Your browser doesn't support javascript.
loading
Show: 20 | 50 | 100
Results 1 - 16 de 16
Filter
Add more filters

Journal subject
Affiliation country
Publication year range
1.
Plant Dis ; 2023 Mar 01.
Article in English | MEDLINE | ID: mdl-36856645

ABSTRACT

Ponkan (Citrus reticulata Blanco cv. Ponkan) is a Chinese citrus species with tasty fruit. In November 2021, an unknown postharvest disease of Ponkan fruit caused nearly 15% losses of 2000 fruits in Nanchang, Jiangxi Province (28.68° N, 115.85° E). The initial fruit's surface necrosis was brown (Xu et al. 2022) (Figure 1A). Disease spots spread to the entire fruit, white or grey hyphae appeared, and the fruit rotted. Twenty diseased fruits were surface-disinfested with 2% sodium hypochlorite and 75% ethanol, then rinsed with sterile distilled water to isolate the pathogen. Diseased tissue sections (5 × 3 mm) were incubated on potato dextrose agar (PDA) for 7 days at 25°C. Twelve of 15 monoconidial isolates have similar morphology. On PDA, the isolates produced copious white aerial mycelia. After 5-7 days on straw juice medium, two types of conidia appeared (Rice straw 60 g, Agar 20 g, distilled water 1000 mL) (Figure 1E-I). Macroconidia were abundant, falcate, slender, and slightly curved with 0-8 septa, mostly 4-5 septa (average 41.70 × 3.81 m, n=100) (Figure 1J). Microconidia were globose, oval, or piriform with 0-1 septa, 2.72 to 8.57 × 2.53 to 7.47 m (average 5.49 × 4.52 m, n=50) (Figure 1L), and chlamydospores were not observed. Conidial and colony morphology identified 12 monoconidial isolates as Fusarium graminearum (Fisher et al., 1982; Yulfo-Soto et al., 2021). Genomic DNA was extracted from three isolates using a DNA Extraction Kit (Yeasen, Shanghai, China). The ITS1/4 region combined with partial gene fragments of translation elongation factor-1 alpha (TEF-1α, primer TEF1/2, O'Donnell et al. 1998), RNA polymerase second largest subunit (RPB2, primer fRPB2-5F/7cR, Liu et al. 1999) and ß-tubulin (ß-tub, primer Bt2a/2b, Li et al. 2013) from the isolates were amplified and sequenced. The three tested isolates showed identical gene sequences. Sequences amplified from one representative isolate (PG16) have been submitted to GenBank. BLAST searches revealed that ITS (OM019317), TEF-1α (OM048103), RPB2 (ON364348), and ß-tub (OM048104) had 99 to 100% identity compared with F. graminearum (MH591453.1, KX087136.1, MF662636.1, and MZ078952.1, respectively) in GenBank. The phylogenetic analysis combined ITS - TEF-1α - RPB2 (O'Donnell et al. 2015) concatenated sequences using MEGA7.0 (Mao et al. 2021) showed the isolate was clustered with the F. graminearum clade with 100% bootstrap support (Figure 2). The isolate PG16 was used for pathogenicity tests. Ponkan fruits were surface-disinfested with 75% ethanol and rinsed with sterile distilled water three times. Then, 30 punctured wound fruits (2-mm-diameter, 2-mm-depth) with a sterile needle and 30 unwounded fruits were inoculated with conidial suspension (10 µL, 3.0 × 105 conidia/mL). while the control fruits were inoculated with 10 µL sterile distilled water. All fruits were incubated at 25°C and 90% relative humidity. Two days later, all wounded fruits inoculated with conidial suspension showed disease spots, similar symptoms to the original rotten fruits (Figure 1D). Control and conidial-inoculated unwounded fruits were healthy (Figure 1B-C). The Pathogenicity test was repeated twice, and similar symptoms were observed. Morphologically and molecularly, the re-isolated fungus matched the inoculated isolate. First report of F. graminearum causing Ponkan fruit rot in China. As Ponkan is an important citrus crop with high economic value in China, identification of the causing agent, F. graminearum, for fruit rot allows the development of control measures to manage this disease.

2.
Plant Dis ; 2023 Jul 16.
Article in English | MEDLINE | ID: mdl-37455409

ABSTRACT

Wisteria (Wisteria sinensis) is a well-known ornamental plant for environmental protection in the garden, which also has a high value for medicinal use. In December 2021, leaf spots were observed on W. sinensis plants growing on the campus of Jiangxi Agricultural University in Nanchang, Jiangxi Province (28.45° N, 115.49° E), with a incidence rate of 40% plants were infested (n = 100 investigated plants). Initially leaf spots were small and pale brown (Approx. 2 mm in diameter), which gradually expanded into round or irregular dark brown spots as disease progressed, and lesions developed greyish-white necrotic tissues in the center at later stages, eventually causing the leaves to rot. To isolate the pathogen, tissues (5 × 5 mm) at the margin of lesions were cut from ten symptomatic leaves, surface disinfected with 75% ethanol for 30 s followed by 2% sodium chloride (NaClO) for 1 min, rinsed three times with sterile distilled water, and the dried tissues were cultured on potato dextrose agar (PDA) at 28 ± 1℃ in darkness for 3 days. After culture purification, five isolates were obtained and the representative single spore isolate (ZTTJ1) was used for subsequent identification tests. After 10 days of incubation on PDA medium, colonies had dense aerial mycelium with a gray center and dark gray-green mycelium outward, with orange-red conidial masses distributed in a ring on the surface. The underside of the colonies was light gray to dark gray. Conidia were cylindrical, with ends obtuse-rounded, 11.83 to 20.74 × 3.34 to 5.33 µm (av=16.11 µm × 4.26 µm, n = 50) in size. These morphological characteristics were consistent with Colletotrichum gloeosporioides (Shi et al, 2019). Six conserved regions of isolate (ZTTJ1), internal transcribed spacer (ITS), glyceraldehyde-3-phosphate dehydrogenase (GAPDH), calmodulin (CAL), ß-tublin (TUB), actin (ACT), and chitin synthase 1 (CHS1) gene regions were amplified using ITS1/ITS4 (Gardes et al, 1993), GDF/GDR (Templeton et al, 1992), CL1C/CL2C (Li et al, 2018), Bt2a/Bt2b (Prihastuti et al, 2009), ACT-512F/ACT-783R and CHS-79F/CHS-345R (Carbone et al, 1999) primers, respectively. Using BLAST, ITS, GAPDH, CAL, TUB, ACT, and CHS1 gene sequences (GenBank Accession No. OP703312, OP713773, OP713775, OP713776, OP713772, OP713774, respectively) were over 99% identical to C. gloeosporioides (GenBank Accession No. MK967281, MH594288, MT449307, MN624110, MN107239 and MN908602, respectively). A maximum likelihood (ML) phylogenetic analysis based on ITS-ACT-GAPDH-CHS1-CAL-TUB2 sequences using MEGA7.0, placed isolate (ZTTJ1) within C. gloeosporioides. To complete Koch's postulates, 10 µL spore suspension (1.0 × 106 conidia/mL) of ZTTJ1 (7-day-old culture on PDA medium) was dropped onto 10 leaves wounded with a sterilized needle and 10 non-wounded leaves, respectively. Ten wounded leaves were inoculated with sterile water as controls. All leaves were incubated at 28 ± 1 °C and 90 % relative humidity (12 h/12 h light/dark). After 7 days, all wounded leaves inoculated with C. gloeosporioides developed symptoms as previously observed, while the control and non-wounded leaves remained healthy. The fungus re-isolated from the inoculated leaves were identified as C. gloeosporioides by morphological and molecular identification; the pathogen causing disease in W. sinensis was determined to be C. gloeosporioides. To our knowledge, this is the first report of C. gloeosporioides causing anthracnose on W. sinensis in China. This work has identified the pathogenic species of the disease, which helps to take targeted measures to control its spread, providing a basis for the prevention and treatment of the disease.

3.
Plant Dis ; 2023 May 15.
Article in English | MEDLINE | ID: mdl-37189040

ABSTRACT

Sweet persimmon is native to Japan and valued for its fruit, which are high in sugar and vitamins. In October 2021, symptoms were observed on persimmon (Diospyros kaki L. cv. Yangfeng) fruits in cold storage room in Suiping county, Henan Province (32.59 °N, 15 113.37 °E). Initially, small circular dark-brown spots were visible on the fruit rind, turning into irregular sunken dark areas, and eventually rotting 15% of 200 fruits after four weeks of cold storage (10°C, 95% relative humidity). To isolate the causal agent, 10 fruits of symptomatic tissues (4 mm2) were surface-sterilized in 2% sodium hypochlorite (NaOCl) for 1 minute, washed three times in sterile distilled water, then aseptically transferred to potato dextrose agar (PDA) and incubated for 7 days at 25°C. Fungal colonies were isolated from plant tissue, and on three colonies of similar morphology, single-spore isolation was performed. On PDA, the isolates produced circular colonies of fluffy aerial mycelia, gray-brown in the center with gray-white margins. Conidia were dark brown, obclavate or pyriform, with 0 to 3 longitudinal septa and 1 to 5 transverse septa, and a size range of 19.2 - 35.1 × 7.9 - 14.6 µm (n=100). Conidiophores were olivaceous, septate, straight, or bent, with a length of 18 - 60 × 1 - 3 µm (n=100). These morphological characteristics identify the isolates as Alternaria alternata (Simmons. 2007). Genomic DNA was extracted from a representative isolate YX and re-isolated strain Re-YX by cetyltrimethylammonium bromide (CTAB). The primers of ITS1/4, Alt-F/R, GPD-F/R, EF1/2, EPG-F/R (Chen et al. 2022), RPB2-5F/7cR (Liu et al. 1999), and H3-1a/1b (Lousie et al. 1995) were used to amplify the partial internal transcribed spacer (ITS) region, Alternaria major allergen (Alt a1), Glyceraldehyde-3-phosphate dehydrogenase (GAPDH), translation elongation factor 1-alpha (TEF), endo-polygalacturonase (endoPG), RNA polymerase second largest subunit (RPB2) and Histone 3 (His3), respectively. GenBank accession No of ITS, Alt a1, GAPDH, TEF, endoPG, RPB2, His3 were ON182066, ON160008 to ON160013 for YX and OP559163, OP575313 to OP575318 for Re-YX respectively. Sequence data of Alternaria spp. were downloaded from GenBank and the BLAST analysis showed 99%-100% homology between various A. alternata strains (ITS: MT498268; Alt a1: MF381763; GAPDH: KY814638; TEF: MW981281; endoPG: KJ146866; RPB2: MN649031; His3: MH824346). A phylogenetic analysis based on ITS, Alt a1, GAPDH, TEF, and RPB2 sequences using MEGA7 (Molecular Evolutionary Genetics Analysis) revealed that the isolate YX and Re-YX were clustered in A. alternata clade (Demers M. 2022). For the pathogenicity test, seven-day-old cultures were used to create spores suspensions (5.0 × 105 spores/mL) of each of the three isolates. Ten µL aliquots from each isolate were inoculated onto ten needle-wounded persimmon fruits; ten additional fruits were inoculated with water only to serve as controls. The pathogenicity test was three replications. Fruits were deposited in a climate box at 25°C, 95% relative humidity. Seven days post-inoculation, the wounded fruit treated with spore suspensions displayed black spot symptoms similar to the symptoms on the original fruit. There were no symptoms on the control fruits. The strain Re-YX was re-isolated from the symptomatic tissue of inoculated fruits and its identity confirmed using the morphological and molecular methods previously mentioned, fulfilling Koch's postulates. The persimmon fruit rot caused by A. alternata had been reported in Turkey and Spain (Kurt et al., 2010, Palou et al., 2012). According to our knowledge, this is the first report of black spot disease on persimmon fruits caused by A. alternata in China. The disease could infect persimmon fruits during cold storage, so more control methods should be developed to prevent postharvest disease of persimmon in the future.

4.
Plant Dis ; 2023 Jan 23.
Article in English | MEDLINE | ID: mdl-36691267

ABSTRACT

Chamaedorea elegans, native to Mexico and Guatemala, is a commonly planted indoor and small-scale garden ornamental due to its stately appearance, tolerance of low light levels, and its ability to improve air quality (El-Khateeb et al. 2010). In December 2021, an unknow leaf-spot disease was observed on C. elegans in Ganzhou City of Jiangxi Province, China (25.83 °N, 114.93 °E). The symptoms were small brown spots on the leaves, gradually expanded into irregular dark brown spots with necrotic tissue forming in the center of the lesions (Figure 2 A-1 and A-2). To isolate the pathogen, the diseased leaves were surface sterilized in 75% ethanol for 30 s. Small pieces of tissue (5 × 5 mm) were taken from the margin between diseased and healthy tissue, disinfected 1% NaClO for 45 s, washed three times in sterile water, and then placed on PDA at 25 ± 1°C for 5 days. Later, five isolates were purified from single spores and each of the five isolates has the same properties as described below. The isolates had abundant pale purple flocculent hyphae with purple pigmentation (Figure 2 C-1 and C-2). Macroconidia were falciform, straight or slightly curved, 1-2 septate, 11.75 to 22.99 × 3.06 to 4.44 µm (µ=16.08 µm × 3.37 µm, n=50) (Figure 2 D-1). Microconidia were oval or elliptical, a septate, 4.03 to 9.19 × 1.92 to 3.73 µm (µ=5.88 µm × 2.66 µm, n=50) (Figure 2 D-2). Chlamydospores formed singly or in pairs, and were terminal or intercalary in hyphae (Figure 2 D-3). Based on morphological characteristics, the fungus was preliminarily identified as a Fusarium sp. (Leslie et al. 2006). To confirm the identification, primers ITS1/ITS4 (White et al. 1990), RPB2-5f2/RPB2-7cr (O'Donnell et al. 2010; Liu et al. 1999) and TEF 1-αF/TEF 1-αR (O'Donnell et al. 2000) were used to amplify and sequence apportion of the ITS, RPB2 and TEF (Table 1). The sequences (Genebank accession number: OM780148, OM782679, OM782680) shared 100% idnetity with Fusarium oxysporum (Genebank accession number: MH866024.1, MH484930.1, MH485021.1). The maximum likelihood (ML) phylogenetic analysis of the concantenated ITS, RPB2 and TEF sequences was performed in MEGA7.0. (Sudhir et al. 2016), assigning the isoaltes to the F. oxysporum species complex (Figure 1). To confirm the pathogenicity, nine pots of healthy 3-year-old C. elegans plants were inoculated in the greenhouse (12 h light/12 h dark cycle, RH 90 %, three for wounded inoculation, three for nonwounded inoculation and three for control). Fifty disinfected leaves were wounded with sterile needles and fifty remained unwounded. The wounded (Figure 2 B-1 and B-2) and unwounded leaves were inoculated with a 10 µL spore suspension (1.0 × 106 conidia/ml) which was taken from each of the five isolates cultured for 7 days. Fifty leaves were mock-inoculated with sterile water (Figure 2 B-3 and B-4). After incubation for 7 days, the wounded leaves inoculated with the spore suspension had similar symptoms to the original diseased leaves, while the unwounded leaves and the control leaves did not develop symptoms. The experiment was repeated three times and the pathogens was reisolated from wound-inoculated leaves with the same morphological characteristics to the original pathogens, and identified as F. oxysporum by morphological and molecular analysis, completing Koch's postulates. F. oxysporum, a pathogen with a broad spectrum of hosts, ranks 5th among the top 10 fungal plant pathogens (Amjad et al. 2018.) and has been reported to Carpinus betulus, Citrullus lanatus, Pinus pinea (Mao et al. 2021; Muhammad et al. 2021; Monther et al. 2021). To our knowledge, this is the first report of leaf spot disease on C. elegans caused by F. oxysporum in China. C. elegans is an important ornamental plant in China with high economic value, so the disease has the potential to be a threat to its cultivation industry.

5.
Plant Dis ; 2023 Sep 05.
Article in English | MEDLINE | ID: mdl-37669177

ABSTRACT

Paeonia lactiflora Pall is a traditional famous flower with long cultivated history in China, and has important medical and ornamental functions (Duan et al. 2022). In the middle of June 2022, anthracnose disease was observed nearly 25% (n=90) on P. lactiflora in Poyang County, Shangrao City, Jiangxi Province (29.00° N, 116.67° E) (Figure 1 E). The symptoms of the disease were small, round, light brown spots then grew bigger to round or irregular dark brown lesions (5 to 7 mm diameter) progressively on the leaves with disease spread (Figure 1 A). Subsequently, necrotic tissue was formed in the center and caused fade and wilt on the leaves ultimately, which reduced the medicinal and aesthetic value severely. Small pieces of diseased tissue (5 × 5 mm) were cut from the diseased junction, disinfected with 75% ethanol for 30 to 45 seconds, then 1% NaClO for 1 to 2 minutes, rinsed three times with sterile water. To identify the pathogen, tissues were placed on PDA and incubated for 3 days at 28°C. Single spore isolates were cultured on PDA, the colonies of one representative strain (SY4) were originally white with a lot of aerial mycelium after 5 to 7 days at 28°C in the incubator. The center of the colony turned greyish-white, released tiny orange-yellow particles (conidia) (Figure 1 F and 1 G), which were single, colorless, elongated ovals with rounded ends and measured 11.29 to 23.24 × 3.94 to 5.60 µm (av=15.89 µm × 4.74 µm, n=50) (Figure 1 H and 1 I). The isolate SY4 was identified to Colletotrichum fructicola based on morphological characteristics (Yang et al. 2021; Li et al. 2022b). For further molecular identification, the rDNA-ITS, actin gene (ACT), glyceraldehyde-3-phosphatedehydrogenase (GAPDH), chitin synthase (CHS) and calmodulin gene (CAL) genes were amplified and sequenced with primers of ITS1/ITS4 (Gardes et al. 1993), ACT-512F/ACT-783R, GDF/GDR (Templeton et al. 1992), CHS-79F/CHS-345R (Carbone et al. 1999) and CL1C/ CL2C (Weir et al. 2012) respectively. The accession numbers in GenBank were OP523977 (ITS-rDNA), OP547618 (ACT), OP605733 (GAPDH), OP605732 (CHS), and OP605731 (CAL). The BLAST analysis revealed that these sequences were identical more than 99% with those of C. fructicola (GenBank accession Nos. MZ437948.1, MN525803.1, MN525860.1, MZ13360.1 and ON188684.1) (Figure 2). To confirm pathogenicity, the leaves were cleaned with 75% ethanol, rinsed with sterile water. After the leaf surface was dried naturally, 20 leaves were pricked at two symmetrical places on either side of the main veins of the leaf with a sterilized inoculum needle (2.0 mm in diameter), half of the wounded leaves were inoculated with 20 µL spore suspension (1.0 × 106 spores/mL) (Figure 1 C and 1 D), while the other half were inoculated with sterile water as controls (Figure 1 B). Inoculated leaves were grown for 5 days in an incubator at 28 °C and above 90% relative humidity, repeated three times. The results demonstrated that the wounded leaves with C. fructicola showed the same signs of wilting with the original disease leaves, while control leaves remained healthy. The same fungus was reisolated from the diseased leaves which confirmed with Koch's postulates. The same fungus was re-isolated from the diseased leaves while it was not isolated from control leaves, confirmed with Koch's postulates. In China, it had been reported that C. fructicola caused anthracnose on Persea americana (Li et al. 2022a) and Myrica rubra (Li et al. 2022b). To the best of our knowledge, this is the first report of anthracnose on P. lactiflora caused by C. fructicola in China. The results will help to develop effective control strategies for anthracnose on P. lactiflora.

6.
Plant Dis ; 2022 Mar 09.
Article in English | MEDLINE | ID: mdl-35263153

ABSTRACT

Ophiopogon jaburan (Liliaceae), named white lilyturf, is widely cultivated as an ornamental plant in south China. During 2017-2019, leaf spots on O. jaburan were observed all year in Zhanjiang, Guangdong, China (N21°9'3"; E110°17'47"). Almost all plants were infected and the disease incidence on affected leaves was about 80% in the field. Initially, spots were brown, round or oval, and gradually enlarged to irregular shapes. The color of the spots changed from rusty-brown to grayish-white with rusty-brown borders. Subsequently, the spots expanded until the leaves withered and died. Infected tissues were surface-sterilized with 75% ethanol for 30s followed by 1% NaClO solution for 1 min, then rinsed thrice with sterile water, before placed on potato dextrose agar (PDA) containing 50mg/L ampicillin, and incubated in darkness at 25℃ with 90% relative humidity. Colonies growing on PDA were cushion-like, pale greenish grey to grayish black on the front side and clearly dark gray on the reverse. Colony diameter was av. 86.0 mm (n = 15) grown in the dark at 25 ℃ for 10 days. Conidia with oil droplets were colorless, hyaline, smooth-walled, aseptate, slightly curved, and tapered gradually to each end, 12.3-28.9 × 2.2-6.6 µm (av. 20.9×4.2µm, n=200). Setae were brown to dark brown, 2-4 septate, with the base slightly inflated, and measured 40.0-130.3 × 2.2-5µm (av. 84.3 × 3.3µm, n=23). On PDA, scattered or loosely clustered appressoria were elliptical or irregular, smooth-walled, aseptate, and dark brown. To confirm the identification, partial regions of the internal transcribed spacer (White et al. 1990), beta-tubulin (Aveskamp et al. 2009) and actin (Carbone et al 1999) were amplified and sequenced (MW989743, MZ014461 and MZ014462). The blast results showed these sequences had >99.59% homology with sequences of Colletotrichum liriopes holotype strain CBS 119444 (NR_111449, GU228098 and GU227902). Maximum likelihood analysis and Bayesian inference were performed from concatenated sequences using RAxML v.1.0.0 and MrBayes v.3.2.1 software respectively. Several C. liriopes strains clustered in the same clade. Based on morphological-molecular characteristics, the fungus was identified as C. liriopes (Damm et al 2009; Chen et al. 2019). To confirm pathogenicity, healthy leaves were surface disinfected with 75% ethanol and rinsed thrice with sterile water. On ten leaves, three sites were wounded by pricking with needles, and inoculated 20 µL of 106 conidia/ml suspension or mycelium in contact with blade surface using 6-mm mycelial plugs. Similarly, the inoculation was done for three unwounded sites each leaf. Sterile water and medium plugs (without fungus) served as controls. All leaves were incubated on sterile wet filter paper at 25-28℃ with 90% relative humidity. After 7 days, all the inoculated leaves showed symptoms similar to those of field diseases, whereas control leaves remained healthy. The fungus with morphological-molecular features identical to the original isolate was reisolated from the disease lesions. C. liriopes causes anthracnose on Bletilla ochracea, Eria coronaria, Hemerocallis fulva, Pleione bulbocodioides (Jayawardena et al 2016) and Liriope sp. (Yang et al 2020; Chen et al 2019) in China. This is the first report of C. liriopes causing anthracnose on O. jaburan in China. Anthracnose could greatly affect ornamental value of O. jaburan, and this work can alert gardeners to prevent and control of the disease.

7.
Plant Dis ; 2022 Mar 06.
Article in English | MEDLINE | ID: mdl-35253487

ABSTRACT

Solanum muricatum is native to South America and well known for its sweet, attractive, nutritious fruits. S. muricatum has been cultivated in China since the 1980s and increasingly popular (Li et al. 2015). In November 2021, an unknown fruit rot was observed in Shilin County of Yunnan Province (24.77 °N, 103.28 °E). The incidence of this disease was about 16% of 500 postharvest S. muricatum fruits after 7 d in storage room (25°C, 90% relative humidity). The initial symptoms were small brown spots on the fruit surface, which gradually expanded into irregular brown or black lesions, and gray-white mold developed in the center of the lesions, eventually the fruit turned rot. To isolate the pathogen, ten fruits with typical symptoms were collected and surface-sterilized with 75% ethanol for 45 s. Small fragments (5 × 5 mm) from the margin of lesions on fruit were disinfected with 1% sodium hypochlorite for 60 s, washed three times with sterile water then transferred to potato dextrose agar (PDA), and incubated at 28 ± 1℃ for 3 days (Li et al. 2022). Two fungal isolates with the same morphology were obtained and purified by single-spore isolation method. The colony was covered with thick fluffy aerial mycelia and the center was dark brown or black with white margins. Conidia were brown, pyriform or ellipsoid, with 1 to 3 longitudinal and 2 to 6 transverse septa, 15.12 to 34.01 × 6.90 to 12.73 µm (21.22 × 9.69 µm on average, n=50) in size. These morphological characteristics were consistent with Alternaria alternata (Li et al. 2015; Xiang et al. 2021; Alberto. 1992). For molecular identification, genomic DNA was extracted from a representative isolate, and primers ITS1/ITS4 (Gardes et al. 1993), TEF-F/TEF-R (Lawrence et al. 2013), Alt-F/Alt-R (Hong et al. 2005), GPD-F/GPD-R (Berbee et al. 1999) and EPG-F/EPG-R (Peever et al. 2004) were used to amplify the internal transcribed spacer (ITS), translation elongation factor 1-alpha (TEF), Alternaria major allergen (Alt a1), glyceraldehyde-3-phosphate dehydrogenase (GAPDH) and endo-polygalacturonase (endoPG), respectively. The obtained DNA sequences (ITS: OM049821; TEF: OM069656; Alt a1: OM069655; GAPDH: OM069654 and endoPG: OM069653) showed over 99% homology with that of A. alternata (GenBank Accession No. MN856355.1 (565/573 bp); MN258023.1 (267/267 bp); KY923227.1 (491/501 bp); LC131645.1 (608/609 bp) and MN698284.1 (452/454 bp)). A phylogenetic tree based on the combined ITS, TEF, Alt a1, GAPDH, and endoPG sequences using the maximum likelihood methods with Kimura 2-parameter model, bootstrap nodal support for 1000 replicates in MEGA7.0 (Li et al. 2019) revealed that the isolate was assigned to A. alternata. To confirm pathogenicity, 10 µL spore suspension (1.0 × 106 conidia/ml) obtained from 7-day-old PDA cultures of each isolate were inoculated on 15 needle-wounded and 15 non-wounded surface-disinfected fruits, respectively. Healthy fruits were inoculated with sterile water as controls and the experiment was repeated 3 times. All fruit were incubated at 25 ± 1℃, 90% relative humidity. After 7 days, all the wounded and non-wounded fruit inoculated with A. alternata showed similar symptoms to those observed on the previously fruits, while the control fruits remained healthy. The same pathogen was again isolated from the inoculated fruits, thus Koch's postulates were fulfilled. A. alternata causing fruit rot of Prunus avium and Mangifera indica in China were reported in previous studies (Ahmad et al. 2020; Liu et al. 2019). As far as we know, this is the first report of postharvest fruit rot on S. muricatum caused by A. alternata in southwest China. This work provides a basis for the development of control strategies of the disease in the future.

8.
Plant Dis ; 2022 Jul 08.
Article in English | MEDLINE | ID: mdl-35802012

ABSTRACT

Citrus sinensis (L.) Osbeck is popular with consumers for its delicious taste. In December 2020, a rot symptom causing about 15% losses of a total of 450 fruits was observed on 'Newhall' navel oranges after 70 d storage (20℃, 85%-90% RH) at Jiangxi Key Laboratory for Postharvest Technology and Nondestructive Testing of Fruits & Vegetables (28.68° N, 115.85° E). The fruits were harvested from an orchard in Ganzhou City, Jiangxi Province, China (25.53° N, 114.79° E). Incipiently, the pedicles of infected fruits were brown, the peels became softened and showed yellowish-brown lesions which, gradually expanded and had white hyphae (Fig. S1A). To isolate the pathogen, the surface of diseased fruits was disinfected with NaClO (2%) for 2 min and ethanol (75%) for 0.5 min, then washed with sterile water three times. Tissues (5 × 5 mm) around the lesion were incubated on potato dextrose agar (PDA) at 28 ± 1℃ (L: D=12: 12) for 5 days. Five cultures with similar morphology were obtained and colonies initially produced white aerial hyphae and became khaki then turned pink on PDA (Fig. S1F, G, H). Abundant microconidia, macroconidia and rare chlamydospores were observed after 10 days on PDA and no glucose PDA media (Zhang et al. 2020). Macroconidia were falciform and curved to lunate, 2-4 septa, 29.38 × 3.75 µm in size (n=50) (Fig. S1K, Fig. S3). Microconidia were oval, napiform or pyriform, 0-1 septa, 12.00 × 3.43 µm in size (n=50) (Fig. S1L1 to L4, Fig. S3). Chlamydospores were found in hyphae, ellipsoidal or orbicular (Fig. S1I-1 to I-2, J-1 to J-2). The morphological features of five isolates were similar to Fusarium (Leslie and Summerell 2006). Genomic DNA of five isolates was extracted with DNA Extraction Kit (Yeasen, Shanghai, China), ITS1/ITS4, EF1Ha/EF2Tb and fRPB2-5F/fRPB2-7cR primers were used to amplify the internal transcribed spacer region (ITS), and the transcriptional elongation factor-1 alpha (TEF-1α), and RNA polymerase II (RPB2) gene sequences (White et al. 1990; Carbone and Kohn 1999; Liu et al. 1999). The ITS, TEF-1α and RPB2 sequences of five isolates were deposited in GenBank and showed 99-100% identity with corresponding sequences from F. tricinctum (Table S1). A phylogenetic tree was constructed with ITS-TEF-1α-RPB2 concatenated sequences in MEGA7.0 (Li et al. 2021) and all five isolates were placed in F. tricinctum clade with 100% bootstrap support (Fig. S2). To confirm pathogenicity, ten healthy C. sinensis fruits were surface-sterilized with 75% ethanol and inoculated with 10 µL spore suspension (1.0 × 106 spore/mL) including five wounded (with sterilized needle) and five unwounded (Fig. S1B to E). Control fruits were inoculated with 10 µL sterile water. All fruits were incubated at 28 ± 1℃, 90% RH for 7 days. The experiment was conducted three times. The lesion diameter of inoculated wounded fruits was 21.01 ± 2.52 mm and showed similar symptoms to original rotten fruits. However, the control and unwounded fruits remained healthy. To fulfill Koch's postulates, F. tricinctum was re-isolated from the inoculated fruits and deposited in Collaborative Innovation Center of Postharvest Key Technology and Quality Safety of Fruits and Vegetables in Jiangxi Province. To our knowledge, F. tricinctum has been reported on apple tree and kiwi plant in China (Zhang et al. 2021; Ma et al. 2022), but this is the first report of F. tricinctum causing fruit rot on navel orange in China. This finding provides important information for preventing postharvest disease of citrus.

9.
Plant Dis ; 2022 Aug 19.
Article in English | MEDLINE | ID: mdl-35984392

ABSTRACT

Nanfeng tangerine (Citrus reticulata Blanco) is highly regarded for its nutritional and economic value. In January 2022, an unknown fruit rot was observed on Nanfeng tangerine fruits harvested from Nanfeng County (27.22 °N, 116.53 °E), Fuzhou City, Jiangxi Province after 70 days in storage (25 °C, 90% relative humidity). The disease mostly started from the pedicel or a wound. Symptoms initiated with dark brown lesions that rapidly expanded between the fruit center and pulp capsule causing total fruit rot. The surface of symptomatic fruit was sterilized with 75% ethanol for 30 s and 2% NaClO for 30 s. Small diseased tissue pieces (2 mm2) between diseased and healthy tissues were placed on potato dextrose agar (PDA) and put in an incubator (25 ± 1 °C) for 3 days. The representative isolate NFMJ-1 was subcultured onto PDA using single-spore purification. Colonies on PDA were light yellow to white, with abundant flocculent aerial hyphae. Microconidia were oval, obovoid to allantoid, 0 septate, occasionally 1 septate, 4.07 to 17.53 × 1.69 to 3.56 µm (average=7.40 µm × 2.55 µm, n=50). Macroconidia were slender, with a beaked apical cell and a foot-shaped basal cell, 3 to 5 septate, 22.99 to 81.12 × 2.34 to 3.81 µm (average=45.04 µm × 3.12 µm, n=50). According to morphological characteristics, the isolate was tentatively identified as Fusarium sp. (Leslie and Summerell 2006). To confirm the identification, the internal transcribed spacer (ITS), translation elongation factor 1-alpha (TEF), RNA polymerase II second largest subunit (RPB2), beta-tubulin gene (TUB2), and calmodulin gene (CaM) sequences were amplified with primers ITS1/ITS4 (Gardes et al. 1993), TEF1/TEF2 (O'Donnell et al. 2010), RPB2-5f2/RPB2-7cr (Liu et al. 1999), Bt2a/Bt2b (Glass and Donaldson 1995), and CL1C/CL2C (Weir et al. 2012), respectively. The obtained sequences (ON184033, ON212051, ON212052, ON212053, ON212054) showed homology with F. concentricum ITS (MW016417.1; 514/514 bp), TEF (MK609902.1; 667/667 bp), RPB2 (LC631461.1; 941/972 bp), TUB2 (MT942588.1; 331/337 bp), and CaM (MK609916.1; 558/597 bp). A phylogenetic analysis of concatenated ITS-RPB2-TEF sequences was performed by MEGA7.0 with the maximum likelihood and Kimura 2-parameter model, revealing that the isolate was placed in the F. concentricum clade. To confirm pathogenicity, 36 healthy tangerine fruits were surface sterilized with 75% alcohol, then 18 disinfected fruits were wounded with sterile needles and 18 remained unwounded. Half of the wounded and un-wounded fruits were inoculated with 10 µL of a conidial suspension (1.0 × 106 conidia/ml) of isolate NFMJ-1 cultured for 7 days on PDA. Half of the wounded and un-wounded fruits were mock-inoculated with sterile water as controls. After incubation in an incubator (25 ± 1°C, 90% relative humidity) for 7 days, the wounded fruits inoculated with F. concentricum showed similar symptoms to the original diseased fruits, while the mock-inoculated fruits were asymptomatic. The pathogenicity test was repeated three times. The pathogen was re-isolated from the wound-inoculated fruits and identified as F. concentricum by morphological and molecular analysis, completing Koch's postulates. F. concentricum has been reported as a pathogen of Podocarpus macrophyllus (Dong et al. 2022), Capsicum annuum (Wang et al. 2013) and Zea mays (Du et al. 2020) in China. This is the first report of fruit rot caused by F. concentricum on Citrus reticulata in China. Appropriate prevention and control measure of the pathogen need to be developed to preserve marketability of this economically important citrus fruit.

10.
Plant Dis ; 2022 Apr 05.
Article in English | MEDLINE | ID: mdl-35380465

ABSTRACT

Begonia lanternaria Irmsch., an ornamental plant endemic in China, which is commonly used in landscape and interior decoration. In March 2021, an estimated 30% B. lanternaria plants were observed with anthracnose-like symptoms at a botanical garden conservation greenhouse in Mengla County of Yunnan Province (21.91° N, 101.21°E). Initially, small black spots developed on the disease leaves, which gradually expanded into irregular necrotic lesions surrounded by a yellowish halo, eventually turned wilting and defoliating. Twenty diseased leaves were collected and surface-disinfested with 75% ethanol for 30 s. Small fragments (5 × 5 mm) from the margin of lesions were disinfected with 1% NaClO for 120 s, washed with sterile water three times, and cultured on potato dextrose agar (PDA) at 28 ± 1℃. After 3 days single spores from four fungal colonies with identical morphology were isolated. Colonies on PDA were 70-75 mm diam in 7 d (7.5-10.6 mm/d), with dense white to gray-white mycelia attached with brown to black-brown acervulus. The underside of the culture was yellow to yellowish-brown concentric circle. Conidia were single-celled, hyaline, straight to slightly curved, cylindrical, 12.88 to 16.66 × 6.25 to 7.97 µm (av=14.65 µm × 7.22 µm, n=50) in size. For molecular identification, genomic DNA was extracted from a representative isolate, and the internal transcribed spacer, glyceraldehyde-3-phosphate dehydrogenase, calmodulin gene, ß-tublin, actin, and chitin synthase 1 genes were amplified with ITS1/ITS4 (Gardes et al, 1993), GDF/GDR (Templeton et al, 1992), CL1C/CL2C (Li et al, 2018), Bt2a/Bt2b (Prihastuti et al, 2009), ACT-512F/ACT-783R and CHS-79F/CHS-345R (Carbone et al, 1999) primers, respectively. The obtained DNA sequences showed over 99% homology with Colletotrichum karsti (GenBank Accession No. ITS: NR144790; GAPDH: KX578772; CAL: KY039988; TUB2: KX578804; ACT: LC412408; CHS1: KU251855), and the results of sequences were deposited into GenBank with accession No. MZ496954 (522/522 bp), MZ504978 (238/238 bp), MZ504979 (737/737 bp), MZ504982 (472/472 bp), MZ504981 (273/273 bp), MZ504980 (282/284 bp). The phylogenetic tree combined with ITS-ACT-GAPDH-CHS 1-CAL-TUB2 concatenated sequences using the maximum likelihood methods showed that the isolate was C. karsti. To confirm pathogenicity, Koch's postulates were conducted on intact plants, 10 µl spore suspension (1.0 × 106 conidia/ml) of each of four isolates (7-day-old culture on PDA) was inoculated on 15 wounded with a sterilized needle or non-wounded healthy living leaves, and 15 wounded leaves were inoculated with sterile water as controls. All leaves were incubated at 28 ± 1°C and 90% relative humidity (12 h/12 h light/dark). After 5 days, all wounded leaves inoculated with C. karsti showed symptoms similar to those previously observed, while the control and non-wounded leaves remained healthy. Colletotrichum karsti was re-isolated from inoculated leaves. C. karsti was previously reported to cause disease on Nicotiana tabacum L. (Zhao et al, 2020), Stylosanthes guianensis (Jia et al, 2017) and Fatsia japonica (Xu et al, 2020) in China. To our knowledge, this is the first report of C. karsti causing anthracnose of B. lanternaria Irmsch. in China. This disease reduces the ornamental and economic value of B. lanternaria Irmsch., and this work will provide a basis for the prevention and treatment of the disease in the future.

11.
Plant Dis ; 2020 Nov 24.
Article in English | MEDLINE | ID: mdl-33231522

ABSTRACT

Tetradium ruticarpum, previously and commonly known as Evodia rutaecarpa, is a tree that produces a fruit which is one of the most important traditional Chinese medicine herbs in China (Zhao et al. 2015). In July 2019, an investigation of diseases of T. ruticarpum was conducted in the farmland of Ruichang County (29.68° N, 115.65° E), Jiujiang City, China. An unknown fruit rot disease was observed and the incidence rate was estimated to be 60% to 70% within a 5,000 m2 area. The early symptoms appeared as small circular to irregular dark brown or black spots on the fruit, which gradually coalesced to a light brown-to-black discoloration and caused fruit rot. To identify the causal agent of the disease, 10 diseased fruits were collected and surface disinfected with 2% sodium hypochlorite for 2 min, 70% ethanol for 30 s, rinsed in sterile water and dried on filter paper. Tissues from non-symptomatic tissue as well as from the margin between healthy and affected edge were incubated on potato dextrose agar (PDA) at 25±1°C (12 h light/dark) with 90% relative humidity for 5 days. The colonies were brown to black with abundant whitish margins. Conidiophores were brown and measured 20.40 - 43.10×1.30 - 4.20 µm (25.47 × 2.35 µm on average, n=50). Conidia produced in single or branched chains, were obclavate or ovoid, approximately 9.90 - 32.80×6.50 - 14.50 µm (28.75×12.57 µm on average, n=50) with 2 to 5 transverse septa and 0 to 3 longitudinal septa. The colonies were consistent with Alternaria alternata (Simmons 2007). For molecular identification, the f partial internal transcribed spacer (ITS) regions, Glyceraldehyde-3-phosphate dehydrogenase (gapdh) genes, translation elongation factor 1-alpha (TEF) and Alternaria major allergen (Alt a1) gene of the isolate were amplified using primers ITS1/ITS4 (White et al. 1990), GDF/GDR (Templeton et al. 1992), EF1-728F/EF1-986R (Carbone and Kohn 1999) and Alt-for/Alt-rev (Hong et al. 2005). Sequence data showed 100% homology to A. alternata (GenBank accessions No.MN625176.1 (570/570 bp), MK683866.1 (618/618 bp), MK637432.1 (281/281 bp), KT315515.1 (488/488 bp)), respectively and the sequence data were deposited into GenBank with accession numbers MN897753 (ITS), MT041998 (gapdh), MT041999 (TEF), and MT042000 (Alt a1). Based on both morphological and molecular characteristics, the pathogen was identified as A. alternata. To confirm pathogenicity, 10 µl of a spore suspension (1.0 × 106 conidia/ml) obtained from 5-day-old PDA cultures of the strain were inoculated on 20 wounded (using sterile needle) and 20 nonwounded healthy T. ruticarpum fruits previously disinfected in 75% ethanol. Control fruits including 20 wounded fruits and 20 nonwounded fruits were inoculated with sterilized water. All fruits were incubated at 25±1°C (12 h light/dark) with 90% relative humidity. Four days later, all the wounded and non-wounded fruits showed the initial symptoms of black rot which was similar to that observed in the field, while the wounded and nonwounded fruits treated with sterile water remained healthy. The same pathogen was again isolated from the inoculated fruits. The pathogenicity experiment was repeated three times with the same results. As far as we know, this is the first report of A. alternata causing fruits rot on T. ruticarpum in China, and the identification of the pathogen will provide useful information for developing effective control strategies.

12.
Front Nutr ; 9: 1081476, 2022.
Article in English | MEDLINE | ID: mdl-36532563

ABSTRACT

Ascorbic acid (AsA) is an important nutritious substance in fruits, and it also can maintain the biological activity of fruits during storage. This research investigated the effect of exogenous melatonin (MT) on AsA metabolism in postharvest kiwifruit. Our results indicated that exogenous MT delayed the decrease of fruit firmness and titratable acid (TA), inhibited the increase of soluble solids content (SSC), reduced the respiration rate and ethylene production, and maintained a higher AsA content in kiwifruit during storage. The high expression of L-galactose pathway key genes in the early storage and regeneration genes in the later storage maintained the AsA content in postharvest kiwifruit. MT treatment enhanced the expression levels of AsA biosynthesis (AcGME2, AcGalDH, and AcGalLDH) and regeneration (AcGR, AcDHAR, and AcMDHAR1) genes. Meanwhile, the expression of the degradation gene AcAO was inhibited in MT-treated kiwifruits.

13.
Front Plant Sci ; 13: 1097733, 2022.
Article in English | MEDLINE | ID: mdl-36589109

ABSTRACT

Botryosphaeria dothidea is a major postharvest causal agent of soft rot in kiwifruit. Methyl jasmonate (MeJA) is an important plant hormone that participates as a plant defense against pathogens from a signal molecule. However, the impact and regulatory mechanism of MeJA on the attenuation of kiwifruit fungal decay remains unknown. This work investigated the effects of exogenous MeJA on the enzyme activity, metabolite content and gene expression of the phenylpropanoid and jasmonate pathways in kiwifruit. The results revealed that MeJA inhibited the expansion of B. dothidea lesion diameter in kiwifruit (Actinidia chinensis cv. 'Hongyang'), enhanced the activity of enzymes (phenylalanine ammonia lyase, cinnamate 4-hydroxylase, 4-coumarate: coenzyme A ligase, cinnamyl alcohol dehydrogenase, peroxidase and polyphenol oxidase), and upregulated the expression of related genes (AcPAL, AcC4H, Ac4CL, and AcCAD). The accumulation of metabolites (total phenolics, flavonoids, chlorogenic acid, caffeic acid and lignin) with inhibitory effects on pathogens was promoted. Moreover, MeJA enhanced the expression of AcLOX, AcAOS, AcAOC, AcOPR3, AcJAR1, AcCOI1 and AcMYC2 and reduced the expression of AcJAZ. These results suggest that MeJA could display a better performance in enhancing the resistance of disease in kiwifruit by regulating the phenylpropanoid pathway and jasmonate pathway.

14.
Front Chem ; 10: 1104805, 2022.
Article in English | MEDLINE | ID: mdl-36583154

ABSTRACT

Two novel amides, named clauphenamides A and B, and twelve other known compounds were isolated from the twigs and leaves of Clausena lansium Lour. Skeels (Rutaceae). Their structures were elucidated on the basis of extensive spectroscopic analysis and comparison with data reported in the literature. Clauphenamide A (1) featured in the unit of N-2-(4,8-dimethoxyfuro [2,3-b]quinolin-7-yl)vinyl, and clauphenamide B (2) was a unprecedented N-phenethyl cinnamide dimer. Other known compounds belong to pyrrolidone amides (3 and 4), furacoumarins (7-10), simple coumarins (11-14), lignan (5) and sesquiterpene (6). Compounds 5, 6, 10 and 12 were separated from the genus (Clausena) for the first time, while 13 was isolated in the species (C. lansium) for the first time. The antifungal activities of the isolated compounds were assayed. As a result, at the concentration of 100 µg/ml, compared with the control (chlorothalonil, inhibition rate of 83.67%), compounds 1 and 2 were found to exhibit moderate antifungal activity against B. dothidea with inhibition rates of 68.39% and 52.05%, respectively. Compounds 11-14 also exhibited moderate activity against B. dothidea and F. oxysporum, with inhibition rates greater than 40%. In addition, compared with the control (chlorothalonil, inhibition rate of 69.02%), compounds 11-14 showed strong antifungal activity to P. oryzae, with inhibition rates greater than 55%. Among them, compound 14 has the strongest antifungal activity against P. oryzae, and the inhibition rate (65.44%) is close to that of the control chlorothalonil. Additionally, the structure-activity relationships of the separated compounds are also discussed preliminarily in this paper.

15.
Pest Manag Sci ; 77(3): 1355-1360, 2021 Mar.
Article in English | MEDLINE | ID: mdl-33089666

ABSTRACT

BACKGROUND: Clausena lansium (Lour.) Skeels, belonging to the genus Clausena of the family Rutaceae, has a wide range of medical and agricultural activities. Previous studies on agricultural activities have shown that C. lansium extracts and some components have obvious herbicidal activities. In order to study systematically herbicidal activity of this plant, we studied the herbicidal effect of ethyl acetate (EtOAc) extract from the stems and leaves of this plant and further isolated the active compounds. RESULTS: The EtOAc extract inhibited the growth of roots and shoots of Echinochloa crus-galli (L.) Beauv., and the inhibitory effect of the EtOAc extract on roots were stronger than those on shoots with half maximal inhibitory concentration (IC50 ) values of 420.45 and 585.05 mg L-1 , respectively. Fifteen compounds were subsequently isolated and identified from the stems and leaves of C. lansium, including nine O-monoterpenoid furanocoumarins and six cinnamamides. Our results showed that most compounds exhibited varying degrees of herbicidal activities to E. crus-galli. Among them, compounds 3, 8, and 13-15 showed the best inhibitory activities on the growth of E. crus-galli roots, with inhibition rate values ranging from 70% to 83% at a concentration of 300 mg L-1 . Compounds 1 and 2 are two new compounds, and their structures were established as 5-O-monoterpenoid furanocoumarin and 8-O-monoterpenoid furanocoumarin, and named as claulansicoumarin-A and -B, respectively. CONCLUSION: The EtOAc extract and pure compounds showed noticeable herbicidal activities against E. crus-galli and indicated a great potential for these natural compounds to be developed as a herbicide. © 2020 Society of Chemical Industry.


Subject(s)
Clausena , Furocoumarins , Herbicides , Herbicides/pharmacology , Monoterpenes , Plant Leaves
16.
Food Chem ; 305: 125483, 2020 Feb 01.
Article in English | MEDLINE | ID: mdl-31610420

ABSTRACT

Kiwifruit (Actinidia deliciosa cv. Jinkui) were treated with 0.1 mmol/L methyl jasmonate (MeJA) to investigate the effects on disease resistance to soft rot caused by Botryosphaeria dothidea. The results showed that MeJA treatment significantly reduced the diameter of lesions after inoculation with B. dothidea. This treatment significantly enhanced the activities of related antioxidant protective enzymes, defence-related enzymes including catalase (CAT), peroxidase (POD), superoxide dismutase (SOD), polyphenol oxidase (PPO), chitinase (CHI), ß-1,3 glucanase (GLU) and increased the accumulation of total phenolic content, while the degree of membrane lipid peroxidation was reduced. MeJA treatment effectively enhanced gene expression of AcPOD, AcSOD, AcCHI and AcGLU. The results from this research suggest that MeJA treatment is a promising and safe strategy for controlling postharvest rot soft of kiwifruit.


Subject(s)
Acetates/pharmacology , Actinidia/microbiology , Ascomycota/drug effects , Cyclopentanes/pharmacology , Disease Resistance/drug effects , Oxylipins/pharmacology , Actinidia/chemistry , Actinidia/metabolism , Ascomycota/physiology , Chitinases/genetics , Chitinases/metabolism , Fruit/chemistry , Fruit/metabolism , Fruit/microbiology , Gene Expression/drug effects , Lipid Peroxidation/drug effects , Peroxidase/genetics , Peroxidase/metabolism , Phenols/analysis , Superoxide Dismutase/genetics , Superoxide Dismutase/metabolism
SELECTION OF CITATIONS
SEARCH DETAIL