identified in babaco using high-throughput sequencing (HTS) (Cornejo-Franco et al. 2020, (Reyes-Proaño et al. 2023). In 2021, total RNA from a babaco plant showing distinctive leaf yellowing was extracted using the PureLink RNA Mini Kit (Thermo Fischer Scientific, USA) and subjected to HTS on an Illumina NovaSeq6000 system as 150 paired-end reads (Macrogen Inc., South Korea). Library construction was done using the TruSeq Stranded Total RNA Sample kit with Plant Ribo-Zero, as described (Villamor et al. 2022). Reads were processed using BBDuk and de novo assembled using SPAdes 3.15. both implemented in Geneious 2022. Contig analysis was done by BLASTx using the NCBI viral sequence database (as of November 2022). HTS generated 54 million reads, of which 12% assembled into contigs corresponding to genomes of previously reported babaco viruses including babaco virus Q (BabVQ), babaco nucleorhabdovirus 1 (BabRV1) and babaco ilarvirus 1 (BabIV1). Interestingly, 144 reads (0.0003%) assembled into seven contigs ranging from 100 to 480 nucleotides (nt) in length. These contigs showed homology, with 97% amino acid (aa) identity (100% query coverage), to regions of the RNA-dependent-RNA-polymerase (RdRp) of beet western yellows virus (BWYV, Acc. No. NC_004756), a member of the Polerovirus genus. To confirm the occurrence of BWYV in babaco, double-stranded RNA (dsRNA) was extracted from 15 g of leaf tissue from the original sample as described (Dodds et al. 1984) and used as template for reverse-transcription (RT)-PCR using overlapping primers designed to span all short contigs. RT-PCR amplified fragments were cloned into a pGEM®T-easy vector (Promega, USA) and sequenced by the Sanger method (Macrogen Inc., South Korea). The sequences were assembled into a single 2.7 kbp BWYV genome fragment comprising the complete protein 1 (P1) and partial RdRp gene (GenBank Acc. No. PP480670). Sequence alignments between the partially sequenced genome of the babaco isolate and its corresponding fragment from the closest BWYV isolate (NC_004756) revealed 94% and 97% identities at the nt and aa levels, respectively. To assess the prevalence of BWYV in babaco, 30 leaf samples showing yellowing symptoms from Pichincha (n=15) and Azuay (n=15) provinces were tested by RT-PCR using total RNA. Total RNA extraction and reverse transcription were done using the methodology described by Halgren et al. (2007). For RT-PCR, the primer set BWYV_Bab_F: 5'-CAGTGTCCTCCAAGTGCAACAT-3' / BWYV_Bab_R: 5'GGTTCCTTCCCAGTTTGGTGGT-3', which amplifies a 235 nt-long P1 region, was used. Three RT-PCR products from each positive sample were purified using the GeneJET PCR clean-up kit (Thermo Scientific, USA) and sequenced. BWYV was confirmed in 9 out of 15 samples (60%) from Pichincha, and in 10 out of 15 samples (64%) from Azuay. Samples were also tested for additional babaco viruses as described (Reyes-Proaño et al. 2023). All BWYV-infected plants turned out positive for papaya ringspot virus (PRSV), babaco mosaic virus (BabMV), BabVQ, and BabIV1. Hence, the impact of BWYV infection on babaco plants in single and mixed infections warrants further investigation. To the best of our knowledge, this is the first report of BWYV in a crop in Ecuador, and the first time it has been found in a Caricaceae species.

5.
Plant Dis ; 2024 Aug 08.
Artigo em Inglês | MEDLINE | ID: mdl-39115951

RESUMO

Carica papaya (papaya) in Guam, USA may experience soft rot symptoms, often referred to as mushy canker disease. Disease symptoms first appear as expanding water-soaked dark-green stem lesions or leaf spotting with chlorotic halos. Defoliation at petiole-stem junctions and crown necrosis leads to plant death. Papaya diseases caused by Erwinia spp. are documented in nearby tropic regions such as the Northern Mariana Islands (Trujillo and Schroth 1982), the Philippines (Dela Cueva et al. 2017), Japan (Hanagasaki et al. 2020), and Indonesia (Suharjo et al., 2021). The pathogen was isolated from symptomatic papaya stem sections (cv. Red Lady) from a nursery at the University of Guam Agriculture and Life Sciences building in April 2023. Approximately 20% of seedlings collapsed from stem soft rot, with nearly all plants showing varying degrees of water-soaked lesions on leaves or stems. Stem tissue from lesion margins was excised, surface sterilized with 70% EtOH, and macerated in sterile water. Macerate was plated onto nutrient agar (NA) and incubated at 28°C, yielding colonies that were clear to white in color, smooth, circular and mucoid on NA plates for five suspect isolates (JGD231-235). Strains produced blue diffusible pigment on King's B (KB) media, were Gram-negative rods, and exhibited swimming motility on semi-solid (0.5% agar) NA plates. Crown stab inoculation of ten papaya plants (cv. Red Lady) with isolates resulted in mushy canker symptoms within seven days, while negative control plants stabbed with a sterile probe remained asymptomatic. Koch's postulates were fulfilled by drench-inoculating spontaneous rifampicin-resistant (100µg/ml) mutants, JGD233r and JGD235r, onto ten papaya plants (cv. Solo Sunrise). Nine days post-inoculation, bacterial strains were recovered from symptomatic stem tissue macerate plated on rifampicin (100µg/ml) NA and incubated at 28°C. No symptoms or bacterial cells were recovered from the tissue of negative control plants. Cell morphology, culture phenotypes, and disease symptoms suggested the isolates were Erwinia spp., and blue pigment production on KB further suggested E. papayae (Gardan et al. 2004). Partial 16S rDNA sequences of Guam strains JGD231-235 (sequenced using PCR forward primer 5' - AGAGTTTGATCMTGGCTCAG - 3' and reverse primer 5' - GGTTACCTTGTTACGACTT - 3', GENEWIZ (South Plainfield, NJ)) were deposited into GenBank (OR577627- 631). Highest NCBI BLAST results for all strains showed a 16S rDNA sequence identity of 98.17-98.91% with those of Erwinia sp. I-leaf (LC590218) and E. mallotivora BT-MARDI (HQ456230). A maximum likelihood phylogenetic tree based on concatenated partial atpD, infB, and rpoB sequences of strains JGD232 (PP669340, PP669346, PP669343), JGD233 (PP669341, PP669347, PP669344), and JGD235 (PP669342, PP669348, PP669345) (Brady et al. 2008) constructed using MEGA11 (Tamura et al. 2021) showed all strains formed a monophyletic group with Erwinia sp. I-leaf (Hanagasaki et al. 2020) and E. papayae NCPPB 4294T (Gardan et al. 2004), supported with 98% bootstrap. This note documents the first occurrence of E. papayae as a papaya pathogen in Guam. Papaya cultivation supports sustainable food security for Guam (Bevacqua and Sayama 2023), and Erwinia spp. pathogens threaten papaya on other Pacific islands like Hawaii. These findings convey the need for effective quarantine practices, local disease management, and further research on this pathogen.

6.
Plant Dis ; 2024 Aug 07.
Artigo em Inglês | MEDLINE | ID: mdl-39110610

RESUMO

Winter jasmine (Jasminum nudiflorum Lindl.) is a medium-sized, deciduous shrub native to China that has become a popular choice among gardeners and landscapers. In 2020 to 2021, symptoms of anthracnose including brown necrotic spots, enlarged irregular lesions and leaf blight were observed on leaves of 20 winter jasmine shrubs in a public garden (22°34'58'' N; 113°56'23'' E) in Shenzhen, China, and with an estimated disease incidence of 65%. Tissues samples (6 × 6 mm2) surrounding the necrotic spots were surface sterilized with 75% ethanol for 30 s, followed by 2% NaClO for 1 min, then rinsed with sterile water for three times and dried with sterile filter paper. Tissues were placed on potato dextrose agar (PDA) medium and incubated at 25℃. After 3 to 7 d, pure cultures were obtained by transferring hyphal tips to new plates and 32 isolates producing Colletotrichum-like colonies were obtained from 40 tissues (isolation frequency=32/(4×10)=80%). Three representative isolates YCH09, YCH23 and YCH32 were selected for further study. Three selected isolates were identical in morphological characteristics. Colonies on PDA after 5 d at 25℃ were white to gray with cottony mycelia and grayish-white on the underside of the culture. Conidia (n = 60) measured 15.4 ± 1.1 µm (13.0 to 17.1 µm) in length and 5.4 ± 0.3 µm (4.9 to 6.0 µm) in width and were hyaline, single-celled, cylindrical with rounded ends. Appressoria (n = 15) measured 7.1 ± 0.1 µm (5.3 to 8.9 µm) in length and 5.2 ± 0.2 µm (4.1 to 6.2 µm) in width and were brown to dark brown, ovoid. These morphological features were aligned with those of Colletotrichum spp. (Weir et al. 2012). Sequences of five genetic markers of representative isolates YCH09, YCH23 and YCH32 including the rDNA internal transcribed spacer region, chitin synthase, partial actin, ß-tubulin 2 and Apn2-Mat1-2 intergenic spacer and partial mating type (Mat1-2) region were 99.3 to 100% identical to the ex-type isolate of C. fructicola strain ICMP 18581 (Zhang et al., 2020). From the maximum likelihood phylogenetic tree which was constructed based on concatenated sequences, three representative isolates (YCH09, YCH23 and YCH32) were clustered with other isolates of C. fructicola. The above morphological and molecular characteristics suggest that causal agent was C. fructicola. Pathogenicity was tested using a whole-plant assay. Five healthy plants were inoculated by spraying a conidial suspension (1.5×104 conidia/ml; 20 ml per plant) of the isolate YCH23 onto the foliage (Marshall et al., 2023). Three noninoculated control plants were sprayed with sterile water. All plants were placed in a greenhouse at 25±2℃ with approximately 75% relative humidity. Yellow lesions appeared on leaves of inoculated plants as early as 4 days after inoculation (DAI), and irregularly shaped brown spots similar to those observed in the field were formed on 10 DAI. Noninoculated plants remained asymptomatic. Colletotrichum isolates resembling morphological characters of YCH23 were reisolated from all inoculated plants, then identified as C. fructicola by DNA sequence analysis. C. fructicola is a well-known fungus causing anthracnose on more than 63 plant species including agricultural and horticultural plants worldwide (Talhinhas and Baroncelli, 2021). To our knowledge, this is the first report of C. fructicola infecting J. nudiflorum plants in China. Since its potential risk to other horticultural plant species, precautions may be necessary to minimize the spread of this fungi.

7.
Plant Dis ; 2024 Aug 07.
Artigo em Inglês | MEDLINE | ID: mdl-39110608

RESUMO

Stemphylium leaf spot can result in significant losses to spinach seed, processing, and fresh market crops. Stemphylium isolates (n = 1,775) collected from 2000 to 2022 from spinach seed, leaves, and seed crop stem residues were used to assess the diversity of species associated with spinach. Eleven Stemphylium species were identified based on cmdA sequences: S. vesicarium (63.6% of isolates), S. beticola (48.9%), S. amaranthi (5.1%), S. eturmiunum (4.5%), S. astragali (4.0%), S. simmonsii (3.4%), and S. lucomagnoense, S. drummondii, S. gracilariae, S. lycopersici, and S. chrysanthemicola (each 0.6 to 1.7%). Only isolates of S. beticola, S. drummondii, and S. vesicarium were pathogenic to spinach. The incidence of spinach seed on which Stemphylium was observed ranged from 2.5 to 73.5% per seed lot, with S. vesicarium and S. beticola predominant. However, only 60.7 and 62.3% of isolates tested for these two species were pathogenic to spinach, respectively. Therefore, the incidence of Stemphylium species on spinach seed may not reflect accurately the risk of a seed lot carrying pathogenic isolates. Fused MAT1-1 and MAT1-2 genes were detected in isolates of S. vesicarium, but only MAT1-1 was detected in S. beticola isolates, which corroborates previous studies that have proposed the two species to be self-fertile. The duration of ascospore dispersal of S. beticola and S. vesicarium from spinach seed crop stem residues in western Washington, the primary region of spinach seed production in the USA, occurred from mid-winter to late spring or early fall, potentially serving as inoculum for the next season's spinach seed crops. Growers should incorporate residues into the soil after harvest to reduce inoculum production of these pathogens on spinach seed crop residues.

8.
Plant Dis ; 2024 Aug 07.
Artigo em Inglês | MEDLINE | ID: mdl-39110614

RESUMO

European chestnut (Castanea sativa Mill.) currently reaches 1,470 ha, distributed from the Maule region to the Los Rios region in Chile. Almost 3000 tons of fruit have been exported in the last three years. A survey was carried out in January 2023 in an eight-year-old orchard located in Vilcún (38°34'46.22"S 72° 9'58.61"O), Araucanía Region. Chestnut trees with branch die back and reduced growth and vigor were detected. The incidence in the orchard was 3% (6 out of 200 trees) estimated by visual observation. Cross and longitudinal sections of the woody trunk of two trees were collected and examined, and an internal dark-brown discoloration to partial necrosis lesion was observed. To identify the causal agent, small pieces of wood from the edge of the symptomatic area were surface sterilized with 70% ethanol, rinsed twice with sterile distilled water, blotted on dry sterile filter paper, plated on potato dextrose agar (PDA) and incubated at 22°C. Fungal colonies were consistently isolated, and after 5 days, pure cultures were obtained by transferring mycelium to new PDA plates, preliminarily identified as Gnomoniopsis sp. (Visentin et al. 2012, Shuttleworth 2012). All cultures exhibited characteristics consistent with the description of G. castaneae (Syn. G. smithogilvyi), such as concentric development of greyish-brown mycelium, abundant stroma, hyaline conidia of 7.2 ±0.54 (6.1-8.1) X 2.3 ±0.26 (1.5-2.9) µm (n= 30), mainly biguttulate and fusoid. Total DNA was extracted, rDNA amplified using ITS1/ITS4 primers (White et al. 1990), and the fragment was Sanger sequenced and the sequence was deposited in GenBank (OR665735). BLAST analysis revealed a 99% identity to G. castaneae (MH384925). In addition, the DNA of the isolate was evaluated in a species-specific multiplex PCR (Silva-Campos et al. 2022), and the amplicons were electrophoretically separated, giving a similar band profile to G. smithogilvyi RGM 2903 and RGM 2904 strain from Chilean Collection of Microbial Genetic Resources. Pathogenicity of G. castaneae isolate (CV-11) was tested on ten replicates of 3-year-old C. sativa plants. Two wounds were made on the same season growing shoot and two on the previous season shoot. Longitudinal wounds (5 mm long, 4 mm wide and 2 mm depth) were made using a scalpel without removing the outer bark to inoculate the plants. Each wound was inoculated with a 5-mm mycelium plug, covered with the outer bark, and wrapped with Parafilm. Plugs of PDA were placed onto the wounds of two plants as control. The plants were kept in a growth chamber (22 ±1 0C and 90± 5% RH). All plants showed dark brown cankers measuring 20 to 40 mm long two weeks after inoculation. Also, most plants inoculated in the same season shoot presented wilted and chlorotic foliage. Mature conidiomata with cirri developed in most of the cankers. No symptoms were observed in the control. Fungal colonies of G. castaneae were reisolated on PDA from all inoculated chestnut plants and were not recovered from the controls. Recently, G. smithogilvyi has been identified as the causal agent of brown rot on chestnut nuts in Chile (Cisterna Oyarce et al. 2022); however, in several countries, it has also been associated as the causal agent of cankers in branch and stem of chestnut, as well as an endophyte in different hardwood species. Future studies on the incidence of this pathogen and its impact on chestnut yield should be carried out in the producing regions because it represents an emerging threat to Chilean chestnut production.

9.
Plant Dis ; 2024 Aug 07.
Artigo em Inglês | MEDLINE | ID: mdl-39110611

RESUMO

Diverse Phytophthora species, including many important plant pathogens, have been widely detected among surface water irrigation sources. In the past decade, metabarcoding has been used to characterize waterborne Phytophthora populations. Metabarcoding typically involves amplification of portions of the nuclear ribosomal internal transcribed spacer (ITS)1 or ITS2 from Phytophthora species, followed by indexed high throughput sequencing. However, full-length sequences of the entire ITS region are required for resolution of many Phytophthora species. We used metabarcoding with PacBio sequencing of full-length ITS amplicons to analyze populations of Phytophthora in waterways of the Stockton East Water District (SEWD) in the northern San Joaquin Valley of California. This approach yielded species-level resolution of many members of the Phytophthora community. Results were compared to those obtained by using ITS1 or ITS2 regions alone and were found to provide superior species resolution for P. pini, P. capsici, and P. gregata. Samples were collected throughout the 2021 irrigation season from five waterways across the SEWD. Thirty-eight Phytophthora species were detected in the waterways, including tree-crop pathogens P. acerina, P. cactorum, P. pini, P. ×cambivora, P. niederhauserii, P. mediterranea, and P. taxon walnut. These pathogenic species were detected throughout the SEWD during most of the irrigation season. The results demonstrated the utility of full-length ITS amplicon sequencing for identifying Phytophthora species in environmental samples and suggested that some disease risk may be incurred by orchardists irrigating with SEWD water. Additional epidemiological studies will be required to critically evaluate this risk.

10.
Plant Dis ; 2024 Aug 07.
Artigo em Inglês | MEDLINE | ID: mdl-39110616

RESUMO

Dandelion (Taraxacum mongolicum), belonging to the Asteraceae family, is one of the main associated species in the alpine meadow, and is famous for its both feeding and medicinal values (Lin et al. 2022; Wang et al. 2022). In September 2021 (vigorous growth period), a newly emerging leaf spot disease of T. mongolicum were observed on natural grassland in Ruoergai County, Aba (Ngawa) Tibetan and Qiang Autonomous Prefect, China (33°59'51'' N, 102°44'57'' E, alt. 3414 m). Leaf disease incidence was ranged from 10% to 15%. The symptoms appeared as brown to dark brown, circular or irregular, sunken spots; eventually, the infected sites of leaves formed a hole in the middle position of lesions. For isolation, 21 tissue pieces (5mm × 5mm) from 7 symptomatic leaf samples of 4 different plants were surface sterilized with 70% ethanol for 30 s and rinsed three times with sterilized distilled water. Then, these tissues were placed on potato dextrose agar (PDA) at 25°C and incubated in the dark for 2 to 7 days. Finally, six pure Didymella strains with consistent colony characteristics were obtained from hyphal tips as described by Xue et al. (2023). Colonies on PDA were brown to black with concentric circles, and abundant black pycnidia were visible; reverse similar in color. Conidia were ellipsoidal, ovoid, hyaline, 0 to 1-septate, 4.07 to 8.67 × 2.74 to 5.35 µm (average 3.60 × 6.55 µm; n = 50). Seven-week-old healthy plants were obtained by growing T. mongolicum seeds in pots (two plants per pot). The six pure Didymella strains were subsequently used to inoculate healthy plants as follows: for each strain, eight pots were spray inoculated with a mycelial suspension of about 4 × 104 CFU/ml, referring to our previous method (Xue et al. 2023). In addition, eight pots considered as non-inoculated controls were sprayed with sterilized distilled water. All pots were individually covered with transparent polyethylene bags for 5 days to maintain high relative humidity and placed in a greenhouse at 23 to 29°C. After incubation for 10 days, the typical symptoms consisted of brown to dark brown, sunken spots, which were similar to those previously observed in nature grassland; however, symptoms were not observed on the non-inoculated plants (controls). The same fungus was reisolated from the lesions and confirmed by the morphological and molecular methods described in this note, thus fulfilling Koch's postulates. To further identify this fungal pathogen, ITS-rDNA, and two other protein-coding genes (rpb2 and tub2) of the representative strain REG28 were amplified with primers described by Chen et al. (2022). Sequences were deposited in GenBank (PP385777 for ITS, PP781948 for rpb2, and PP781947 for tub2). A maximum likelihood (RAxML) phylogenetic tree based on the combined ITS, rpb2, and tub2 alignments showed REG28, and ex-type CGMCC 3.20069 of D. uniseptata (Chen et al. 2022) formed a subclade with 100% bootstrap support (Fig. S1). The causal agent of this disease was confirmed as D. uniseptata by the morphological, molecular, and pathogenic features described above. Recently, D. macrophylla has been reported as the first record on T. officinale in Russia (Gomzhina et al. 2020). To our knowledge, this is the first report of D. uniseptata causing leaf spots on T. mongolicum worldwide. This information will be useful for the diagnosis, detection, pathogen identification and future control of this disease on T. mongolicum in natural grassland.

11.
Plant Dis ; 2024 Aug 14.
Artigo em Inglês | MEDLINE | ID: mdl-39143817

RESUMO

The Ralstonia solanacearum species complex (RSSC) causes vascular wilt of many crops and is considered one of the most destructive plant pathogenic bacteria worldwide. The species complex was recently resolved into a stable taxonomy of three species aligning with the previously determined phylotypes, namely R. solanacearum (phylotype II), R. pseudosolanacearum (phylotype I and III), and R. syzygii (phylotype IV). Knowing which Ralstonia species and subspecies are established in Australia is important to Australia's biosecurity and market access. The goal of this study was to analyse Australia's Ralstonia culture collections and to assign the isolates to the modern taxonomic groups. The results shed light on the identity, distribution, and pathogenicity of the Ralstonia strains in Australia. Ralstonia solanacearum, R. pseudosolanacearum phylotype I, and R. syzygii phylotype IV-11 are present in Australia but have limited geographic ranges. We identified two aberrant RSSC strains that have genetic similarity to R. syzygii based on sequevar analysis, but do not yield a phylotype IV multiplex PCR band, similar to the known aberrant strain ACH732. The aberrant strains may represent a novel species. Three new sequevars were determined, 72, 73 and 74. Several Ralstonia lineages remain undetected in Australia, providing evidence that they are absent. These include R. pseudosolanacearum phylotype III and the phylotype I mulberry infecting strains; R. solanacearum strains IIC and the Moko causing strains; and R. syzygii subsp. celebesensis, and R. syzygii subsp. syzygii. This study fulfilled Koch's postulates for the Australian strains, R. solanacearum wilted potato plants, and R. pseudosolanacearum wilted blueberry plants, the hosts from which they were initially isolated. The data supports the hypothesis that Australia has native and introduced strains of Ralstonia.

12.
Plant Dis ; 2024 Aug 14.
Artigo em Inglês | MEDLINE | ID: mdl-39143813

RESUMO

Take-all of wheat (Triticum aestivum L.), caused by Gaeumannomyces tritici (syn. G. graminis var. tritici), is perhaps the most important soil-borne disease of wheat globally and can cause substantial yield losses under several cropping scenarios in Oregon. Though resistance to take-all has not been identified in hexaploid wheat, continuous cropping of wheat for several years can reduce take-all severity through the development of suppressive soils, a process called "take-all decline" (TAD). Extensive work has shown that TAD is driven primarily by members of the Pseudomonas fluorescens complex that produce 2,4-diacetlyphloroglucinol (DAPG), an antibiotic that is associated with antagonism and induced host resistance against multiple pathogens. Field experiments were conducted to determine the influence of agronomically relevant first year wheat cultivars on take-all levels and ability to accumulate DAPG-producing pseudomonads within their rhizospheres in second-year field trials and in greenhouse trials. One first year wheat cultivar consistently resulted in less take-all in second-year wheat and accumulated significantly more DAPG-producing pseudomonads than other cultivars, suggesting a potential mechanism for take-all reduction associated with that cultivar. An intermediate level of take-all suppression in other other cultivars was not clearly associated with population size of DAPG-producing pseudomonads, however. The first year cultivar effect on take-all dominated in subsequent plantings, and its impact was not specific to the first year cultivar. Our results confirm that wheat cultivars may be used to suppress take-all when deployed appropriately over cropping seasons, an approach that is cost effective, sustainable, and currently being utilized by some wheat growers in Oregon to reduce take-all.

13.
Plant Dis ; 2024 Jul 01.
Artigo em Inglês | MEDLINE | ID: mdl-38949762

RESUMO

Since its debut in 1982, The Land has embodied Walt Disney's vision, capturing the attention of millions of EPCOT guests with venues focusing on agriculture and environmental stewardship and sustainability. The Land pavilion spans over eight acres in the World Nature section of EPCOT at the Walt Disney World Resort in Lake Buena Vista, Florida. The pavilion houses three attractions, namely Soarin' Around the World, Awesome Planet, and the Living with The Land boat ride, complemented by a greenhouse walking tour entitled Behind the Seeds and two restaurants. Each attraction derives inspiration from nature and challenges mankind to be responsible stewards of planet earth. This feature article focuses on the Living with The Land boat ride attraction, which traverses greenhouses showcasing agricultural technologies and crops from around the world. The sections below describe both how various show elements are designed to engage guests and how the show is made possible by applying relevant science and technology.

14.
Plant Dis ; 2024 Jul 13.
Artigo em Inglês | MEDLINE | ID: mdl-39003505

RESUMO

Turfgrasses are susceptible to a wide variety of ectotrophic root-infecting (ERI) fungi that cause root rot (Tredway et al., 2023). Among the root rot diseases, fairway patch, caused by Phialocephala bamuru P.T.W. Wong & C. Dong sp. nov., was recently identified and characterized in Australia infecting bermudagrass (Cynodon dactylon) and kikuyu (Pennisetum clandestinum) grass (Wong et al., 2015). Symptoms begin as small, 5-10 cm diameter patches of yellowed turf that may coalesce into larger areas of diseased grass. A characteristic sign of fairway patch is roots colonized by dark brown to black, ectotrophic mycelium. In June 2020, many tan colored, irregular-shaped patches ranging from 10-30 cm in diameter developed on a hard fescue (Festuca brevipila) cultivar 'Beacon' turfgrass field in North Brunswick, New Jersey, USA. The centers of these patches later died and became sunken or filled in partially by recovering hard fescue. The patches grew into tan irregular-shaped rings with diameters up to 3 m by Aug 2023. Symptoms were indicative of a root disease. Five 'Beacon' hard fescue soil cores at the interface of the symptomatic and non-symptomatic area were sampled in Aug 2023. Root and crown samples were observed under a dissecting microscope and dark ectotrophic hyphae were observed on both. Roots with visible ectotrophic mycelium were removed, rinsed in sterile water three times, cut into 5 mm pieces, and plated onto 10% potato dextrose agar amended with streptomycin and gentamicin at 100 mg/L (PDA+). The plates were incubated at 25°C in the dark for 5 days. The most abundant colonies being characteristic long, septate hyphae that were hyaline at the edge and dark brown to black in the center and resembled the fungus described in Wong et al., 2015. These colonies were subcultured onto PDA+ medium and selected for molecular identification. Other less abundant colonies could be identified using morphology after subcultured and had no record being pathogenic to turfgrass. To confirm the isolate's identity, its internal transcribed spacer (ITS) region was amplified in PCR using the ITS1F/ITS4 primers (Bellemain et al., 2010). The amplicon was then sequenced with both ITS1 and ITS4 primers by Sanger sequencing. Sequences were assembled (GenBank #PP000819). The consensus sequence was then BLASTn analyzed with default settings, and the result showed 99.64% sequence identity with P. bamuru (GenBank #MG195534.1). Koch's postulate was conducted in an environmentally controlled growth chamber. Six healthy 'Beacon' hard fescue plugs were sampled from the field. Three of the six plugs (treatment) were each inoculated with P. bamuru by placing 20 g of P. bamuru colonized millets beneath and around the plug before filling the pots with sand. The other three plugs (control) received the same treatment except the P. bamuru colonized millets were autoclaved. The pots were incubated in the growth chamber with a 16 h light period and 25/20°C day/night temperatures. Symptoms resembling those observed in the field appeared on the treatment pots after 21 days of incubation while the control pots remained healthy. The roots from the treatment pots were examined under the dissecting microscope to confirm the colonization of P. bamuru on the roots, and P. bamuru was reisolated and confirmed using the aforementioned morphological traits and molecular assays (GenBank #PP000820). This is the first report of a turfgrass root rot disease caused by P. bamuru in the United States. Further epidemiological, disease ecological, and pathogen biological studies are required to clarify the importance of this disease in the United States and establish proper disease containment or control measures.

15.
Plant Dis ; 2024 Jul 06.
Artigo em Inglês | MEDLINE | ID: mdl-38971960

RESUMO

Onion (Allium cepa L.) is the most produced vegetable after tomato worldwide and is grown on about 15,000 ha in Germany. In Lampertheim, Hesse in southwest Germany (49°40'02.3"N, 8°26'00.0"E) bulbs of the cultivar 'Red Baron F1' were harvested in September 2023 in an apparently healthy state. Four months later some of the onions showed rotting symptoms, which could not be assigned to a known storage disease. At first, the bulbs became glassy, later they showed soft rot. They originated from a field located in a growing region severely affected by "Syndrome Basses Richesses" (SBR). 'Candidatus Arsenophonus phytopathogenicus' as well as 'Candidatus Phytoplasma solani' are associated with this disease in sugar beet (Gatineau et al. 2002). Moreover, 'Ca. A. phytopathogenicus' was recently reported in association of bacterial wilt and yellowing in potato (Behrmann et al. 2023). Both phloem-restricted bacteria are vectored by the polyphagous planthopper Pentastiridius leporinus (Therhaag et al. 2024), which is highly abundant in this region. To examine, if the unknown symptoms in onion might be related to the presence of these pathogens, DNA of 69 bulbs showing a different degree of softening were analyzed. The samples were tested for the presence of 'Ca. Phytoplasma solani' in a TaqMan assay (Behrmann et al. 2022). All showed negative results. To demonstrate the presence of 'Ca. A. phytopathogenicus', universal and genus-specific primers for the amplification of 16S rDNA and a real-time qPCR assay amplifying an hsp20 fragment were employed (Christensen et al. 2004, Zübert and Kube 2021). Two bulbs of the five positive samples were in an apparently healthy state, the other three showed light to moderate softening symptoms. The 16S rDNA fragments from two samples were sequenced on both strands and aligned. Both fragments were homologous. One fragment of 1474 bp fragment showing 100% homology to the 16S rDNA from SBR (accession no. AY057392) was submitted to GenBank (accession no. PP400342). Other taxa of 'Ca. Arsenophonus' showed 16S rDNA homologies of less than 99.3 %. To corroborate the finding onion samples were subjected to PCR reactions employing genus-specific primers for the conserved tufB, secY and manA gene, which had been derived from multiple alignments of 'Ca. A. spp' sequence submissions (Sela et al. 1989, Lee et al. 2010). The tufB, secY and manA primers amplified fragments of about 980 bp, 640 bp and 930 bp, respectively, from all previously positive samples. Samples which had been tested negative for 'Ca. P. phytopathogenicus' remained negative. Fragments from two accessions were sequenced and the sequences from both isolates were 100 % identical. A BLAST search of the partial tufB gene (acc. no. PP950434) showed 98.57 % sequence identity to a yet unnamed Arsenophonus endosymbiont (acc. no. OZ026540) and 91.85 to 91.83 % to 'Ca. A. nasoniae' and 'Ca. A. apicola', respectively. A similar result was obtained for the partial secY sequence (acc. no. PP950433). The manA sequence (acc. no. PP942231) was identical to a partial sequence of 'Ca. A. phytopathogenicus' strain HN (acc. no. OK335757) and 97.42 % to 'Ca. A. nasoniae and about 87 % to related Arsenophonus species. The finding of 'Ca. A. phytopathogenicus' in onion is novel and might indicate an expanding host range of vector and pathogen in the regional crop rotation. As a correlation between the pathogen and the soft rot symptom is unclear at present, further investigations are needed.

16.
Plant Dis ; 2024 Jul 06.
Artigo em Inglês | MEDLINE | ID: mdl-38971962

RESUMO

Xanthium strumarium, known as cocklebur, is an annual herb and has been used in traditional Chinese medicine. In October 2020, powdery mildew-like disease signs and symptoms were observed on X. strumarium grown in a crop field, Xinxiang city, Henan Province, China (35.36076° N, 113.93467° E). The specimen (PX-XS2023) was stored in Xinxiang Key Laboratory of Plant Stress Biology. White colonies in irregular or coalesced circular shaped-lesions were abundant on both ad- and abaxial surfaces of leaves and covered up to 99 % of the leaf area. Some of the infected leaves were senesced. More than 70 % of plants (n = 130) exhibited these signs and symptoms. Conidiophores were straight or slightly curved, 55 to 160 × 11 to 13 µm composed of foot-cells, shorter cells and conidia. Conidia were ellipsoid to oval, 29 to 40 × 14 to 20 µm (n = 50), with a length/width ration of 2.0 to 2.5, containing fibrosin bodies. Dark brown to black chasmothecia were found on infected leaves. The appendages were mycelium-shaped and at the base of scattered or gregarious chasmothecia (n = 50, 70 to 120 µm in diameter). Asci were 55 to 80 × 50 to 65 µm (n=30). These morphological characteristics were consistent with those of Podosphaera xanthii (Braun and Cook 2012). The internal transcribed spacer (ITS) region and Glyceraldehyde-3-phosphate dehydrogenase (GAPDH) region of the fungus (PX-XS2023) were amplified and sequenced with primers ITS1/ITS4 (White et al. 1990) and GAPDH1/GAPDH3R (Bradshaw et al. 2022) according to a previously reported method (Zhu et al. 2022). The resulting sequences were respectively deposited into GenBank (Accession No. MW300956 and PP236083). BLASTn analysis indicated that the sequences were respectively 99.82 % (564/565) and 100% (272/272) identical to P. xanthii (MT260063 and ON075658). The phylogenetic analysis indicated that the strain PX-XS2023 and P. xanthii were clustered into a same branch. Therefore, the causal agent of powdery mildew on X. strumarium was P. xanthii. To conduct pathogenicity assays, mature leaves of five healthy X. strumarium (height in 50 centimeters) were inoculated with fungal conidia by gently pressing surfaces of infested leaves onto leaves of healthy plants (Zhu et al. 2020). Five untreated plants served as controls. The controls and inoculated plants were separately maintained in greenhouses (humidity, 60%; light/dark, 16 h/8 h; temperature, 18°C). Eight days post-inoculation, signs of powdery mildew were detectable on inoculated plants, however, the controls were asymptomatic. Thus, the fungal pathogen was morphologically and molecularly identified and confirmed as P. xanthii. This powdery mildew caused by P. xanthii was previously reported on X. strumarium in Korea, Russia and India (Farr and Rossman, 2021). In addition, P. xanthii was recorded on X. strumarium in Xinjiang Province, China (Tai 1979). However, this is the first report of P. xanthii on X. strumarium in central China, where is around 3000 km away from Xinjiang Province with geographically differences. The sudden presence of powdery mildew caused by P. xanthii may adversely affect plant health and thus reduce medical value of X. strumarium. Therefore, the identification and confirmation of P. xanthii infecting X. strumarium enhance the knowledge on the hosts of this pathogen in China and will provide fundamental information for disease control in the future.

17.
Plant Dis ; 2024 Jul 10.
Artigo em Inglês | MEDLINE | ID: mdl-38985507

RESUMO

Blue honeysuckle (Lonicera caerulea L.) has contributed to maintaining the forest's ecological balance and remarkable frost-resistant abilities, helping it withstand extremely cold conditions (-46 °C) and a wide pH range (5 to 8) (Sharma and Lee 2021). Between September 2022 and September 2023, leaf spots were observed on approximately 30% of blue honeysuckle plants of the 'Lanjingling' cultivar grown in a 1.13 ha field in Da Hinggan Ling Prefecture (50.32° N, 124.13° E) in Heilongjiang Province, China. The leaves of the affected plants displayed black-colored spots. To identify the causal agents, 10 healthy and symptomatic leaves were randomly collected from ten healthy and infected individual plants, respectively. Small (3 to 4 mm) segments of the symptomatic tissues were immersed in 5% sodium hypochlorite (NaOCl) for 3 min, rinsed three times with sterile distilled water, dried in a paper towel, and plated on 9-cm Petri dishes containing potato dextrose agar (PDA). Ten fungal colonies developed on the PDA plates with an isolation frequency of 100% from 10 symptomatic leaves, and all colonies displayed a morphology consistent with Cladosporium spp. (Bensch et al. 2018). Cladosporium-like fungi were not isolated from healthy leaves. Dark olive-colored mycelia were observed, with straight unbranched conidiophores bearing terminal light brown-colored limoniform conidia (1.80 to 4.50 × 2.10 to 12.60 µm) and surrounded by a thin line of white mycelium (Delisle-Houde et al. 2024). To confirm this identification, PCR amplification of two representative strains LD-299 and LD-300 genomic DNA was performed with ITS1/ITS4 (White et al. 1990) and ACT512F/ACT783R (Carbone and Kohn 1999) primers. Basic local alignment search tool (BLAST) analyses of the National Center for Biotechnology Information database showed that sequences of the ITS (PP600316, PP600317) and ACT (PP624334, PP624335) all revealed 100% (493/493 nt, 493/493 nt; 181/181 nt, 181/181 nt) shared identity with Cladosporium pseudocladosporioides strain ex-type MF473195 and HM148674 (Bensch et al. 2010), respectively. Using a neighbor-joining phylogenetic analysis based on the ITS and ACT sequences, isolates LD-299 and LD-300 clustered in the same clade of C. pseudocladosporioides. Therefore, based on its morphological characteristics and molecular phylogeny, the two isolates were identified as C. pseudocladosporioides (Cosseboom and Hu 2023). A pathogenicity test was performed using nine healthy two-year-old blue honeysuckle Lanjingling plants. Three plants were inoculated with either the LD-299 or the LD-300 conidial suspension (1 × 106 spores/ml) or with clean water as an experimental control (Aydogdu et al. 2023). All plants were cultured in a greenhouse (28℃, 75% relative humidity, 12 h light and dark cycle), and each experiment was replicated three times. Typical leaf spot symptoms were first observed on the inoculated leaves after 10 days. Morphological and molecular characterization of re-isolated pathogens from the artificially infected leaves indicated that the two isolates were identical, thereby confirming Koch's postulates. Cladosporium pseudocladosporioides previously caused leaf spot disease on artichoke (Cynara scolymus) in Türkiye (Aydogdu et al. 2023). To the best of our knowledge, this is the first report of C. pseudocladosporioides causing leaf spots on blue honeysuckle in China. Blue honeysuckle production losses due to the leaf spots are critical for growers. Therefore, further focus should be given to investigate the host range and geographic distribution of C. pseudocladosporioides.

18.
Plant Dis ; 2024 Jul 10.
Artigo em Inglês | MEDLINE | ID: mdl-38985511

RESUMO

Fluazinam, a fungicide widely used in agriculture and turf management, was traditionally thought to pose a low risk of resistance. However, our in vitro sensitivity test conducted in 2021 revealed reduced sensitivity of fluazinam in dollar spot, highlighting the need for more careful field monitoring. In 2022 and 2023, we evaluated the field responses of four Clarireedia jacksonii isolates with different in vitro sensitivity to fluazinam. Fluazinam was used at a full labeled rate (0.5 oz/1,000 ft2) and a half-rate (0.25 oz/1,000 ft2) to evaluate the effectiveness of isolate-inoculated plots in the field. In 2022, natural and sensitive isolates showed significantly better control than insensitive isolates in both half- and full-rate treatments. However, in 2023, half-rate fluazinam demonstrated limited control in high disease pressure, providing relative disease control of dollar spot less than 45% across all treatments. In contrast, full-rate fluazinam maintained significantly better control of natural and sensitive isolates compared to insensitive isolates. Our results showing in vitro insensitivity leading to field insensitivity under inoculated field conditions suggest the development of fluazinam insensitivity in the C. jacksonii population. This highlights the need for judicious use of the fungicide fluazinam and the establishment of continuous resistance monitoring. Furthermore, the loss of control observed when applied at half-rates under high disease pressure highlights the importance of careful use of fungicides.

19.
Plant Dis ; 2024 Jul 17.
Artigo em Inglês | MEDLINE | ID: mdl-39021152

RESUMO

Citrus chlorotic dwarf disease (CCDD) seriously affects the citrus industry. Citrus chlorotic dwarf-associated virus (CCDaV) is speculated to be the causal agent of CCDD. However, this speculation has not been confirmed by fulfilling Koch's postulates. In this study, an infectious clone was constructed that comprises a 1.6-fold tandem CCDaV genome in the binary vector pBinPLUS and agro-inoculated into Eureka lemon (Citrus limon) seedlings through vacuum infiltration. At 60 days post inoculation, 25% of the Eureka lemon seedlings developed symptoms of crinkling and curling that are the same as those associated with the wild-type virus. Western blotting and graft transmission assays confirmed that the infectious clone systemically infected Eureka lemon seedlings. In addition, CCDaV can establish infection on three more Citrus species and one hybrid, although at different infection rates. These findings support that CCDaV is the primary causal agent of CCDD. The infectious CCDaV clone will allow further studies on the functions of viral proteins and molecular interactions of CCDaV with its hosts.

20.
Plant Dis ; 2024 Jul 19.
Artigo em Inglês | MEDLINE | ID: mdl-39030659

RESUMO

Pecan is a valuable nut crop cultivated in the southeastern US. Among the major yield-limiting factors in the region is scab, caused by the plant pathogenic fungus Venturia effusa. Managing scab in tall trees (15 to 25+ m) in pecan orchards is challenging due to the limitations of getting sufficient spray coverage throughout the canopy. We explored the effects of hedge-pruning on scab in three orchards: 14 m tall cv. Desirable trees winter hedge-pruned on alternate sides to 11 m (site 1), 18 m tall cv. Stuart trees hedge-pruned on both sides simultaneously to 11 m (site 2), and 15 m tall cv. Caddo trees winter hedge-pruned in winter vs. summer to 11 m (site 3). At site 1 and 2 hedge-pruned trees were compared to non-pruned control trees. All trees received recommended fungicide applications to control scab via air-blast sprayer. Disease incidence and/or severity was assessed at different sample heights on shoots, foliage and fruit during three seasons (2020, 2021, and 2022). At site 1 the hedge pruned trees often had significantly or numerically more severe scab on foliage and fruit compared to the control trees, although the differences were mostly small. The frequency of mature fruit with scab severity <10% was greatest on control trees in 2021 and 2022. At site 2, there were few differences between hedge-pruned and control trees (on fruit, scab severity was either significantly less on hedge-pruned trees, or not different to the control), but the frequency of mature fruit with scab severity <10% was consistently greatest on hedge-pruned trees. At site 3, scab intensity was low, and there were no significant differences in scab severity between winter- and summer-pruning treatments. At sites 1 and 2 there was generally more severe scab at greater sample heights compared to low in the canopy. At site 3 there was little effect of height on disease. The benefit of hedge-pruning likely increases with tree height in scab-susceptible cultivars. If a tree is >~15 m tall, a greater proportion of the fruit will be within reach of efficacious spray coverage from air-blast sprayers.

SELEÇÃO DE REFERÊNCIAS
DETALHE DA PESQUISA