Your browser doesn't support javascript.
loading
Mostrar: 20 | 50 | 100
Resultados 1 - 20 de 77
Filtrar
1.
Plant Physiol ; 194(4): 2249-2262, 2024 Mar 29.
Artículo en Inglés | MEDLINE | ID: mdl-38109500

RESUMEN

Desiccation is typically fatal, but a small number of land plants have evolved vegetative desiccation tolerance (VDT), allowing them to dry without dying through a process called anhydrobiosis. Advances in sequencing technologies have enabled the investigation of genomes for desiccation-tolerant plants over the past decade. However, a dedicated and integrated database for these valuable genomic resources has been lacking. Our prolonged interest in VDT plant genomes motivated us to create the "Drying without Dying" database, which contains a total of 16 VDT-related plant genomes (including 10 mosses) and incorporates 10 genomes that are closely related to VDT plants. The database features bioinformatic tools, such as blast and homologous cluster search, sequence retrieval, Gene Ontology term and metabolic pathway enrichment statistics, expression profiling, co-expression network extraction, and JBrowser exploration for each genome. To demonstrate its utility, we conducted tailored PFAM family statistical analyses, and we discovered that the drought-responsive ABA transporter AWPM-19 family is significantly tandemly duplicated in all bryophytes but rarely so in tracheophytes. Transcriptomic investigations also revealed that response patterns following desiccation diverged between bryophytes and angiosperms. Combined, the analyses provided genomic and transcriptomic evidence supporting a possible divergence and lineage-specific evolution of VDT in plants. The database can be accessed at http://desiccation.novogene.com. We expect this initial release of the "Drying without Dying" plant genome database will facilitate future discovery of VDT genetic resources.


Asunto(s)
Briófitas , Desecación , Proteínas de Plantas/genética , Proteínas de Plantas/metabolismo , Plantas/metabolismo , Genoma de Planta/genética , Transcriptoma/genética , Briófitas/genética
2.
Plant J ; 115(4): 1100-1113, 2023 08.
Artículo en Inglés | MEDLINE | ID: mdl-37177875

RESUMEN

Phyllosphere-associated microbes play a crucial role in plant-pathogen interactions while their composition and diversity are strongly influenced by drought stress. As dioecious plant species exhibited secondary dimorphism between the two sexes in response to drought stress, whether such difference will lead to sex-specific differences in phyllosphere microbiome and associated pathogen resistance between male and female conspecifics is still unknown. In this study, we subjected female and male full siblings of a dioecious poplar species to a short period of drought treatment followed by artificial infection of a leaf pathogenic fungus. Our results showed that male plants grew better than females with or without drought stress. Female control plants had more leaf lesion area than males after pathogen infection, whereas drought stress reversed such a difference. Further correlation and in vitro toxicity tests suggested that drought-mediated sexual differences in pathogen resistance between the two plant sexes could be attributed to the shifts in structure and function of phyllosphere-associated microbiome rather than the amount of leaf main defensive chemicals contained in plant leaves. Supportively, the microbiome analysis through high-throughput sequencing indicated that female phyllosphere enriched a higher abundance of ecologically beneficial microbes that serve as biological plant protectants, while males harbored abundant phytopathogens under drought-stressed conditions. The results could provide potential implications for the selection of suitable poplar sex to plants in drought or semi-drought habitats.


Asunto(s)
Microbiota , Populus , Sequías , Hojas de la Planta/fisiología , Hongos , Populus/genética
3.
Int J Mol Sci ; 25(1)2024 Jan 01.
Artículo en Inglés | MEDLINE | ID: mdl-38203739

RESUMEN

Arthrinium phaeospermum is the major pathogen responsible for the significant stem disease "blight" in B. pervariabilis × D. grandis. The interacting proteins of the key pathogenic factor ApCtf1ß, BDUbc and BDSKL1, have previously been obtained by two-hybrid, BiFC, GST pull-down yeast assays. However, the functions of these interacting proteins remain unknown. This study successfully obtained transgenic plants overexpressing BDUbc, BDSKL1, and BDUbc + BDSKL1 via Agrobacterium-mediated gene overexpression. qRT-PCR analysis revealed significantly increased expression levels of BDUbc and BDSKL1 in the transgenic plants. After infection with the pathogenic spore suspension, the disease incidence and severity index significantly decreased across all three transgenic plants, accompanied by a marked increase in defense enzyme levels. Notably, the co-transformed plant, OE-BDUbc + BDSKL1, demonstrated the lowest disease incidence and severity index among the transgenic variants. These results not only indicate that BDUbc and BDSKL1 are disease-resistant genes, but also that these two genes may exhibit a synergistic enhancement effect, which further improves the resistance to blight in Bambusa pervariabilis × Dendrocalamopsis grandis.


Asunto(s)
Bambusa , Queratoconjuntivitis , Agrobacterium , Bioensayo , Plantas Modificadas Genéticamente , Saccharomyces cerevisiae
4.
Planta ; 257(6): 109, 2023 May 05.
Artículo en Inglés | MEDLINE | ID: mdl-37145304

RESUMEN

MAIN CONCLUSION: Serine/arginine-rich (SR) proteins participate in RNA processing by interacting with precursor mRNAs or other splicing factors to maintain plant growth and stress responses. Alternative splicing is an important mechanism involved in mRNA processing and regulation of gene expression at the posttranscriptional level, which is the main reason for the diversity of genes and proteins. The process of alternative splicing requires the participation of many specific splicing factors. The SR protein family is a splicing factor in eukaryotes. The vast majority of SR proteins' existence is an essential survival factor. Through its RS domain and other unique domains, SR proteins can interact with specific sequences of precursor mRNA or other splicing factors and cooperate to complete the correct selection of splicing sites or promote the formation of spliceosomes. They play essential roles in the composition and alternative splicing of precursor mRNAs, providing pivotal functions to maintain growth and stress responses in animals and plants. Although SR proteins have been identified in plants for three decades, their evolutionary trajectory, molecular function, and regulatory network remain largely unknown compared to their animal counterparts. This article reviews the current understanding of this gene family in eukaryotes and proposes potential key research priorities for future functional studies.


Asunto(s)
Proteínas de Unión al ARN , Serina , Animales , Proteínas de Unión al ARN/genética , Proteínas de Unión al ARN/metabolismo , Serina/genética , Serina/metabolismo , Proteínas Nucleares/genética , Empalme del ARN/genética , Empalme Alternativo/genética , Precursores del ARN/genética , Precursores del ARN/metabolismo , Proteínas de Plantas/metabolismo , ARN Mensajero/genética , ARN Mensajero/metabolismo , Factores de Empalme de ARN/metabolismo , Arginina
5.
J Exp Bot ; 74(6): 2188-2202, 2023 03 28.
Artículo en Inglés | MEDLINE | ID: mdl-36738293

RESUMEN

Microorganisms associated with the phyllosphere play a crucial role in protecting plants from diseases, and their composition and diversity are strongly influenced by heavy metal contaminants. Dioecious plants are known to exhibit sexual dimorphism in metal accumulation and tolerance between male and female individuals. Hence, in this study we used male and female full-siblings of Populus deltoides to investigate whether the two sexes present differences in their phyllosphere microbiome structures and in their associated resistance to the leaf pathogenic fungus Pestalotiopsis microspora after exposure to excess soil cadmium (Cd). We found that Cd-treated male plants grew better and accumulated more leaf Cd than females. Cd stress reduced the lesion areas on leaves of both sexes after pathogen infection, but male plants exhibited better resistance than females. More importantly, Cd exposure differentially altered the structure and function of the phyllosphere microbiomes between the male and female plants, with more abundant ecologically beneficial microbes and decreased pathogenic fungal taxa harbored by male plants. In vitro toxicity tests suggested that the sexual difference in pathogen resistance could be attribute to both direct Cd toxicity and indirect shifts in the phyllosphere microbiome. This study provides new information relevant for understanding the underlying mechanisms of the effects of heavy metals involved in plant-pathogen interactions.


Asunto(s)
Metales Pesados , Microbiota , Populus , Cadmio/toxicidad , Suelo , Hongos
6.
J Exp Bot ; 74(3): 976-990, 2023 02 05.
Artículo en Inglés | MEDLINE | ID: mdl-36346205

RESUMEN

Plants have evolved a two-layer immune system comprising pattern-triggered immunity (PTI) and effector-triggered immunity (ETI) that is activated in response to pathogen invasion. Microbial patterns and pathogen effectors can be recognized by surface-localized pattern-recognition receptors (PRRs) and intracellularly localized nucleotide-binding leucine-rich repeat receptors (NLRs) to trigger PTI and ETI responses, respectively. At present, the metabolites activated by PTI and ETI and their roles and signalling pathways in plant immunity are not well understood. In this study, metabolomic analysis showed that ETI and PTI induced various flavonoids and amino acids and their derivatives in plants. Interestingly, both glutathione and neodiosmin content were specifically up-regulated by ETI and PTI, respectively, which significantly enhanced plant immunity. Further studies showed that glutathione and neodiosmin failed to induce a plant immune response in which PRRs/co-receptors were mutated. In addition, glutathione-reduced mutant gsh1 analysis showed that GSH1 is also required for PTI and ETI. Finally, we propose a model in which glutathione and neodiosmin are considered signature metabolites induced in the process of ETI and PTI activation in plants and further continuous enhancement of plant immunity in which PRRs/co-receptors are needed. This model is beneficial for an in-depth understanding of the closed-loop mode of the positive feedback regulation of PTI and ETI signals at the metabolic level.


Asunto(s)
Inmunidad de la Planta , Plantas , Retroalimentación , Plantas/metabolismo , Transducción de Señal , Receptores de Reconocimiento de Patrones/metabolismo , Enfermedades de las Plantas
7.
Plant Cell Rep ; 43(1): 13, 2023 Dec 23.
Artículo en Inglés | MEDLINE | ID: mdl-38135780

RESUMEN

KEY MESSAGE: ZmMRPA6 was cloned and characterized as the first ATP-binding cassette (ABC) transporter in maize to be proven to participate in cold and salt tolerance. Homologous genes AtABCC4 and AtABCC14 of ZmMRPA6 also responded to salt stress. ATP-binding cassette (ABC) proteins are major transmembrane transporters that play significant roles in plant development against various abiotic stresses. However, available information regarding stress-related ABC genes in maize is minimal. In this study, a maize ABC transporter gene, ZmMRPA6, was identified through genome-wide association analysis (GWAS) for cold tolerance in maize seeds germination and functionally characterized. During germination and seedling stages, the zmmrpa6 mutant exhibited enhanced resistance to cold or salt stress. Mutated of ZmMRPA6 did not affect the expression of downstream response genes related cold or salt response at the transcriptional level. Mass spectrometry analysis revealed that most of the differential proteins between zmmrpa6 and wild-type plants were involved in response to stress process including oxidative reduction, hydrolase activity, small molecule metabolism, and photosynthesis process. Meanwhile, the plants which lack the ZmMRPA6 homologous genes AtABCC4 or AtABCC14 were sensitive to salt stress in Arabidopsis. These results indicated that ZmMRPA6 and its homologous genes play a conserved role in cold and salt stress, and functional differentiation occurs in monocotyledonous and dicotyledonous plants. In summary, these findings dramatically improved our understanding of the function of ABC transporters resistance to abiotic stresses in plants.


Asunto(s)
Arabidopsis , Zea mays , Zea mays/genética , Zea mays/metabolismo , Tolerancia a la Sal/genética , Transportadoras de Casetes de Unión a ATP/genética , Transportadoras de Casetes de Unión a ATP/metabolismo , Estudio de Asociación del Genoma Completo , Plantas Modificadas Genéticamente/genética , Estrés Salino , Arabidopsis/genética , Estrés Fisiológico/genética , Adenosina Trifosfato/metabolismo , Regulación de la Expresión Génica de las Plantas/genética , Proteínas de Plantas/genética , Proteínas de Plantas/metabolismo , Frío
8.
Plant Dis ; 2023 Apr 05.
Artículo en Inglés | MEDLINE | ID: mdl-37018208

RESUMEN

Walnut (Juglans regia) is a deciduous tree of the Juglandaceae family, widely cultivated in China, and provides value in a variety of ways, including the usage of the wood and nuts, and offers substantial economic, social, and environmental advantages (Wang et al, 2017). Nevertheless, a fungal disease of causing walnut trunk rot was observed in approximately 30% of 50 counted ten-year-old J. regia in Chongzhou City (30°33'34″N, 103°38'35″E, 513 m), Sichuan Province, China, and this disease has greatly delete healthy growth of walnut. The infected bark exhibited purple necrotic lesions, and the sick parts were surrounded by water-soaked plaques. From 10 trunks of the 10 diseased trees, 20 isolated fungal colonies were the same. The ascospores placed in 60 mm plates were almost entirely covered with mycelium within 8 days, colonies on the PDA changed from initial pale to white, ad then turned yellowish to light orange or rosy to yellow-brown (25℃, 90% relative humidity, 12-h photoperiod). On the host, Ectostromata were immersed to erumpent, globose to subglobose, purple and brown, and measured 0.6 - 4.5 × 0.3 - 2.8 mm (x̄ = 2.6 × 1.6 mm, n = 40); Ascomata were flask-shaped to subglobose, dark brown, and measured 0.1 - 0.6 × 0.1 - 0.4 mm (x̄ = 0.35 × 0.25 mm, n = 40); Asci were numerous, cylindrical to subclavate, contained 8 uniseriate ascospores, and measured 80 - 150 × 10 - 20 µm (x̄ = 115 × 15 µm, n = 40), and Ascospores were ellipsoid, 2-celled, dark brown to black, plump or attenuated towards, apices with 1 large drop per cell, and measured 14 - 20 × 6.5 - 9 µm (x̄ = 17 × 7.8 µm, n = 40). These morphological characteristics are consistent with the species Myrmaecium fulvopruinatum (Berk.) Jaklitsch & Voglmayr (Jaklitsch et al. 2015). The genomic DNA of a representative isolate SICAUCC 22-0148 was extracted. The ITS, LSU region, tef1-α, rpb2 genes region were amplified using the primer pairs ITS1/ITS4 primers (White et al. 1990), LR0R/LR5 (Moncalvo et al. 1995), EF1-688F/986R (Alves et al. 2008), fRPB2-5f/fRPB2-7cr (Liu et al. 1999), respectively. The sequences were deposited in NCBI with accession numbers ON287043 (ITS), ON287044 (LSU), ON315870 (tef1-α), and ON315871 (rpb2), rspectively, which showed 99.8, 99.8, 98.1, and 98.5% identities with M. fulvopruinatum CBS 139057 holotype (accession numbers KP687858, KP687858, KP688027, and KP687933 respectively). Based on the analyses of phylogenies and morphologies, the isolates were identified as M. fulvopruinatum. The pathogenicity of SICAUCC 22-0148 was tested by inoculating surface-sterilized trunk wounds of four-year-old trees of J. regia with a mycelial plug (Desai et al. 2019). Sterile PDA plugs were used as controls. Wounds were covered with a film, to ensure humidity and prevent contamination. Each inoculation was repeated twice and included two plants, control and inoculated. A month later, the symptoms observed on inoculated trunks were similar to those in the wild, and M. fulvopruinatum was re-isolated from the inoculated trunk, confirming Koch's postulates. Previous research has reported M. fulvopruinatum as an important fungal species that cause canker delete symptoms on Chinese sweet chestnut in China (Jiang et al. 2018). We carried the taxonomy work of the fungi that caused trunk rot on walnut, and this is the first time that M. fulvopruinatum has been linked to walnut trunk rot on J. regia. Trunk rot of walnut will not only cause weakening of trees, but also affect the yield and quality of walnuts, bringing huge economic losses. This study was supported by the Sichuan Science and Technology Program under Grant 2022NSFSC1011. References: Alves, A., et al. 2008. Fungal Diversity 28:1-13. Desai, D.D., et al. 2019. International Journal of Economic Plants 6:147-149. Jaklitsch., W.M., et al. 2015. Fungal Diversity 73(1):159-202. Jiang, N., et al. 2018. Mycosphere 9(6):1268-1289. Liu, Y.L., et al. 1999. Mol Biol Evol 16:1799-1808. Moncalvo, J.M., et al. 1995. Mycologia 87:223-238. Wang, Q.H., et al. 2017. Australasian Plant Pathology 46:585-595. White, T.J., et al. 1990. Page 315 in: PCR Protocols: A Guide to Methods and Applications. Academic Press, San Diego, CA.

9.
Plant Dis ; 2023 Jan 06.
Artículo en Inglés | MEDLINE | ID: mdl-36607332

RESUMEN

Prunus sibirica L. (Siberian apricot) is a member of the Rosaceae family and an ecologically important tree species in China (Buer et al., 2022). Shot hole symptoms on the leaves were observed in five Siberian apricot groves in Chengdu (103.81 E, 30.97 N), Sichuan province in July 2020. The symptoms first appeared as small purplish-brown spots with yellow rings around them. As the disease progressed, the damaged area (diameter 1.5-3.0 cm) became necrotic and fell off. The disease incidence was about 60% and the disease index was 28.6 of leaves in the grove. in most severe cases. Fifteen symptomatic leaves were collected from 5 different trees in an orchard. Pathogen isolation was performed from symptomatic leaf tissue (5 × 5 mm) though surface disinfection (in 70% ethanol and 2% NaClO) and incubation on Potato Dextrose Agar (PDA) at 28℃ for 3 days. Overall 10 isolates with similar colony morphology were obtained from the 15 infected tissue pieces, and three representative isolates (XCK 2-4) were selected for further study. Colonies of the isolates on PDA were initially cottony, pale white to grayish-green with abundant aerial hyphae and produced conidial masses after 7 days. Conidiogenous cells were clavate and aggregated in acervuli. Conidia were smooth-walled, single-celled, straight, and slightly obtusely rounded at both ends, 12.8 to 18.7 × 4.3 to 5.7 µm in size (Fig. 1). The morphological characteristics of the three isolates were consistent with the description of species in the Colletotrichum gloeosporioides complex. DNA was amplified using the following primers pairs for the internal transcribed spacer (ITS) region of rDNA and partial sequences of beta-tubulin (TUB2), glyceraldehyde-3-phosphate dehydrogenase (GAPDH), chitin synthase (CHS-1), and translation elongation factor (TEF-1), respectively: ITS1/ITS4, T1/Bt2b, GDF/GDR, CHS-F/CHS-R, and EF-F/EF-R (Vieira et al., 2014). Accession numbers (MW228049, MW284974, MW284976, MW284975 and MW284977, respectively) were obtained afterepositing all the resulting sequences in GenBank. Nucleotide blast showed 99 to 100% identities with Colletotrichum fructicola (GenBank accessions nos. MZ961683, MW284974, MN525881, MN525860, MF627961). Phylogenetic analysis of combined ITS-TUB-GAPDH genes using the Mrbayes inference method showed that the three isolates clustered with three reference isolates of C. fructicola as a distinct clade (Fig. 2). To verify Koch's postulates, ten 3-year-old healthy potted plants of P. sibirica were inoculated by spraying a conidial suspension (6 × 105 conidia/mL) of isolate XCK2 on both sides of leaves, and the control leaves were sprayed with sterile water. Then, all treatments were placed in a moist environment (25±2°C, 80% relative humidity, natural light). The inoculated plants showed typical symptoms of plants with natural infections, while the controls remained asymptomatic after 14 days. The pathogen C. fructicola was re-isolated from all inoculated plants, and the culture and fungus characteristics were the same as those of the original isolate. Colletotrichum fructicola was not isolated from the control plants. The results indicated that C. fructicola is the causal agent of the disease. Colletotrichum fructicola was reported as a leaf pathogen on Camellia chrysantha in China (Zhao et al., 2021). This is the first report of C. fructicola causing P. sibirica leaf shot-hole in the world. The identification of C. fructicola could provide relevant information for applying management strategies and research on the Siberian apricot disease.

10.
Plant Dis ; 2023 May 04.
Artículo en Inglés | MEDLINE | ID: mdl-37142964

RESUMEN

Taxus chinensis var. mairei is the endemic, endangered, and first-class protected tree species in China. This species is considered as an important resource plant because it can produce Taxol which is an effective medicinal compound against various cancers (Zhang et al., 2010). Stem blight was observed in two plant nurseries in Ya'an (102°44'E,30°42'N), Sichuan province in April 2021. The symptoms first appeared as round brown spots on the stem. As the disease progressed, the damaged area gradually expanded into an oval or irregular shape, which was dark brown. About 800 square meters of planting area were investigated and the disease incidence was up to approximately 64.8%. Twenty obviously symptomatic stems which exhibited the same symptoms as above were collected from 5 different trees in the nursery. To isolate the pathogen, the symptom margin was cut into small blocks (5 x 5 mm), and the blocks were surface sterilized in 75% ethanol for 90 s and 3% NaClO solution for 60 s . Finally incubated on Potato Dextrose Agar (PDA) at 28℃ for 5 days. Ten pure cultures were isolated by transferring hyphal and the three strains (HDS06, HDS07 and HDS08) were selected as representative isolates for further study. Initially, colonies on the PDA of three isolates were white and cotton-like, and then gradually turned gray-black from the center. After 21 days, conidia were produced and were smooth-walled, single-celled, black, oblate, or spherical, measuring 9.3 to 13.6 × 10.1 to 14.5 µm in size (n = 50). Conidia were present at the tip of conidiophores on hyaline vesicles. These morphological features were generally consistent with those of N. musae (Wang et al., 2017). To validate the identification, DNA were extracted from the three isolates, followed by the amplification of transcribed spacer region of rDNA (ITS), the translation elongation factor EF-1 (TEF-1), and the Beta-tubulin (TUB2) sequences with the respective primer pairs ITS1/ITS4 (White et al., 1990), EF-728F/EF-986R (Vieira et al., 2014) and Bt2a/Bt2b (O'Donnell et al., 1997) .The sequences were deposited in GenBank with the accession numbers ON965533, OP028064, OP028068, OP060349, OP060353, OP060354, OP060350, OP060351 and OP060352, respectively. Phylogenetic analysis of combined ITS, TUB2, and TEF genes using the Mrbayes inference method showed that the three isolates clustered with Nigrospora musae as a distinct clade (Fig. 2). Combine with morphological characteristics and phylogenetic analysis, three isolates were identified as N. musae. 30 2-year-old healthy potted plants of T. chinensis were used for pathogenicity test. 25 of these plants were inoculated by injecting 10 µL of the conidia suspension (1 × 106 conidia/mL) into stems and then wrap around the seal to moisturize. The remaining 5 plants were injected with the same amount of sterilized distilled water as a control. Finally, all potted plants were placed in a greenhouse at 25°C and 80% relative humidity. After 2 weeks, the inoculated stems developed lesions similar to those observed in the field, whereas controls were asymptomatic. N. musae was re-isolated from the infected stem and identified by both morphological characteristics and DNA sequence analysis. The experiments repeated three times showed similar results. As far as we know, this is the first report of N. musae causing T. chinensis stem blight in the world. The identification of N. musae could provide a certain theoretical basis for field management and further research of T. chinensis.

11.
Int J Mol Sci ; 24(6)2023 Mar 09.
Artículo en Inglés | MEDLINE | ID: mdl-36982311

RESUMEN

The formation of mature mRNA requires cutting introns and splicing exons. The occurrence of splicing involves the participation of the spliceosome. Common spliceosomes mainly include five snRNPs: U1, U2, U4/U6, and U5. SF3a2, an essential component of spliceosome U2 snRNP, participates in splicing a series of genes. There is no definition of SF3a2 in plants. The paper elaborated on SF3a2s from a series of plants through protein sequence similarity. We constructed the evolutionary relationship of SF3a2s in plants. Moreover, we analyzed the similarities and differences in gene structure, protein structure, the cis-element of the promoter, and expression pattern; we predicted their interacting proteins and constructed their collinearity. We have preliminarily analyzed SF3a2s in plants and clarified the evolutionary relationship between different species; these studies can better serve for in-depth research on the members of the spliceosome in plants.


Asunto(s)
Ribonucleoproteínas Nucleares Pequeñas , Empalmosomas , Empalmosomas/metabolismo , Filogenia , Ribonucleoproteínas Nucleares Pequeñas/genética , Empalme del ARN/genética , Ribonucleoproteína Nuclear Pequeña U2/química , Ribonucleoproteína Nuclear Pequeña U2/genética , Ribonucleoproteína Nuclear Pequeña U2/metabolismo , ARN Mensajero/metabolismo
12.
Int J Mol Sci ; 24(6)2023 Mar 10.
Artículo en Inglés | MEDLINE | ID: mdl-36982373

RESUMEN

Alternative splicing (AS) is one of the major post-transcriptional regulation mechanisms that contributes to plant responses to various environmental perturbations. Darkness and heat are two common abiotic factors affecting plant growth, yet the involvement and regulation of AS in the plant responses to these signals remain insufficiently examined. In this study, we subjected Arabidopsis seedlings to 6 h of darkness or heat stress and analyzed their transcriptome through short-read RNA sequencing. We revealed that both treatments altered the transcription and AS of a subset of genes yet with different mechanisms. Dark-regulated AS events were found enriched in photosynthesis and light signaling pathways, while heat-regulated AS events were enriched in responses to abiotic stresses but not in heat-responsive genes, which responded primarily through transcriptional regulation. The AS of splicing-related genes (SRGs) was susceptible to both treatments; while dark treatment mostly regulated the AS of these genes, heat had a strong effect on both their transcription and AS. PCR analysis showed that the AS of the Serine/Arginine-rich family gene SR30 was reversely regulated by dark and heat, and heat induced the upregulation of multiple minor SR30 isoforms with intron retention. Our results suggest that AS participates in plant responses to these two abiotic signals and reveal the regulation of splicing regulators during these processes.


Asunto(s)
Proteínas de Arabidopsis , Arabidopsis , Empalme Alternativo , Arabidopsis/metabolismo , Empalme del ARN , Proteínas de Arabidopsis/genética , Proteínas de Arabidopsis/metabolismo , Isoformas de Proteínas/metabolismo , Estrés Fisiológico/genética , Regulación de la Expresión Génica de las Plantas
13.
Int J Mol Sci ; 24(19)2023 Sep 29.
Artículo en Inglés | MEDLINE | ID: mdl-37834210

RESUMEN

Chinese pepper rust is a live parasitic fungal disease caused by Coleosporium zanthoxyli, which seriously affects the cultivation and industrial development of Z. armatum. Cultivating and planting resistant cultivars is considered the most economical and environmentally friendly strategy to control this disease. Therefore, the mining of excellent genes for rust resistance and the analysis of the mechanism of rust resistance are the key strategies to achieve the targeted breeding of rust resistance. However, there is no relevant report on pepper rust resistance at present. The aim of the present study was to further explore the resistance mechanism of pepper by screening the rust-resistant germplasm resources in the early stage. Combined with the analysis of plant pathology, transcriptomics, and metabolomics, we found that compared with susceptible cultivar TJ, resistant cultivar YK had 2752 differentially expressed genes (DEGs, 1253 up-, and 1499 downregulated) and 321 differentially accumulated metabolites (DAMs, 133 up- and 188 down-accumulated) after pathogen infection. And the genes and metabolites related to phenylpropanoid metabolism were highly enriched in resistant varieties, which indicated that phenylpropanoid metabolism might mediate the resistance of Z. armatum. This finding was further confirmed by a real-time quantitative polymerase chain reaction analysis, which revealed that the expression levels of core genes involved in phenylpropane metabolism in disease-resistant varieties were high. In addition, the difference in flavonoid and MeJA contents in the leaves between resistant and susceptible varieties further supported the conclusion that the flavonoid pathway and methyl jasmonate may be involved in the formation of Chinese pepper resistance. Our research results not only help to better understand the resistance mechanism of Z. armatum rust but also contribute to the breeding and utilization of resistant varieties.


Asunto(s)
Transcriptoma , Zanthoxylum , Zanthoxylum/genética , Zanthoxylum/metabolismo , Fitomejoramiento , Metaboloma , Flavonoides/metabolismo , Resistencia a la Enfermedad/genética , Enfermedades de las Plantas/genética , Enfermedades de las Plantas/microbiología
14.
Int J Mol Sci ; 24(5)2023 Mar 02.
Artículo en Inglés | MEDLINE | ID: mdl-36902308

RESUMEN

Phomopsis capsici (P. capsici) causes branch blight of walnuts, which leads to significant economic loss. The molecular mechanism behind the response of walnuts remains unknown. Paraffin sectioning and transcriptome and metabolome analyses were performed to explore the changes in tissue structure, gene expression, and metabolic processes in walnut after infection with P. capsici. We found that P. capsici caused serious damage to xylem vessels during the infestation of walnut branches, destroying the structure and function of the vessels and creating obstacles to the transport of nutrients and water to the branches. The transcriptome results showed that differentially expressed genes (DEGs) were mainly annotated in carbon metabolism and ribosomes. Further metabolome analyses verified the specific induction of carbohydrate and amino acid biosynthesis by P. capsici. Finally, association analysis was performed for DEGs and differentially expressed metabolites (DEMs), which focused on the synthesis and metabolic pathways of amino acids, carbon metabolism, and secondary metabolites and cofactors. Three significant metabolites were identified: succinic semialdehyde acid, fumaric acid, and phosphoenolpyruvic acid. In conclusion, this study provides data reference on the pathogenesis of walnut branch blight and direction for breeding walnut to enhance its disease resistance.


Asunto(s)
Juglans , Juglans/genética , Transcriptoma , Fitomejoramiento , Metaboloma
15.
Plant Dis ; 2022 Aug 31.
Artículo en Inglés | MEDLINE | ID: mdl-36044368

RESUMEN

Phyllostachys aureosulcata McClure 'Spectabilis' C.D. Chu. et C.S. Chao is predominantly native to subtropical to warm temperate areas and is widely cultivated for landscaping in China (Neményi et al. 2015). In November 2020 (10 - 16 ℃), culm blight symptoms were observed on P. aureosulcata 'Spectabilis' in Wangjiang Tower Park (all kinds of plant areas are about 9.8 ha), Chengdu City (104°09'30.42″ E, 30°63'18.89″ N). Fifty plants were surveyed, and disease incidence was recorded as approximately 30%. Initially, chlorotic necrotic patches appeared on the culms, and gradually the patches became white, expanded to both ends, and encircled the whole culm with black edge and conidiomata, which eventually led to wilt and death. Five samples from different bamboos were collected and one of them were used for morphological observation. Five single conidia isolates were carried out on potato dextrose agar (PDA) at 25±1℃ (Chomnunti et al. 2014). Colonies were initially white and then yellowish in the center with abundant aerial mycelia. On the culm, conidiomata were dry, black, and filamentous. Conidiophores were reduced to conidiogenous cells. Conidiogenous cells were smooth, hyaline, ampulliform to doliiform. Conidia were ellipsoid to globose, dark brown, smooth and aseptate, measuring 5.2 to 9.4 × 4.4 to 7.3 µm, (=8.2 × 6.5µm, n=50). On the PDA medium, conidia were globose to subglobose, olive green to pale brown, and smooth, larger than those from the host in size, measuring 9.0 to 18 × 7.5 to 9.5 µm ( =36.6 × 18.8 µm, n=50). These asexual structures were extremely similar to Apiospora locuta-pollinis (F. Liu & L. Cai) X.G. Tian & Tibpromma (Zhao et al. 2018). DNA was extracted from the representative strain (SICAUCC 22-0036), and the internal transcribed spacer (ITS), translation elongation factor 1-alpha (tef1-α), beta-tubulin (tub2), 28S large subunit rDNA (LSU) were amplified and sequenced with primers ITS1/ITS4 (White et al. 1990), EF1-728F (Carbone & Kohn 1999)/EF2 (O'Donnell et al. 1998), T1 (O'Donnell & Cigelnik 1997)/Bt2b (Glass & Donaldson 1995) and LR0R/LR5 (Rehner & Samuels 1994). The newly generated sequences were deposited in GenBank with accession nos. ON228609 (ITS), ON324018 (tef1-α), ON237657 (tub2), and ON228665 (LSU). Nucleotide blast showed 98.97%, 100% and 99.46% identities with A. locuta-pollinis (LC11683, ex-holotype) (accession nos. MF939595, MF939622, MF939616), and LSU data missing. Phylogenetic analyses using maximum likelihood showed a 92% bootstrap support value in a clade with A. locuta-pollinis (Fig 2). Eight healthy plants (2-year-old) were used for the pathogenicity test. Culms of four healthy bamboos were wounded via sterile double-edged blade and sprayed with conidial suspension (105 conidia/ml) prepared from 4-week-old cultures that were incubated on PDA at 25℃. The other four bamboos were sprayed with sterile distilled water as controls. Inoculated plants were placed in a growth chamber (25℃, 90% relative humidity, 12-h photoperiod). About 60 days later, necrotic patches similar to those observed in the field were found on the inoculated culms, and no symptoms were observed on the controls. The pathogen was reisolated from the diseased culms with identical morphology as previously described. To our knowledge, this is the first report of culm blight on P. aureosulcata 'Spectabilis' caused by A. locuta-pollinis. The risk of this pathogen needs further evaluation, and effective control measures should be taken.

16.
Plant Dis ; 2022 Feb 11.
Artículo en Inglés | MEDLINE | ID: mdl-35147449

RESUMEN

Bambusa pervariabilis × Dendrocalamopsis grandis is the main cultivated bamboo species used for ecological construction in the Yangtze River basin. This species has the advantages of easy reproduction, wide adaptability and strong resistance and has high economic, ecological and social benefits (Peng et al. 2020). One area of B. pervariabilis × D. grandis with basal rot disease was discovered in Renshou County, Sichuan Province, China (29°41'N, 104°11'E) in June 2020. The disease occurrence area was 68 hm2 in Renshou County, with an incidence rate of 34.8%, and 5% of the B. pervariabilis × D. grandis with basal rot disease died. The pathogen initially invaded from the first section of the base of the bamboo stalk, appearing as black to yellowish brown strips or lumps of disease spots, and rapidly developed horizontally and vertically, which caused the whole plant to wither in severe cases. Diseased tissues were collected from the base of a 4-year-old bamboo stalk with a sterile blade. 100 pieces (5 × 5 × 2 mm) of diseased tissues were sterilized with 3% NaClO for 30 s and in 75% ethanol for 90 s, rinsed three times with sterile distilled water, dried with sterile surface water on sterile filter paper, plated onto potato dextrose agar amended with streptomycin sulfate (Solarbio, 50 µg/ml), and incubated at 25 °C for 7 days with light. A total of five isolates were obtained, of which four isolates were similar in morphology. Using the method of monospore isolation (Leslie and Summerell 2006) and culturing it on PDA, the fungus produced round colonies with a diameter of approximately 8.4 mm and a surface color ranging from white to purple within 7 days at 25 °C. For identification by typical spores, the fungus was cultured on carnation leaf agar (CLA) medium at 25 °C for 7 days. The microconidia by the isolates BD2002, BD2004, BD2008 and BD2010 cultured on CLA medium were elliptical, ovoid, without septum, and measured 4.56 to 15.53 µm long × 1.36 to 6.98 µm wide (n=100). The macroconidia were rod-shaped or slightly curved, tapering apically with three to five septa, and measured 18.86 to 52.99 × 1.56 to 6.42 µm in size (n=100). According to the morphological characteristics of macroconidia and microconidia, the isolates were identified as Fusarium sp. (Leslie and Summerell 2006). For molecular identification, fungal DNA of isolates BD2002, BD2004, BD2008 and BD2010 was extracted by a fungal genomic DNA extraction kit. Polymerase chain reactions (PCRs) were performed with primers ITS1/ITS4 for the internal transcribed spacer (ITS) rDNA region (White et al. 1990), primers Bt2a/Bt2b for the ß-tubulin (TUB) region (Glass and Donaldson 1995), primers EF1F/EF2R for the translation elongation factor 1α (TEF) region (Carbone et al. 1999), primers 5f2/7cr for the RNA polymerase II genes (RPB2) region (O'Donnell et al. 2010), primers H3-1a/H3-1b for the histone H3 (HIS) region (Jacobs et al. 2010), and primers NMS1/NMS2 for the mitochondrial small subunit (mtSSU) rDNA region (Stenglein et al. 2010). Using BLASTn to search GenBank for ITS, TUB, TEF, RPB2, HIS and mtSSU sequences, all isolates showed the highest similarity with Fusarium proliferatum (Matsushima) Nirenberg. The representative isolate BD2010 showed that ITS had 99.61% similarity to F. proliferatum Z23-28 (FJ648201.1); HIS had 99.57% similarity to F. proliferatum M06A_4G_4 (KX681532.1); and the TUB, TEF, RPB2, and mtSSU sequences showed 99.67%, 99.10%, 99.06%, and 99.57% similarity, respectively, to F. proliferatum ITEM2287 (accession numbers LT841243.1, LT841245.1, LT841252.1, and LT841247.1 in GenBank). The GenBank numbers of the representative isolate BD2010 were ITS, OK325614; TUB, OK377026; TEF, OK377027; RPB2, OK377028; HIS, OK377029; and mtSSU, OK338638. To confirm the pathogenicity, thirty 4-year-old healthy bamboo plants were grown in 30 pots. Each five plants were inoculated with one isolate, and a total of twenty-five plants were inoculated with five isolates. A conidia suspension (1 × 106 conidia/ml) of the fungus was inoculated (100 µl each) into plants that had been acupunctured at the base by a sterile syringe. Five control plants were inoculated only with the same amount of sterile distilled water. The inoculation site was wrapped with wet gauze to maintain moisture. All bamboo plants were watered every seven days. The illumination conditions were 12 h light and 12 h dark. All plants were cultured in a greenhouse at 25-28 °C and 70-80% relative humidity. One month later, twenty plants inoculated with conidial suspensions of BD2002, BD2004, BD2008 and BD2010 showed the same symptoms as those observed in the field, whereas plants inoculated with the other fungus and the control treatment remained asymptomatic. The pathogenicity test was conducted three times, and the experimental results were consistent. Furthermore, the fungi were reisolated from the diseased part and were identified as F. proliferatum by morphological and molecular comparison. To our knowledge, this is the first report of basal rot disease caused by F. proliferatum on B. pervariabilis × D. grandis in China. This research is conducive to laying the foundation for the development of effective control strategies for basal rot disease in this species.

17.
Plant Dis ; 2022 Mar 29.
Artículo en Inglés | MEDLINE | ID: mdl-35350888

RESUMEN

Juglans regia L. is one of the major cultivated walnut species in China for nuts and wood (Pollegioni et al. 2012). In June 2020, branches with blight symptoms were observed in an orchard at Chongzhou City (30°33'34″N, 103°38'35″E). In an orchard of 30 hectares, disease incidence was around 50%. A total of 15 plants were sampled and 40% of their branches were affected by this disease. Firstly, brown and irregular spots appeared, then the spots gradually expanded and encircled the branch, which eventually killed the branch. Five samples of diseased branches from different trees were collected and a single fungal isolate was obtained from each of the five samples using the single ascospore isolation (Chomnunti et al. 2014). Colonies of the five isolates on potato dextrose agar (PDA) were identical that initially appeared white on the top, becoming light to dark brown with age. On the host, ascostroma were black, globose to subglobose, short-papillate, ostiolate, 260 - 410 × 210 - 320 µm (x = 335 × 265 µm, n = 20). Asci were 8-spored, bitunicate, cylindrical, short pedicellate, 55 - 78 × 8 - 12 µm (x = 67.5 × 10 µm, n = 40). Ascospores were 1-septate, fusiform to ellipsoidal, slightly curved, guttulate, 12 - 17 × 3 - 5 µm (x = 14.5 × 4 µm, n = 40). These sexual morphological characteristics are consistent with the Palmiascoma qujingense Phook. & K.D. Hyde (Monkai et al. 2021). Asexual morphs were formed on PDA in incubator after 17 days (25℃, 90% relative humidity, 12-h photoperiod). Conidiomata were black, globose to subglobose, 220 - 300 × 240 - 380 µm (x = 270 × 310 µm, n = 20). Conidia were oblong to ellipsoidal, aseptate and smooth-walled, 3 - 7 × 2 - 4 µm (x = 4.9 × 3 µm, n = 50). The genomic DNA of a representative isolate SICAUCC 21-0013 was extracted, and the internal transcribed spacers (ITS) region, large subunit rDNA (LSU) region, small subunit rDNA (SSU) region, and the largest subunit of RNA polymerase II (rpb2) gene were amplified and sequenced with primers ITS5/ITS4 (White et al. 1990), LR0R/LR5 (Rehner et al. 1994), NS1/NS4 (White et al. 1990), and fRPB2-5F/fRPB2-7cR (Liu et al. 1999), respectively. The sequences were deposited in NCBI with accession numbers MZ983549, MZ959419, MZ951112, and MZ818772, respectively, which showed 100%, 100%, 99.14%, and 99.59% identities with P. qujingense KUMCC 19-0201 (holotype) (accession numbers MT477185, MT477186, MT477183, MT495782respectively). Phylogenetic analysis (maximum likelihood) based on a concatenated dataset showed 93% bootstrap support values with P. qujingense. To verify Koch's postulates, 9 healthy branches from three 1-year-old seedlings were inoculated with conidial suspension (106 conidia/ml) from 4-week-old cultures via pin-prick inoculation (Desai et al. 2019), and the same number of seedlings and branches were inoculated with sterile water as controls. Plants were placed in a greenhouse at 25℃ and 90% RH on a 12-h fluorescent light/dark regime. After 28 days, brown spots were formed on P. qujingense-inoculated branches and similar to those observed in the field, while the controls remained asymptomatic. The pathogen was re-isolated from the lesions and identified by morphology and phylogeny. To our knowledge, this is the first report of P. qujingense causing branch blight on J. regia in the world. This disease potentially impacts the growth and yield of J. regia, and control measures should be made.

18.
Plant Dis ; 2022 Aug 30.
Artículo en Inglés | MEDLINE | ID: mdl-36040226

RESUMEN

Iron walnut (Juglans sigillata Dode) is a temperate deciduous tree indigenous to China. It is mainly distributed in southwestern China, and valued for its wood and nuts (Feng et al. 2018). In September 2020, symptoms of canker on J. sigillata were observed in an orchard measuring 2 hectares located in Chongzhou City, Sichuan Province (31°5' 25″N, 105°27'36″E, 365 m altitude). Twenty percent of plants showed canker symptoms during the 50 surveyed plants. The infected trunk showed necrotic lesions with black pycnidia, that led to necrosis of branches and death of the whole plant in severe cases (Fig. 1). Six specimens from different diseased plants were collected for pathogen isolation and morphological observation. Pure cultures were obtained from single conidium on potato-dextrose agar (PDA) media according to the method described by Chomnunti (Chomnunti et al. 2014). Colonies grew fast and reached 3 cm after 5 days. The aerial mycelium was abundant, which was initially white and then grayish. Conidiomata on the host were measured 160-280 µm × 140-190 µm (average: 220 × 165 µm, n = 20), stromatic, uniloculate, dark brown to black, immersed, and erumpent when mature. Pycnidial walls 32-58 µm wide, were composed of 5-7 layers of brown to dark brown cells. Conidia were hyaline, and ellipsoidal with rounded apex and base, widest at the middle, thick-walled, and unicellular, with a size 21.5-31 µm × 11.5-15.7 µm (average: 27 × 13.5 µm, n = 50). Morphological characteristics fit the description of Lasiodiplodia pseudotheobromae A.J.L. Phillips, A. Alves & Crous (Aives et al. 2008). The internal transcribed spacers (ITS), 18S small subunit rRNA (SSU), 28S large subunit rDNA (LSU), translation elongation factor 1-alpha (tef1-α), and beta-tubulin (tub2) were amplified by polymerase chain reaction and sequenced with primers ITS1/ITS4, NS1/NS4, LR0R/LR5, EF1-728F/EF1-986R and Bt2a/Bt2b, respectively (Li et al. 2018). The sequences of the representative isolate (SICAUCC 22-0079) were deposited in NCBI with accession numbers ON090365 (ITS), ON090406 (SSU), ON090418 (LSU), ON112377 (tef1-α), and ON112378 (tub2), respectively. Nucleotide blast showed 100% similarity of all the analyzed and NCBI submitted isolates with L. pseudotheobromae (CBS116459; holotype) (accession numbers EF622077, EU673199, EU673256, EF622057, EU673111). Phylogenetic analyses based on a combined dataset showed 100% bootstrap support values in a clade with L. pseudotheobromae complexes (Fig. 2). Based on morphological and molecular analyses, the fungal pathogen was identified as L. pseudotheobromae. To conduct Koch's postulates, four 2-year-old healthy plants of J. sigillata were inoculated with 10 µL spore suspension (105 conidia/mL) onto the wounded sites via sterile pin. As control, four healthy plants were treated with sterile distilled water. The inoculated and untreated plants were placed in a growth chamber at 25°C with relative humidity >90% and 12-h photoperiod. Trunk canker symptoms appeared on inoculated plants after 15-20 days, and the pathogen was re-isolated and the controls were symptomless, confirming Koch's postulates. L. pseudotheobromae is widely distributed in various plants all over the world, usually as a pathogen associated with damping-off, wilt, die-back, root rot, collar rot, witches' brooms, or fruit rots (Zhao et al. 2010). To our knowledge, this is the first report of trunk canker on J. sigillata caused by L. pseudotheobromae in China. Trunk canker caused by L. pseudotheobromae is becoming a potential threat to walnut production, and some necessary measures for integrated management should be made.

19.
Plant Dis ; 2022 Jun 13.
Artículo en Inglés | MEDLINE | ID: mdl-35698254

RESUMEN

Juglans sigillata Dode, an endemic walnut species native in southwest China, is mainly used as nuts in Sichuan Province (Jin et al., 2019). In May 2021, symptoms of branches blight were observed in an orchard measuring 10 hectares located in Mianyang City, Sichuan Province (31°5' 25″N, 105°27'36″E, 365 m above sea level). About 40% of plants were diseased in the quadrat consisting of twenty walnut trees, and 20% of branches were dead on each affected tree. Initially, light brown spots appeared; then, the spots expanded to surround the whole branches; finally, the branches changed from brown to reddish-brown and died. Four symptomatic branches were sampled randomly from different trees. Next, four fungal isolates were obtained from the acervuli of each branch using the single-conidium isolation (Chomnunti et al. 2014) and cultured on potato dextrose agar (PDA). The Petri dishes were placed in an incubator and cultured at 25 °C under a 12-h photoperiod. Colonies were initially white with thin aerial mycelia and gradually turned dark grey with irregular margins. Conidiomata were acervular, black and scattered, with a diameter of 0.3 - 0.7 mm. Conidiophores were narrowly cylindrical, simple or branched at the base, 30 - 43 × 3 - 8 µm (x = 36.5 × 5.5 µm, n = 40). Conidiogenous cells were annellidic with distinct annellations. Conidia were unicellular, brown when mature, narrowly ellipsoid with gelatinous sheaths and truncate scars at the base, 17 - 32 × 7 - 12 µm (x = 27 × 9 µm, n = 40). The genomic DNA of a representative isolate SICAUCC 22-0064 was extracted, and the internal transcribed spacer (ITS) region, guanine nucleotide-binding protein subunit beta gene (ms204), translation elongation factor 1-alpha (tef1-α), and partial sequences of ß-tubulin (tub2) were amplified by polymerase chain reaction and sequenced with primers V9G/LR5 (de Hoog & van den Ende 1998), MS-E1F1/MS-E5R1 (Walker et al. 2012), EF1-728F (Carbone & Kohn 1999)/TEF1LLErev (Jaklitsch et al. 2005), and T1/BtHV2r (Voglmayr et al. 2017), respectively. The sequences of ITS, ms204, tef1-α, and tub2 were deposited in NCBI with accession numbers ON000068, ON112376, ON112374, and ON112375, respectively. With the consideration of the sequence lack of ms204 and tub2 in the ex-type strain (D96) of Juglanconis appendiculata Voglmayr & Jaklitsch, the isolate D140 was used for nucleotide blast. The results showed 99.68%, 100%, 100%, and 100% identities of ITS, ms204, tef1-α, and tub2 with D140 (accession numbers KY427138, KY427157, KY427207, KY427226). Phylogenetic analysis based on a combined dataset showed 100% bootstrap with J. appendiculata, and the morphology was consistent with the asexual stage of J. appendiculata (Voglmayr et al., 2017). To verify Koch's postulates, five branches wounded by pin-prick were sprayed with conidial suspension (1 × 105 conidia/mL) in each plant, and three repetitions were performed on healthy 2-year-old potted plants. The same number of branches were sprayed with sterile distilled water as controls. The plants were placed in a greenhouse at 25 ℃ under 90% relative humidity and a 12-h fluorescent light/dark regime. After five weeks, all the inoculated branches showed brown necrosis similar to that observed in the field, and no symptoms occurred on the controls. The pathogens were re-isolated from the necrotic lesions and identified by morphology and phylogeny. J. appendiculata has been reported on Juglans nigra and J. regia in Austria, France, Spain and Greece (Farr & Rossman 2022). This paper is the first report of branch blight on Juglans sigillata caused by J. appendiculata in China. This result may develop the understanding of walnut diseases and lay a foundation for further management.

20.
Plant Dis ; 2022 Sep 11.
Artículo en Inglés | MEDLINE | ID: mdl-36089675

RESUMEN

Alocasia macrorrhizos (Giant elephant's ear), a perennial herb in the Araceae family, is native to South Asia and the Asia-Pacific (Takano, et al. 2012). It is cultivated as a medicinal and ornamental plant, and has a considerable economic importance in China. In September 2020, a severe infection of unknown leaf spot disease was observed on these plants at the Sichuan Agricultural University, Sichuan, China. The leaf spots first appeared as yellow dots. As these lesions expanded, they became circular to oval and light brown with darker brown edges. Around the lesions, the leaf tissue was chlorotic, thereby creating a yellow halo. When the infection became severe, spots merged into larger irregular lesions. Eventually, the diseased leaves senesced and dried. To identify the pathogen, five leaf samples of diseased plants were collected, and symptomatic tissues were surface-disinfected with 75% ethanol for 30 s followed by 3% NaCl solution for 30 s. Samples were rinsed three times in sterilized water, placed on potato dextrose agar (PDA), and incubated at 25°C ± 1°C in the dark. The colony grown on PDA was white (3 days), the center was brown (5 days), turned pink to dark red (8 days) with fluffy aerial mycelium and pigmentation with age. Ten pure cultures were inoculated into carnation leaf agar (CLA) medium and incubated at 25°C in an incubator (12 h for one light-dark cycle). In CLA medium, pathogen produced hyaline, sickle-shaped, macroconidia with 3 to 5 septa, and an average size of 30 to 50 × 4 to 5 µm (n = 30) macroconidia but no microconidia in 10 days. Chlamydospores were spherical to subspherical (5.4 to 13.8 µm). Morphological characteristics of the all isolates were consistent with the description of the Fusarium asiaticum (Leslie and Summerell 2006). To validate this identification, RNA polymerase II (RPB2) (Liu et al. 1999), translation elongation factor (EF-1) (Geiser et al. 2004), and ß-tubulin (TUB2) gene region of five isolates were amplified and sequenced (O' Donnell et al. 2015; White et al. 1990). The sequence of one representative isolate (ZL10) sequence was submitted to GenBank (ON215729, ON215730, and ON215731). The NCBI BLAST identified the top hits, 100%, 100%, and 99.87% for RPB2, EF, and TUB gene sequences, respectively, all indicating to Fusarium asiaticum. Pairwise matched of RPB2 and EF genes by MycoBank Fusarium MSIL showed the top hit rate of 100% for F. asiaticum (MH582120 and MH582249). For Koch's postulate and pathogenicity test, spore suspensions (1 × 10^7 conidia/ml) collected from PDA and CLA cultures with 0.05% Tween 80 buffer were used to inoculate with a spray bottle on leaves of a one year old A. macrorrhizos plants. Two leaves of each plant (20 pots in total) were inoculated with the spore suspension (approximately 2000 µl per leaf). An equal number of control leaves were applied with water and 0.05% Tween 80 buffer. Twenty days later, the inoculated plants showed similar symptoms to those of the original diseased plants while the controls remained asymptomatic. Fusarium asiaticum was reisolated from the infected leaves and confirmed using morphological characteristics and DNA sequence analysis. The pathogenicity test was repeated three times with similar results. This first report raises awareness of a new leaf spot disease infecting a commercial A. macrorrhizos in China. It provides an insight for a need of systematic survey identifying current spread, disease origin, and ultimately developing disease management strategies. Funding: Funding was provided by Sichuan Agricultural University Subject Dual Support Program (Grant No. 2121993055). Funding was provided by Deyang Science and Technology Bureau (Sichuan Province) for key R&D projects in agriculture and rural areas (Grant No. 2021NZ048). Funding was provided by the Sichuan Provincial Department of science and technology for the Sichuan Provincial Science and technology project for connecting and Promoting Rural Revitalization (Grant No, 2022ZHXC0007) References: Geiser, D. M., et al. 2004. Eur. J. Plant Pathol. 110:473. https://doi.org/10.1023/B:EJPP.0000032386.75915.a0 Crossref, ISI, Google Scholar Leslie, J. F., and Summerall, B. A., eds. 2006. Page 176 in The Fusarium Laboratory Manual. Blackwell Publishing, Ames, IA. https://doi.org/10.1002/9780470278376 Liu, Y. J., et al. 1999. Mol. Biol. Evol. 16:1799. https://doi.org/10.1093/oxfordjournals.molbev.a026092 O'Donnell, K., and Cigelnik, E. 1997. Mol. Phylogenet. Evol. 7:103. https://doi.org/10.1006/mpev.1996.0376 Takano K T, et al. 2012, Plant Bio., 14(4). https://doi.org/10.1111/j.1438-8677.2011.00541.x.

SELECCIÓN DE REFERENCIAS
DETALLE DE LA BÚSQUEDA