Your browser doesn't support javascript.
loading
Mostrar: 20 | 50 | 100
Resultados 1 - 20 de 82
Filtrar
1.
Ecotoxicology ; 2024 Jul 18.
Artigo em Inglês | MEDLINE | ID: mdl-39026048

RESUMO

The use of some organophosphate insecticides is restricted or even banned in paddy fields due to their high toxicity to aquatic organisms. The aim of this study is to elucidate the main pathways and target organs of organophosphate insecticide toxicity to fish exposed via different routes by integrating histopathological and biochemical techniques. Using malathion as the model drug, when the dosage is 20-60 mg/L, the toxicity of whole body and head immersion drugs to zebrafish is much higher than that of trunk immersion drugs. A dose of 21.06-190.44 mg/kg of malathion feed was fed to adult zebrafish. Although the dosage was already high, no obvious toxicity was observed. Therefore, we believe that the drug mainly enters the fish body through the gills. When exposed to a drug solution of 20 mg/L and 60 mg/L, the fish showed significant neurological behavioral abnormalities, and the pathological damage to key organs and brain tissue was the most severe, showing obvious vacuolization and the highest residual amount (8.72-47.78 mg/L). The activity of acetylcholinesterase was the most inhibited (54.69-74.68%). Therefore, brain tissue is the key toxic target organ of malathion in fish. In addition, we compared the bioaccumulation effects of different water-soluble organophosphorus insecticides in fish and their toxic effects. We found that the higher the water solubility of organophosphorus insecticides, the lower their toxicity to fish.

2.
Plant Dis ; 2023 Dec 19.
Artigo em Inglês | MEDLINE | ID: mdl-38115567

RESUMO

Newhall navel orange [Citrus sinensis (L.) Osbeck] is an economically important agricultural product in China. In February 2022, a rare lesion symptom was observed on Newhall navel oranges that were harvested from an orchard Ganzhou city, Jiangxi province, China (25.53° N, 114.79° E) and stored for 90 days (18±2℃, 80 to 90% RH) at the Jiangxi Key Laboratory for Postharvest Technology and Non-destructive Testing of Fruits and Vegetables (28.68° N, 115.85° E). Approximately 2% (15/750) of the oranges exhibited symptoms, with normal appearance but ink-black flesh and juice, yellowish lesions on edges of the symptoms, and no unusual odor. To isolate the pathogen, three 5 × 5 mm pieces of symptomatic tissue from a diseased orange were disinfected in 75% ethanol for 30 s, rinsed three times with sterile water, and inoculated on potato dextrose agar (PDA) at 25±1℃ and a 12:12 h photoperiod for 7 days. A pure isolate named ND-hsp was obtained. The colony was light yellow center with pale edge on the top and brown on the bottom. Conidia and pycnidia were observed on PDA medium after 2 months. Conidia were long oval, no septa, 2.9 × 3.4 µm (n = 50), and pycnidia were solitary, 39.4 × 43.9 µm (n = 20), with one or no orifice, brown to dark brown. The morphological characteristics of ND-hsp strain on PDA, oatmeal agar and malt extract agar were similar to those of the Didymellaceae (Aveskamp et al. 2010). Ulteriorly, the genomic DNA of the ND-hsp isolate was extracted from its mycelia using a fungal genomic DNA extraction kit (Solarbio, Beijing, China) for subsequent phylogenetic analyses. Four primer sets, LR0R (Rehner and Samuels 1994) /LR7 (Vilgalys and Hester 1990), V9G (Hoog and Gerrits 1998) /ITS4 (White et al. 1990), Btub2Fd/Btub4Rd (Woudenberg et al. 2009) and RPB2-5F2 (Sung et al. 2007)/RPB2-7cR (Liu et al. 1999) were used to amplify the corresponding DNA fragments of large subunit ribosomal RNA (LSU), internal transcribed spacer region (ITS), beta-tubulin gene (TUB2) and RNA polymerase Ⅱ second largest subunit (RPB2), respectively. The obtained sequences were assigned GenBank accession numbers and showed 99 to 100% identity with their counterparts of Vacuiphoma oculihominis UTHSC DI16-308. A phylogenetic tree was constructed in MEGA 7.0 using the concatenated sequences, placing the isolate within the V. oculihominis clade by 100% bootstrap support. Pathogenicity experiments were performed in triplicate. Ten Newhall navel oranges were surface sterilized with 75% ethanol and inoculated with 15µL of a spore suspension (2×106 spores/ml) into a 3 mm-diameter wound on the equator. The control group received sterile water instead of the spore suspension. Treated and control oranges were incubated at 25±1 ℃ and about 90% relative humidity for 20 days. All oranges were cut longitudinally or transversely through the inoculated wound and examined internally. The oranges inoculated with ND-hsp exhibited ink-black flesh and juice symptoms consistent with the initial oranges. The control oranges remained asymptomatic. Under the Koch's rule, V. oculihominis was reisolated from diseased oranges and kept in Collaborative Innovation Center of Postharvest Key Technology and Quality Safety of Fruits and Vegetables in Jiangxi Province. GenBank database analysis confirms that V. oculihominis has been found in human eye secretions and decayed trees. This is the first report of V. oculihominis as a pathogen on navel oranges in China. Our findings contribute to understanding of citrus fruit pathogens.

3.
Plant Dis ; 2023 Mar 01.
Artigo em Inglês | MEDLINE | ID: mdl-36856645

RESUMO

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.

4.
Plant Dis ; 2023 Jul 16.
Artigo em Inglês | MEDLINE | ID: mdl-37455409

RESUMO

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.

5.
Plant Dis ; 2023 May 15.
Artigo em Inglês | MEDLINE | ID: mdl-37189040

RESUMO

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.

6.
Plant Dis ; 2023 Jan 23.
Artigo em Inglês | MEDLINE | ID: mdl-36691267

RESUMO

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.

7.
Plant Dis ; 2023 Sep 05.
Artigo em Inglês | MEDLINE | ID: mdl-37669177

RESUMO

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.

8.
BMC Plant Biol ; 22(1): 108, 2022 Mar 09.
Artigo em Inglês | MEDLINE | ID: mdl-35264115

RESUMO

BACKGROUND: Potassium (K) is important in the regulation of plant growth and development. It is the most abundant mineral element in kiwifruit, and its content increases during fruit ripening. However, how K+ transporter works in kiwifruit postharvest maturation is not yet clear. RESULTS: Here, 12 K+ transporter KT/HAK/KUP genes, AcKUP1 ~ AcKUP12, were isolated from kiwifruit, and their phylogeny, genomic structure, chromosomal location, protein properties, conserved motifs and cis-acting elements were analysed. Transcription analysis revealed that AcKUP2 expression increased rapidly and was maintained at a high level during postharvest maturation, consistent with the trend of K content; AcKUP2 expression was induced by ethylene, suggesting that AcKUP2 might play a role in ripening. Fluorescence microscopy showed that AcKUP2 is localised in the plasma membrane. Cis-elements, including DER or ethylene response element (ERE) responsive to ethylene, were found in the AcKUP2 promoter sequence, and ethylene significantly enhanced the AcKUP2 promoter activity. Furthermore, we verified that AcERF15, an ethylene response factor, directly binds to the AcKUP2 promoter to promote its expression. Thus, AcKUP2 may be an important potassium transporter gene which involved in ethylene-regulated kiwifruit postharvest ripening. CONCLUSIONS: Therefore, our study establishes the first genome-wide analysis of the kiwifruit KT/HAK/KUP gene family and provides valuable information for understanding the function of the KT/HAK/KUP genes in kiwifruit postharvest ripening.


Assuntos
Actinidia/crescimento & desenvolvimento , Actinidia/genética , Etilenos/metabolismo , Frutas/crescimento & desenvolvimento , Frutas/genética , Regulação da Expressão Gênica de Plantas/efeitos dos fármacos , Antiportadores de Potássio-Hidrogênio/metabolismo , China , Produtos Agrícolas/genética , Produtos Agrícolas/crescimento & desenvolvimento , Genes de Plantas , Desenvolvimento Vegetal/efeitos dos fármacos , Desenvolvimento Vegetal/genética , Antiportadores de Potássio-Hidrogênio/genética
9.
Microb Cell Fact ; 21(1): 38, 2022 Mar 12.
Artigo em Inglês | MEDLINE | ID: mdl-35279161

RESUMO

BACKGROUND: Lytic polysaccharide monooxygenases (LPMOs) belonging to the auxiliary activity 9 family (AA9) are widely found in aerobic fungi. These enzymes are O2-dependent copper oxidoreductases that catalyze the oxidative cleavage of cellulose. However, studies that have investigated AA9 LPMOs of aerobic fungi in the herbivore gut are scare. To date, whether oxidative cleavage of cellulose occurs in the herbivore gut is unknown. RESULTS: We report for the first time experimental evidence that AA9 LPMOs from aerobic thermophilic fungi catalyze the oxidative cleavage of cellulose present in the horse gut to C1-oxidized cellulose and C1- and C4-oxidized cello-oligosaccharides. We isolated and identified three thermophilic fungi and measured their growth and AA9 LPMO expression at 37 °C in vitro. We also assessed the expression and the presence of AA9 LPMOs from thermophilic fungi in situ. Finally, we used two recombinant AA9 LPMOs and a native AA9 LPMO from thermophilic fungi to cleave cellulose to yield C1-oxidized products at 37 °C in vitro. CONCLUSIONS: The oxidative cleavage of cellulose occurs in the horse gut. This finding will broaden the known the biological functions of the ubiquitous LPMOs and aid in determining biological significance of aerobic thermophilic fungi.


Assuntos
Celulose , Oxigenases de Função Mista , Animais , Celulose/metabolismo , Cavalos , Oxigenases de Função Mista/metabolismo , Estresse Oxidativo , Oxirredutases/metabolismo , Polissacarídeos/metabolismo
10.
Phytopathology ; 112(2): 355-363, 2022 Feb.
Artigo em Inglês | MEDLINE | ID: mdl-34165320

RESUMO

A 207-amino-acid residue endoglucanohydrolase (EG1) belonging to the glycoside hydrolase 45 (GH45) from Rhizoctonia solani acts as a pathogen-associated molecular pattern (PAMP). However, the mechanism of EG1 inducing plant immunity is unclear. Here, we found that EG1 contains two domains related to its PAMP function. Transient expression showed that EG1-1, the mutation deleting 60 amino acid residues from the N-terminal, still reserved the PAMP function. Further truncation of EG1-1 obtained two truncating mutations: EG1-2, deleting seven amino acid residues from the N-terminal of EG1-1 (SPWAVND), and EG1-3, deleting five amino acid residues from the C-terminal of EG1-1 (GCSRK). Transient expression showed that the two truncating mutations EG1-2 and EG1-3 all lost the PAMP function. Site-directed mutagenesis of EG1-1 showed that the three amino acid residues (P, W, and D) in the region SPWAVND and the two amino acid residues (C and R) in the region GCSRK were involved in the PAMP function. The homology model showed that the two regions were located at a surface on the EG1 and structurally independent. These results demonstrate that there are two functional regions for the plant immune function of the EG1 released by R. solani, and the two functional regions are independent of each other.


Assuntos
Glicosídeo Hidrolases , Moléculas com Motivos Associados a Patógenos , Domínio Catalítico , Doenças das Plantas/genética , Rhizoctonia/genética
11.
Plant Dis ; 2022 Mar 06.
Artigo em Inglês | MEDLINE | ID: mdl-35253487

RESUMO

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.

12.
Plant Dis ; 2022 Jul 08.
Artigo em Inglês | MEDLINE | ID: mdl-35802012

RESUMO

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.

13.
Plant Dis ; 2022 Aug 19.
Artigo em Inglês | MEDLINE | ID: mdl-35984392

RESUMO

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.

14.
Plant Dis ; 2022 Apr 05.
Artigo em Inglês | MEDLINE | ID: mdl-35380465

RESUMO

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.

15.
J Sci Food Agric ; 102(8): 3140-3149, 2022 Jun.
Artigo em Inglês | MEDLINE | ID: mdl-34791654

RESUMO

BACKGROUND: Phomopsis stem-end rot caused by Diaporthe citri, causes significant commercial postharvest losses of pummelo fruit during storage. Carvacrol (CVR) is a known generally recognized as safe and has the ability to prolong the preservation of harvested fruits. In the present study, the inhibitory effects of CVR treatment at the appropriate concentration on Phomopsis stem-end rot development of harvested pummelo fruit inoculated with D. citri were evaluated by the amounts of cell wall components, the activities and gene expressions of related enzymes involved in cell wall modification and lignin biosynthesis. RESULTS: Results indicated that CVR completely inhibited D. citri growth in vitro at 200 mg L-1 and significantly controlled Phomopsis stem-end rot development in harvested pummelo. The CVR treatment delayed peel softening and browning, and retarded electrolyte leakage, superoxide radical (O2 •- ) production, and malondialdehyde content. The CVR-treated fruit maintained higher amounts of cell wall material, protopectin, hemicelluloses, and cellulose, but exhibited lower water-soluble pectin amount. Moreover, in D. citri-inoculated fruit, CVR treatment suppressed the activities and gene expressions of cell wall disassembling-enzymes, including pectin methylesterase, polygalacturonase, cellulase, and ß-galactosidase, while the development of cell wall degradation was reduced. Meanwhile, the CVR treatment enhanced the lignin biosynthesis by increasing the activities and up-regulating the gene expressions of phenylalanine ammonialyase, cinnamic alcohol dehydrogenase, and peroxidase accompanied with elevated level of lignin in pummelo fruit. CONCLUSION: The disease resistance to D. citri in pummelo fruit elicited by CVR treatment is related to delaying cell wall degradation and enhancing lignin biosynthesis. © 2021 Society of Chemical Industry.


Assuntos
Citrus , Frutas , Ascomicetos , Parede Celular/metabolismo , Citrus/metabolismo , Cimenos , Resistência à Doença , Lignina/metabolismo
16.
Protein Expr Purif ; 183: 105865, 2021 07.
Artigo em Inglês | MEDLINE | ID: mdl-33675938

RESUMO

The tresI gene of Myxococcus sp. strain V11 was cloned, and found to encode a trehalose synthase comprising 551 amino acids. The deduced molecular weight of the encoded TreS I protein 64.7 kDa and the isoelectric point (pI) was predicted to be 5.6. The catalytic cleft consists of the Asp202-Glu244-Asp310 catalytic triad and additional conserved residues. The recombinant (His)6-tag enzyme was expressed in Escherichia coli BL21(DE3) and purified by Ni2+-affinity chromatography, resulting in a specific activity of up to 172.7 U/mg. TLC and HPLC results confirmed that rTreS I can convert maltose into trehalose, with a yield of 61%. The KM and Vmax values of recombinant TreS I for maltose were 0.62 mM and 25.5 mM min-1 mg-1 protein, respectively. TreS I was optimally active at 35° and stable at temperatures of <25 °C. TreS I was stable within a narrow range of pH values, from 6.0 to 7.0. The enzymatic activity was slightly stimulated by Mg2+ and strongly inhibited by Fe3+, Co2+ and Cu2+. TreS I was also strongly inhibited by SDS and weakly by EDTA and TritonX-100.


Assuntos
Proteínas de Bactérias , Clonagem Molecular , Glucosiltransferases , Myxococcus , Proteínas de Bactérias/biossíntese , Proteínas de Bactérias/química , Proteínas de Bactérias/genética , Escherichia coli/genética , Escherichia coli/metabolismo , Glucosiltransferases/biossíntese , Glucosiltransferases/química , Glucosiltransferases/genética , Myxococcus/enzimologia , Myxococcus/genética , Proteínas Recombinantes/biossíntese , Proteínas Recombinantes/química , Proteínas Recombinantes/genética
17.
Plant Cell Rep ; 40(12): 2273-2286, 2021 Dec.
Artigo em Inglês | MEDLINE | ID: mdl-34269828

RESUMO

Rising atmospheric CO2 is a key driver of climate change, intensifying drastic changes in meteorological parameters. Plants can sense and respond to changes in environmental parameters including atmospheric CO2 and temperatures. High temperatures beyond the physiological threshold can significantly affect plant growth and development and thus attenuate crop productivity. However, elevated atmospheric CO2 can mitigate the deleterious effects of heat stress on plants. Despite a large body of literature supporting the positive impact of elevated CO2 on thermotolerance, the underlying biological mechanisms and precise molecular pathways that lead to enhanced tolerance to heat stress remain largely unclear. Under heat stress, elevated CO2-induced expression of respiratory burst oxidase homologs (RBOHs) and reactive oxygen species (ROS) signaling play a critical role in stomatal movement, which optimizes gas exchange to enhance photosynthesis and water use efficiency. Notably, elevated CO2 also fortifies antioxidant defense and redox homeostasis to alleviate heat-induced oxidative damage. Both hormone-dependent and independent pathways have been shown to mediate high CO2-induced thermotolerance. The activation of heat-shock factors and subsequent expression of heat-shock proteins are thought to be the essential mechanism downstream of hormone and ROS signaling. Here we review the role of phytohormones in plant response to high atmospheric CO2 and temperatures. We also discuss the potential mechanisms of elevated CO2-induced thermotolerance by focusing on several key phytohormones such as ethylene. Finally, we address some limitations of our current understanding and the need for further research to unveil the yet-unknown crosstalk between plant hormones in mediating high CO2-induced thermotolerance in plants.


Assuntos
Dióxido de Carbono , Reguladores de Crescimento de Plantas/fisiologia , Fenômenos Fisiológicos Vegetais , Termotolerância/fisiologia , Proteínas de Choque Térmico/metabolismo , Resposta ao Choque Térmico/fisiologia , Folhas de Planta/química , Folhas de Planta/fisiologia , Proteínas de Plantas/metabolismo , Estômatos de Plantas/fisiologia , Espécies Reativas de Oxigênio/metabolismo
18.
Pestic Biochem Physiol ; 165: 104505, 2020 May.
Artigo em Inglês | MEDLINE | ID: mdl-32359555

RESUMO

Blue mold caused by Penicillium italicum is an important postharvest disease of citrus fruit. The antifungal activity of a flavonone pinocembroside compound obtained from the fruit of Ficus hirta Vahl., was evaluated against P. italicum. Pinocembroside showed antifungal activity against in vitro mycelial growth of P. italicum, with the minimum inhibitory concentration (MIC) and minimum fungicidal concentration (MFC) of 200 and 800 mg/L, respectively. The blue mold development on 'Newhall' navel oranges was inhibited by pinocembroside in a dose-dependent manner. Moreover, pinocembroside might exert its antifungal activity via membrane-targeted mechanism with increasing membrane permeability, reduction of antioxidant enzyme activity and acceleration of lipid peroxidation in the pathogen. This pioneering study suggested that pinocembroside suppressed postharvest blue mold by direct inhibition of P. italicum mycelial growth via membrane-targeting mechanism, thus providing a novel mode of action against traditional fungicides for controlling blue mold of citrus fruit.


Assuntos
Citrus sinensis , Citrus , Fungicidas Industriais , Penicillium , Frutas
19.
J Sci Food Agric ; 100(5): 2280-2287, 2020 Mar 30.
Artigo em Inglês | MEDLINE | ID: mdl-31944323

RESUMO

BACKGROUND: Hydrogen sulfide (H2 S) is a known signaling molecule in plants, which has the ability to delay fruit ripening. Our previous studies have shown that H2 S treatment could delay the maturation of kiwifruits by inhibiting ethylene production, improving protective enzyme activities, and decreasing the accumulation of reactive oxygen species to protect the cell membrane during storage. The mechanism related to the way in which H2 S affected kiwifruit maturation was still unclear. We performed transcriptome sequencing to explore the influences of H2 S on the softening of kiwifruit. RESULTS: The firmness and the soluble solids content (SSC) of the kiwifruit were significantly better maintained with H2 S treatment compared to the control during the storage period (P < 0.05). Transmission electron microscopy (TEM) showed that degradation of the cell wall was inhibited after H2 S treatment. Based on transcriptome data analysis and quantitative real-time polymerase chain reaction (qRT-PCR), expression levels of endo-1,4-ß-glucanase (ß-glu), ß-galactosidase (ß-gal) and pectinesterase (PME) decreased whereas pectinesterase inhibitor (PMEI) significantly increased in response to H2 S. The members of the signal transduction pathway involved in ethylene were also identified. Hydrogen sulfide inhibited the expression of ethylene receptor 2 (ETR2), ERF003, ERF5, and ERF016, and increased the expression of ethylene-responsive transcription factor 4 (ERF4) and ERF113. CONCLUSION: Hydrogen sulfide could delay the ripening and senescence of kiwifruit by regulating the cell-wall degrading enzyme genes and affecting ethylene signal transduction pathway genes. Our results revealed the effect of H2 S treatment on the softening of kiwifruit at the transcription level, laying a foundation for further research. © 2020 Society of Chemical Industry.


Assuntos
Actinidia/química , Parede Celular/metabolismo , Frutas/metabolismo , Perfilação da Expressão Gênica/métodos , Sulfeto de Hidrogênio/metabolismo , Parede Celular/ultraestrutura , Etilenos , Regulação da Expressão Gênica de Plantas/efeitos dos fármacos , Sulfeto de Hidrogênio/farmacologia , Proteínas de Plantas/metabolismo , Receptores de Superfície Celular/metabolismo , Transdução de Sinais , Transcriptoma
20.
Appl Opt ; 58(7): 1631-1638, 2019 Mar 01.
Artigo em Inglês | MEDLINE | ID: mdl-30874195

RESUMO

Laser-induced breakdown spectroscopy (LIBS) combined with pattern recognition was proposed to discriminate rice species. LIBS spectra in the range of 210-480 nm wavelength from 11 different rice species were collected and preprocessed. Principal component analysis was applied to extract the characteristic variables from LIBS spectral data. Three pattern recognition methods, discriminant analysis, radial basis function neural network, and multi-layer perceptron neural network (MLP) were performed to compare the precision in identifying rice species. The results showed that the performance of the MLP model was better. The average identification rate of rice species reached 100% and 97.9% in the training and test sets, respectively, with MLP. The highest and lowest percentages for correct identification were 100% for early indica rice, Huai rice 5, Yan japonica 6, Lian japonica 8, Xuhan 1, Lvhan 1, Sheng rice 16, Yang japonica 687, and Fenghan 30, and 77.8% for Wuyu japonica rice in test sets. The overall results demonstrate that LIBS combined with MLP could be utilized to rapidly discriminate rice species.

SELEÇÃO DE REFERÊNCIAS
Detalhe da pesquisa