Your browser doesn't support javascript.
loading
Mostrar: 20 | 50 | 100
Resultados 1 - 20 de 59
Filtrar
2.
Funct Integr Genomics ; 24(1): 25, 2024 Feb 07.
Artigo em Inglês | MEDLINE | ID: mdl-38324167

RESUMO

Chemotherapy resistance is the main reason for the poor prognosis of ovarian cancer (OC). FHL1 is an important tumour regulator, but its relationship with the prognosis, drug resistance, and tumour microenvironment of OC is unknown. Immunohistochemistry was used to determine FHL1 expression in OC. Kaplan‒Meier plotter was used for survival analysis. The value of gene expression in predicting drug resistance was estimated using the area under the curve (AUC). Bivariate correlation was used to determine the coexpression of two genes. Functional cluster and pathway enrichment were used to uncover hidden signalling pathways. The relationship between gene levels and the tumour microenvironment was visualised through the ggstatsplot and pheatmap packages. The mRNA and protein levels of FHL1 were downregulated in 426 and 100 OC tissues, respectively. Low FHL1 expression was correlated with good progression-free survival (PFS), postprogression survival, and overall survival (OS) in 1815 OC patients, and was further confirmed to be associated with good OS by immunohistochemistry in 152 OC tissues. Furthermore, FHL1 was downregulated in drug-sensitive tissues, while its high expression predicted drug resistance (AUC > 0.65). Mechanistically, FHL1 was coexpressed with FLNC, CAV1, PPP1R12B, and FLNA at the mRNA and protein levels in 558 and 174 OC tissues, respectively, and their expression was downregulated in OC. Additionally, very strong coexpression of FHL1 with the four genes was identified in at least 23 different tumours. Low expression of the four genes was associated with good PFS, and the combination of FHL1 with the four genes provided better prognostic power. Meanwhile, the expression of all five genes was strongly and positively associated with the abundance of macrophages. Low FHL1 expression acts as a favourable factor in OC, probably via positive coexpression with FLNC, CAV1, PPP1R12B, and FLNA.


Assuntos
Neoplasias Ovarianas , Humanos , Feminino , Macrófagos , RNA Mensageiro , Resistência a Medicamentos , Microambiente Tumoral , Proteínas Musculares , Peptídeos e Proteínas de Sinalização Intracelular , Proteínas com Domínio LIM
3.
Plant Dis ; 2024 Feb 22.
Artigo em Inglês | MEDLINE | ID: mdl-38386296

RESUMO

This study identified a new species (Cercospora Polygonatum) that causes gray leaf spot (GLS) disease in cultivated Polygonatum cyrtonema. This fungal species was isolated from the affected region of GLS on P. cyrtonema leaves. Pathogenicity bioassays were conducted based on Koch's postulates. Morphology was examined based on the features of conidiomata, conidiogenous loci, conidia/conidiophores, and conidiogenous cells. The rDNA internal transcribed spacer region, calmodulin, translation elongation factor 1-alpha, and histone genes were subjected to phylogenetic analysis using MrBayes tool via in Phylosuite. Bootstrap support analysis for phylogenetic placement confirmed the new species, which was significantly different from the closely related species C. senecionis-walkeri and C. zeae-maydis. The morphological characteristics also supported this finding, with the conidiogenous of C. polygonatum being considerably shorter than those of C. senecionis-walkeri or C. zeae-maydis. In addition, C. polygonatum was distinguished by its cultural characteristics. As this fungus was isolated from P. cyrtonema, it was named C. polygonatum F.Q. Yin, M. Liu&W. L. Ma, sp. nov. The type specimen (H8-2) was preserved at the China General Microbiological Culture Collection Center. This is the first report of GLS caused by C. polygonatum on P. cyrtonema leaves in China. The current study enriches the knowledge regarding Cercospora sp., contributes to the identification of a species causing GLS in P. cyrtonema, and provides useful information for the effective management of this disease.

4.
Plant Dis ; 2024 Jan 10.
Artigo em Inglês | MEDLINE | ID: mdl-38197883

RESUMO

Belamcanda chinensis (L.) Redouté, a member of the Iridaceae family, is globally well-known for its medicinal value as clearing away heat, detoxifying, detumescence and pain (Qin 2000). In 2021, spots were observed on 40% B. chinensis leaves and about 28 disease index in Wanzhou District (30°32'N; 108°22'E) of Chongqing. Initial symptoms appeared as circular yellow white, sunken spots lesions, and then expanded into irregular lesions, the center of the spots was beige, external layer was light brown and surrounded by yellow halo. Symptomatic leaf tissues (5 × 5 mm) were cut from the infected margin, surface sterilized with 75% ethanol for 1 min, washed with 3% sodium hypochlorite for 3 min, rinsed three times with sterile water, placed on potato dextrose agar (PDA) medium incubated at 25°C for 7 days in the dark, forty isolates with similar morphology were obtained. Three isolates (SG9、SG20 and SG33) was selected for subsequent research. Colonies color changed from beige to light brown color after 14 days on PDA medium. Fungal colonies transformed from beige to brown at the edges after 28 days and light brown on top. Ascomata dark brown, ellipsoidal to globose 116.6 to 253.3 × 89.6 to 172.6 µm in diamensions. Asci stipitate, cylindrical with obtuse ends, and 69.1 to 114.7 × 10.2 to 24.1 µm (n = 30) in size, with eight overlapping linearly biseriate ascospores. Ascospores brown, narrowly fusiform, straight or slightly curved with three transversely septate, slightly constricted at septa, and 9.7 to 12.6 × 27.6 to 32.6 µm (n = 30). These characteristics are consistent with Phaeosphaeria sp. reported by Quaedvlieg et al in 2013. DNA was extracted from representative isolates. The internal transcribed spacer (ITS) region, the large subunit rDNA (LSU), the small subunit rDNA (SSU) and the RNA polymerase II second largest subunit (RPB2) genes were amplified for Polymerase chain reaction (PCR) by used ITS1/ITS4, LR5/LROR, NS1/NS4, and RPB2-5f2/RPB2-7cr primers (White et al. 1990; Vilgalys et al. 1990; Qi M W. et al. 2008; De G. J. et al. 1992). The sequences were submitted to NCBI GenBank: SG-G9 (ITS, OR701701; LSU, OR701699; SSU, OR701700; RPB2, OR738464); SG-G20 (ITS, OQ748032; LSU, OQ780728; SSU, OQ780723; RPB2, OQ779979); SG-G33 (ITS, OQ748033; LSU, OQ780729; SSU, OQ780722; RPB2, OQ779980). A phylogenetic analysis revealed a 99% similarity to the Phaeosphaeria caricicola CBS 603.86 (ITS, KF251182; LSU, GQ387590; SSU, GQ387529; RPB2, KF252189) sequences. Mycelial agar plugs (5-mm diameter) from a 7-day-old PDA culture of a fungal isolate were placed onto pinpricked leaves of three two-year-old B. chinensis plants. While the sterile PDA plugs inoculated in pinpricked leaves of B. chinensis as controls. Inoculated plants were placed in a greenhouse at 25°C and remained 95±1% relative humidity. The inoculated leaves of treatment developed symptoms after 20 days, whereas no symptoms occurred on controls, fulfilling Koch's postulates. The experiments were repeated three times. The fungus was re-isolated and was identical to original isolate by morphologically and molecularly. As far as we know, P. caricicola can cause diseases on carex plants and has been found in Switzerland. This is the first report of P. caricicola causing leaf spot on B. chinensis in China. Along with recording the occurrence of this disease, plant disease management strategies need to be established to reduce losses.

5.
Int J Biol Macromol ; 258(Pt 1): 128855, 2024 Feb.
Artigo em Inglês | MEDLINE | ID: mdl-38114002

RESUMO

Conductive hydrogels have received widespread attention in the field of flexible sensors. However, a single network structure inside the hydrogel sensor usually makes it difficult to bear larger mechanical loadings, greatly limiting practical applications. Developing a recoverable conductive hydrogel sensor with high toughness and adaptability is still challenging. Herein, a high-performance polyvinyl alcohol (PVA)-based conductive composite hydrogel was constructed, assisted by green cellulose nanofibrils (CNFs), magnesium chloride (MgCl2), ethylene glycol (EG), and liquid metal (LM). The synergistic effects between CNFs and LM enhanced the network structure inside the recoverable hydrogel. This resulted in an excellent tensile strength of 3.86 MPa with an elongation at break of as high as 918.4 % and compressive strength of 4.04 MPa at 80 % strain. In addition, the conductive network composed of MgCl2 and LM endowed the hydrogel good electrical conductivity. Moreover, it could be used as a flexible strain sensor for various application scenarios, e.g., micro-stress monitoring (water droplet falling) and information encryption transmission of Morse code. Such uniqueness will provide a design strategy for developing a new generation of hydrogel sensors.


Assuntos
Celulose , Álcool de Polivinil , Condutividade Elétrica , Hidrogéis , Metais
6.
Biomark Med ; 17(18): 755-765, 2023 09.
Artigo em Inglês | MEDLINE | ID: mdl-38095985

RESUMO

Background: To explore the biological function and the underlying mechanisms of GOT2 in hepatocellular carcinoma (HCC). Materials & methods: The expression level and prognostic value of GOT2 were examined using International Cancer Genome Consortium and International Cancer Proteogenome Consortium databases. The cell counting kit-8 method, clone formation, Transwell® assays and western blotting were used to evaluate the effects of GOT2 on the biological function and autophagy of HCC cells. Results: The expression of GOT2 was downregulated in HCC tissues and correlated with poor prognosis of HCC patients. Knockdown of GOT2 promoted proliferation, migration and invasion of HCC cells and promoted cells' proliferation by inducing autophagy. Conclusion: GOT2 plays a tumor-inhibitory role in HCC and may be a potential therapeutic target for HCC.


Assuntos
Carcinoma Hepatocelular , Neoplasias Hepáticas , Humanos , Carcinoma Hepatocelular/diagnóstico , Carcinoma Hepatocelular/genética , Carcinoma Hepatocelular/metabolismo , Linhagem Celular Tumoral , Movimento Celular/genética , Proliferação de Células/genética , Regulação Neoplásica da Expressão Gênica , Neoplasias Hepáticas/diagnóstico , Neoplasias Hepáticas/genética , Neoplasias Hepáticas/metabolismo , Prognóstico
7.
Plant Dis ; 2023 Nov 15.
Artigo em Inglês | MEDLINE | ID: mdl-37966475

RESUMO

Elaeocarpus decipiens Hemsl., is a member of the Elaeocarpaceae family. It is a broad-leaved evergreen tree (Zhang et al. 2021) and has been widely used in landscape and gardens virescence. In March 2022, leaf spots were observed on E. decipiens leaves with 30-40% disease incidence and about 25 of disease index in Wanzhou District (30°32'N; 108°22'E) of Chongqing. Lesions showed light yellowish brown in color, black fruiting body in the center (Sporodochium), and surrounded by a purplish red halo at the interface between healthy and diseased tissues. The tissue interface of the lesions were cut into small pieces (5×5 mm), sterilized with 75% (vol. -/vol.) ethanol solution for 30 s, and 3% (vol. -/vol.) sodium hypochlorite solution for 3 min, and rinsed three times with sterile water. The sterile leaf tissues were placed on potato dextrose agar (PDA) medium in petri dishes and incubated for 5 days at 28°C in the dark, and produced thirty three uniform fungal colonies with in shape and color. The colonies had petal-shaped edges, with whitish or light pink hyphae, and black sporophores were observed at 14 days after inoculation. Sporodochium were ellipsoidal to globose with a size of 121.7 ~ 232.6 × 97.2 ~ 179.6 µm (n = 40). Conidiogenous cells were simple, tapering, hyaline, and smooth, 8 ~ 16 × 5.3 ~ 13.5 µm in size Its apex was surrounded by a gelatinous coating. Conidia were hyaline, slightly curved to naviculate, rounded to acute apex, smooth-walled, aseptate, and were 9 ~ 14.2 × 1.7 ~ 2.6 µm in size (n = 40). These morphological of the cultures are consistent with those of Coniella sp. reported by Alvarez et al. (2016). The genomic DNA of representative isolates DY4, DY24, and DY28 were extracted. The internal transcribed spacer (ITS) region, translation elongation factor 1-alpha (TEF1), and large subunit ribosomal RNA (LSU) were amplified with primers ITS1/ITS4 (White et al., 1990), EF728/EF986 (Rehner et al., 2005), and LR0R/LR5 (Vilgalys et al., 1990). The sequences were submitted to NCBI GenBank (https://www.ncbi.nlm.nih.gov/). BLASTn searches showed that the ITS (OQ926882-84), TEF1 (OR282454-56), and LSU (OQ926945-47) sequences had the highest similarity to Coniella quercicola with 99% (596/613, 597/613, and 593/613) identity for ITS (KX833595); 94% (315/536, 323/536, and 322/536) identity for TEF1 (KX833698); and 99% (933/898, 871/898, and 932/898) identity for LSU (KX833414), respectively. Phylogenetic analysis was performed using maximum likelihood method in MEGA 11.0 (Tamura et al., 2021), and the phylogenetic tree revealed a 100 % sequence similarity to the C. quercicola CBS 283.76 (ITS, KX833594; TEF1, KX833697; LSU, KX833413) and C. quercicola CBS 904.69. In the pathogenicity test, nine healthy plants of E. decipiens (five-year-old) were selected to use, 10 µL of spore suspension (106 conidia ml-1) were sprayed on the surface of four leaves per plant (six plants in total), and the other three plants were sprayed with sterile distilled water as controls. All plants were placed in a greenhouse with 95±1% relative humidity at 28°C for penetration of the cultures in an alternating dark (12 h) and light (12 h). At 5 days after inoculation, circular lesions symptoms were observed, whereas control plants remained asymptomatic. The fungus was reisolated from diseased leaf tissue and identified as Coniella quercicola according to the methods described as above. Previously, C. quercicola has been reported as a pathogen on Eucalyptus cloeziana in China (Zou et al., 2023), and Quercus robur in Netherlands (Alvarez et al., 2016). To our knowledge, this is the first report of C. quercicola causing leaf spot on E. decipiens in China. This study provides a basis for further elucidating the pathogenic mechanism, and the development of effective management for this disease.

8.
Plant Dis ; 2023 Oct 11.
Artigo em Inglês | MEDLINE | ID: mdl-37822100

RESUMO

Hosta plantaginea is an important horticultural plant with ornamental value and is widely cultivated in China. Since April 2022, leaf rot has been observed in the H. plantaginea plants in Wanzhou District, Chongqing City, China (31º14'58"N, 108º53'25"E), the initial symptom is a yellow and brown lesion on the edge of the leaf, in the late stage, brown blighted tissue caused leaves to curl and abscise. Ten typical diseased leaves were collected, the margins of infected tissues were cut into small pieces (5×5 mm) and were sterilized in 75% Ethanol for 30 s, 3% sodium hypochlorite for 3 min, rinsed three times with sterile water, then dried on sterile filter paper and placed to potato dextrose agar (PDA) medium at 25℃for 4 days. Thirteen isolates with morphological characteristics similar to those of Fusarium spp. (Nelson et al. 1983) were recovered. These isolates had white, pink and yellowish mycelia, two isolates produced irregular colonies, and remaining isolates showed round. Two of each type were selected for intensive study (yz2, yz11, yz9 and yz17). The colony of yz2 reached 62 mm in diameter on PDA medium after seven days, macroconidia were elongated sickle-shaped, 3-5 septa, and 12.92 to 21.49 × 3.42 to 5.90 µm in size, microconidia were oval and measured 5.69 to 12.95 × 3.41 to 9.80 µm in size, conidiophores were whorled and branched, yz9 attained 74 mm in diameter after nine days, macroconidia were curved sickle-shaped and apex cell acuminate, 26.9 to 57.2 × 2.4 to 7.1 µm, 3-5 septa. The microconidia were fusiform, 17.8 to 28.8 × 11.2 to 14.5 µm. Conidiophores variable in length. Genomic DNA was extracted from 7-day-old aerial mycelia of four strains (yz2, yz9, yz11 and yz17). The internal transcribed spacer (ITS) region (White et al. 1990), translation elongation factor (EF-1α) (Cao et al. 2014) and partial RNA polymerase second largest subunit (RPB2) (Wang et al. 2019) gene regions were amplified and multilocus phylogenetic analysis was conducted, their sequences were deposited in NCBI Genbank with the following accession numbers: the strains of yz2 and yz11 with OQ829372 and OR236201 for ITS, OQ848594 and OR282462 for EF-1α, OR492296 and OR492297 for RPB2; yz9 and yz17 with OQ829383 and OR236222 for ITS, OQ848595 and OR282463 for EF-1α, OR492295 and OR492298 for RPB2. The ModelFinder was used to select the best-fit model in PhyloSuite v1.2.2, the Bayesian Inference method (BI) analysis was used to estimate the system relationship, yz9 and yz17 were identified as Fusarium ipomoeae, yz2 and yz11 were identified as Fusarium tricinctum. To verify Koch's postulates, 8 healthy plants of H. plantaginea (two-year-old) grown were rinsed with sterile water, after 5 leaves per plant were stabbed with a sterilized needle, 4 plants were inoculated with conidial suspension (1×106 conidia mL-1), other plants injected with sterile water as control, then placed in a greenhouse maintained with 95% relative humidity at 25 ± 1°C. The symptoms on the leaves were similar to field after inoculation for 7 days, whereas all control leaves remained healthy. The same pathogen was re-isolated and re-identified based on multilocus phylogenetic analysis, fulfilling Koch's postulates. To our knowledge, this is the first report of F. ipomoeae causing leaf rot on H. plantaginea in China. In addition, F. ipomoeae was reported to cause leaf spot in Peanut (Xu et al. 2021), and F. tricinctum can cause fruit rot on navel orange in China (Yang et al. 2023). H. plantaginea as a horticultural plant is popular with some people, but it has long been threatened by Fusarium.spp. The finding can provide a theoretical basis for control leaf rot on H. plantaginea.

9.
Genomics ; 115(6): 110733, 2023 Nov.
Artigo em Inglês | MEDLINE | ID: mdl-37866659

RESUMO

BACKGROUND: Big data mining and experiments are widely used to mine new prognostic markers. METHODS: Candidate genes were identified from CROEMINE and FerrDb. Kaplan-Meier survival and Cox regression analysis were applied to assess the association of genes with Overall survival time (OS) and Disease-free survival time (DFS) in two HCC cohorts. Real-time quantitative polymerase chain reaction (RT-qPCR) and Immunohistochemistry were performed in HCC samples. RESULTS: 21 and 15 genes that can predict OS and DFS, which had not been reported before, were identified from 719 genes, respectively. Survival analysis showed elevated mRNA expression of GLMP, SLC38A6, and WDR76 were associated with poor prognosis, and three genes combination signature was an independent prognostic factor in HCC. RT-qPCR and Immunohistochemistry confirmed the results. CONCLUSIONS: We established a novel computational process, which identified the expression levels of GLMP, SLC38A6, and WDR76 as potential ferroptosis-related biomarkers indicating the prognosis of HCC.


Assuntos
Carcinoma Hepatocelular , Ferroptose , Neoplasias Hepáticas , Humanos , Carcinoma Hepatocelular/metabolismo , Neoplasias Hepáticas/metabolismo , Ferroptose/genética , Biomarcadores Tumorais/genética , Biomarcadores Tumorais/metabolismo , Estimativa de Kaplan-Meier , Prognóstico , Proteínas de Ligação a DNA , Proteínas de Ciclo Celular
10.
Plant Dis ; 2023 Sep 12.
Artigo em Inglês | MEDLINE | ID: mdl-37700475

RESUMO

Bletilla striata (named "Bai Ji" in Chinese) is a plant from the Orchidaceae family that has been employed in traditional Chinese medicine for thousands of years in China. Polysaccharides extracted from B. striata have been shown to have an effect on Alzheimer's disease (Lin et al. 2021). Since 2021, leaf spots have been observed in the B. striata plantation in Chongqing, China. Out of 200 plants, the disease incidence was estimated at 56%, and the disease index was estimated at 32%. The symptoms were necrotic lesions with brown edges and yellow halos; severe infection caused the infected leaves to become blighted, dry and fall off. To identify the causal agent, eighteen leaves with typical symptoms were collected from the B. striata plantation (30.60°N, 108.64°E). The margins of infected tissue areas were cut into small pieces (5×5 mm), surface sterilized with 70% ethanol for 1 min, and rinsed twice with sterile distilled water. The tissue was then surface sterilized in 3% sodium hypochlorite for 2 min, followed by three rinses with sterile water. The tissue was then placed onto potato dextrose agar (PDA) plates and incubated at 25°C for 3 days, pure cultures of fungal isolates were obtained by single-spore isolation, stored on PDA slants and maintained at 4°C. Colonies of the fungal isolates showed three color types, ranging from grayish white to green above with olive green on the reverse, but conidial characteristics were more similar and indicated this was a single fungus. Conidiophores were single, lateral from hyphae or terminal; straight or curved; smooth-walled with 1 to 8 septa; pale brown; usually with only one pigmented terminal conidiogenous site, sometimes with one additional lateral conidiogenous locus; sometimes slightly swollen at the apex; and 15 to 170 µm long, 2.5 to 4.5 µm wide. Conidia were in short or moderately long chains of 2-8 conidia normally, sometimes with more; rarely branched; normally 14.07 to 50 × 5.24 to 10 µm in size; ellipsoid, fusiform, long ellipsoid, obclavate or ovoid with 1 to 11 transverse septa and 2 to 4 longitudinal septa; beakless or with subcylindric or cylindric secondary conidiophores, analogous to the beak 4.25 to 58.6 µm long, 3.2 to 4.8 µm wide. The fungal isolates were tentatively identified as Alternaria sp. The representative isolate BJ8 was selected for the pathogenicity test. The leaves of six healthy plants of B. striata (two years old) grown in pots were washed with sterile water. Ten mL of conidial suspension (1×106 conidia mL-1) contained in 0.05% Tween 80 buffer was brushed onto upper and lower surfaces of all the leaves on three plants, while other plants were brushed with 10 mL 0.05% Tween 80 buffer to serve as controls. Plants were placed in a greenhouse at 25°C and 95±1% relative humidity after inoculation and observed for symptoms. The symptoms initially developed as irregular brown necrotic lesions on the inoculated leaves after 7 days, with a yellow halo around the lesions, consistent with the symptoms in the field. Leaves on the control plants did not produce any symptoms. For molecular identification, the genomic DNAs of representative isolates BJ5, BJ6, and BJ8 were extracted. The internal transcribed spacer (ITS) region and RNA polymerase II second largest subunit (RPB2), translation elongation factor 1-alpha (TEF1), and glyceraldehyde-3-phosphate dehydrogenase (GAPDH) genes were used for polymerase chain reaction (PCR), using primers ITS5/ITS4, GPD1/GPD2, EF-1F/EF-1B and RPB27cR/RPB25F2, respectively (White et al. 1990; Berbee and Pirseyedi et al. 1999; Carbone and Kohn 1999; Liu et al. 1999). The neighbor-joining tree revealed that these isolates are clustered together with the reference strain of A. burnsii. The sequences were deposited in NCBI GenBank BJ5 [ITS: OP897263; GAPDH: OQ544937; TEF1: OQ544941; RPB2: OQ544939], BJ6 [ITS: OP897262; GAPDH: OQ544938; TEF1: OQ544942; RPB2: OQ544940], and BJ8 [ITS: OK285209; GAPDH: OK340046; TEF1: OK340047; RPB2: OQ544936]. All three isolates showed 100% similarity with A. burnsii CBS 107.38 [ITS: KP124420; GAPDH: JQ646305; TEF1: KP125198; RPB2: JQ646457] ex-type sequence, thus the pathogen causing the leaf spot on B. striata was identified as A. burnsii. A. burnsii is an important pathogenic fungus causing blight of cumin (Shekhawat et al. 2013). Furthermore, Al-Nadabi et al. (2018) found that A. burnsii can cause leaf spots on wheat and date palms, and Sunapao et al. (2022) reported that A. burnsii can infect coconuts (Cocos nucifera), causing dirty panicle disease. This is the first report of A. burnsii causing leaf spot on B. striata in China. The new discovery shows that since A. burnsii can readily adapt to a variety of climatic conditions, controlling the fungus is crucial for the healthy growth of B. striata in the future. This study will provide a basis for further elucidating the pathogenic mechanism and development of effective control measures for this disease.

11.
Front Pharmacol ; 14: 1144824, 2023.
Artigo em Inglês | MEDLINE | ID: mdl-37426814

RESUMO

Background: Even 3 years into the COVID-19 pandemic, questions remain about how to safely and effectively vaccinate vulnerable populations. A systematic analysis of the safety and efficacy of the COVID-19 vaccine in at-risk groups has not been conducted to date. Methods: This study involved a comprehensive search of PubMed, EMBASE, and Cochrane Central Controlled Trial Registry data through 12 July 2022. Post-vaccination outcomes included the number of humoral and cellular immune responders in vulnerable and healthy populations, antibody levels in humoral immune responders, and adverse events. Results: A total of 23 articles assessing 32 studies, were included. The levels of IgG (SMD = -1.82, 95% CI [-2.28, -1.35]), IgA (SMD = -0.37, 95% CI [-0.70, -0.03]), IgM (SMD = -0.94, 95% CI [-1.38, -0.51]), neutralizing antibodies (SMD = -1.37, 95% CI [-2.62, -0.11]), and T cells (SMD = -1.98, 95% CI [-3.44, -0.53]) were significantly lower in vulnerable than in healthy populations. The positive detection rates of IgG (OR = 0.05, 95% CI [0.02, 0.14]) and IgA (OR = 0.03, 95% CI [0.01, 0.11]) antibodies and the cellular immune response rates (OR = 0.20, 95% CI [0.09, 0.45]) were also lower in the vulnerable populations. There were no statistically significant differences in fever (OR = 2.53, 95% CI [0.11, 60.86]), chills (OR = 2.03, 95% CI [0.08, 53.85]), myalgia (OR = 10.31, 95% CI [0.56, 191.08]), local pain at the injection site (OR = 17.83, 95% CI [0.32, 989.06]), headache (OR = 53.57, 95% CI [3.21, 892.79]), tenderness (OR = 2.68, 95% CI [0.49, 14.73]), and fatigue (OR = 22.89, 95% CI [0.45, 1164.22]) between the vulnerable and healthy populations. Conclusion: Seroconversion rates after COVID-19 vaccination were generally worse in the vulnerable than healthy populations, but there was no difference in adverse events. Patients with hematological cancers had the lowest IgG antibody levels of all the vulnerable populations, so closer attention to these patients is recommended. Subjects who received the combined vaccine had higher antibody levels than those who received the single vaccine.

12.
Pharm Biol ; 61(1): 839-857, 2023 Dec.
Artigo em Inglês | MEDLINE | ID: mdl-37203204

RESUMO

CONTEXT: Current chemotherapeutic drugs cannot meet the treatment needs of patients with nasopharyngeal carcinoma (NPC), so urgent action is needed to discover novel chemotherapeutic agents. Our previous study revealed that garcinone E (GE) inhibited the proliferation and metastasis of NPC, suggesting that the compound might display promising anticancer activity. OBJECTIVE: To examine the mechanism underlying the anti-NPC activity of GE for the first time. MATERIALS AND METHODS: For MTS assay, NPC cells were treated with 2.5-20 µmol/L GE or dimethyl sulfoxide for 24, 48, and 72 h. Colony formation capacity, cell cycle distribution, and in vivo xenograft experiment of GE were assessed. MDC staining, StubRFP-sensGFP-LC3 observation, LysoBrite Blue staining, and immunofluorescence examined the autophagy of NPC cells after GE exposure. Western blotting, RNA-sequencing, and RT-qPCR measured protein and mRNA levels. RESULTS: GE suppressed cell viability with an IC50 of 7.64, 8.83 and 4.65 µmol/L for HK1, HONE1 and S18 cells. GE inhibited colony formation and cell cycle, increased autophagosome number, and inhibited the autophagic flux partially by blocking lysosome-autophagosome fusion, and repressed S18 xenograft growth. GE dysregulated the expression of autophagy- and cell cycle-related proteins such as Beclin-1, SQSTM1/p62, LC3, CDKs, and Cyclins. Bioinformatics GO and KEGG pathway enrichment analysis of RNA-seq showed that autophagy was enriched in differentially expressed genes upon GE treatment. DISCUSSION AND CONCLUSION: GE acts as an autophagic flux inhibitor, which may have potential chemotherapeutic use for NPC treatment and may have an application in basic research to explore the mechanisms of autophagy.


Assuntos
Apoptose , Neoplasias Nasofaríngeas , Humanos , Carcinoma Nasofaríngeo/tratamento farmacológico , Carcinoma Nasofaríngeo/metabolismo , Proliferação de Células , Autofagia , Linhagem Celular Tumoral , Neoplasias Nasofaríngeas/tratamento farmacológico , Neoplasias Nasofaríngeas/metabolismo , Neoplasias Nasofaríngeas/patologia
13.
Plant Dis ; 2023 Feb 21.
Artigo em Inglês | MEDLINE | ID: mdl-36802293

RESUMO

Polygonatum cyrtonema Hua., is one of the cultivated varieties of Polygonatum sibiricum Redouté., which also an important cash crop in China (Chen, J., et al. 2021). From 2021 to 2022, symptoms resembling gray mold were observed on P. cyrtonema leaves with 30 to 45% disease incidence in Wanzhou District (30°38'1″N, 108°42'27″E) of Chongqing. The symptoms started to occur from April to June and more than 39% of leaves were infected from July to September. Symptoms started as irregular brown spots and progressed to the leaf edges or tips and stems. In dry conditions, the infected tissue appeared dry and thin, light brown in color, and became dry and cracked in the later stages of disease development. When the relative humidity was high, infected leaves developed water-soaked decay with a brown stripe around the lesion, and a gray mold layer appeared. To identify the causal agent, 8 typical diseased leaves were collected, leaf tissues were chopped into small pieces (3×5 mm), surface sterilized for 1 min in 70% ethanol and 5 minutes in 3% sodium hypochlorite, rinsed three times using sterile water, placed onto potato dextrose agar (PDA) amended with streptomycin sulfate (50 µg/ml) and incubated at 25°C for 3 days in dark conditions. Then 6 colonies (3.5 to 4 cm diameter) with similar morphology were transferred onto new plates. In the initial stage of growth of isolates, all hyphal colonies were white, dense, and clustered, and dispersed in all directions. After 21 days, brown to black-colored sclerotia (2.3 to 5.8 mm diameter) were observed embedded on the bottom of the medium. The six colonies were confirmed to be Botrytis sp. based on the morphological characteristics. The conidia were attached in branches on the conidiophores in grape-like clusters. Conidiophores were straight and 150 to 500 µm in length, and the conidia were single-celled, long ellipsoidal, or oval-like, with no septa and 7.5 to 20 × 3.5 to 14 µm (n=50). For molecular identification, DNA was extracted from representative strains 4-2 and 1-5. The internal transcribed spacer (ITS) region and sequences from the RNA polymerase II second largest subunit (RPB2), and the heat-shock protein 60 (HSP60) genes were amplified using primers ITS1/ITS4, RPB2for/RPB2rev, and HSP60for/HSP60rev, respectively (White T.J., et al.1990; Staats, M., et al. 2005). The sequences were deposited in GenBank: 4-2 [ITS; OM655229: RPB2; OM960678: HSP60; OM960679] and 1-5 [ITS; OQ160236: RPB2; OQ164790: HSP60; OQ164791]. These sequences from isolates 4-2 and 1-5 had 100% similarity to the B. deweyae CBS 134649/ MK-2013 [ITS; HG799538.1: RPB2; HG799518.1: HSP60; HG799519.1] ex-type sequences, and phylogenetic analyses based on multi-locus alignment demonstrated strains 4-2 and 1-5 as B. deweyae. Isolate 4-2 was used to verify whether B. deweyae can cause gray mold on P. cyrtonema, by conducting Koch's postulates experiments (Gradmann, C., 2014). The leaves of P. cyrtonema planted in pots were washed with sterile water, and brushed with 10 mL of hyphal tissue in 55% glycerin. Leaves of another plant were brushed with 10 mL 55% glycerin as control, and Kochs' postulates experiments were conducted three times. Inoculated plants were kept in a chamber with 80% relative humidity at 20 ± 1°C. Seven days after inoculation, disease symptoms similar to those in the field were observed on leaves, whereas control plants remained asymptomatic. The fungus was reisolated from inoculated plants and identified as B. deweyae based on multi-locus phylogenetic analysis. To our knowledge, B. deweyae is mostly found on Hemerocallis, is likely to be an important contributor to the development of 'spring sickness' symptoms (Grant-Downton, R.T., et al. 2014.), and this is the first report of B. deweyae causing gray mold on P. cyrtonema in China. Although B. deweyae has a limited host range, it might also become a potential threat to P. cyrtonema. This work will provide a basis for the prevention and treatment of the disease in the future.

15.
Int J Biol Macromol ; 230: 123117, 2023 Mar 01.
Artigo em Inglês | MEDLINE | ID: mdl-36603716

RESUMO

Wearable flexible sensors based on conductive hydrogels have received extensive attention in the fields of electronic skin and smart monitoring. However, conductive hydrogels contain a large amount of water, which greatly affects their performances in harsh environments. It is therefore necessary to prepare hydrogel sensors that are stable at low temperatures. Herein, metal ions (MgCl2) and ethylene glycol (EG) were combined with polyvinyl alcohol (PVA) to obtain a conductive PVA/EG hydrogel with tensile strength and elongation at break of 1.1 MPa and 442.3 %, respectively, which could withstand >6000-fold its own weight. The binary solvent system composed of water and EG contributed to the excellent anti-freezing properties and long-term storage (>1 week), flexibility, and stability of the hydrogel even at -20 °C. The wearable PVA/EG hydrogel as a flexible sensor possessed desirable sensing performances with a competitive GF value of 0.725 and fatigue resistance (50 cycles) when used to monitor various human motions and physiological signals. Overall, this hydrogel sensor shows strong potential for application in the fields of human motion monitoring, written information sensing, and information encryption and transmission.


Assuntos
Temperatura Baixa , Hidrogéis , Humanos , Condutividade Elétrica , Água , Etilenoglicóis
16.
Plant Dis ; 2022 Oct 03.
Artigo em Inglês | MEDLINE | ID: mdl-36190304

RESUMO

Euonymus japonicus Thunb., an evergreen shrub, is popular for landscaping in China. In 2021, leaf spot was observed on E. japonicus (about 150 trees) leaves with 40 to 50% disease incidence in Wanzhou urban forest (30°45'N; 108°27'E) of Chongqing, the infected plants were between 5 and 6 years old. The symptoms started to occur from June to July and approximately 30 to 40% of the leaves exhibited leaf spot symptoms from August to September. Initial symptoms appeared as yellow spots of 1.2 to 4.9 mm in diameter, and then expanded to become large and irregular lesions, having white center surrounded by a brown halo. Under humid conditions, black dots appeared in the central part of the spots. In later stage, split and fall of the tissues occurred from the infected spot. To identify the causal agent, infected tissues from 20 samples (from 5 trees) were cut into small pieces (5 mm2), surface-sterilized for 30 s in 75% ethanol and 3 minutes in 3% sodium hypochlorite, rinsed three times in sterile water, placed onto potato dextrose agar (PDA) amended with streptomycin sulfate (50 µg/ml) and incubated at 25°C in dark conditions. Purified eight fungal colonies were white with undulating margins and light cream on the reverse side, measuring 85 mm diameter after 7 days, dark brown to black conidiomata were irregularly scattered and Conidia were observed in 20 days old colonies. Conidia were spindle-shaped, 4.5 to 6.8 × 15.2 to 23.5 µm (n=50), with 4 diaphragms, the three median cells were light to dark brown and the two end cells were colorless. 1 to 3 accessory filaments (5.2 to 22.5 µm long) protrude from theapical cell while a short stalk (3.5 to 5.5 µm long) was attached to the basal cell, these morphological features suggested that the isolates were most likely Pestalotiopsis. sp. Eight colonies were confirmed to be identical based on morphological characteristics. For molecular identification, DNA was extracted from representative strains (YF-5, YF-13, YF-24). The internal transcribed spacer (ITS) region, ß-tubulin (TUB2), the translation elongation factor-1 alpha gene (TEF1), genes were amplified using primers ITS5/ITS4, TUB2F/TUB2R, and EF-526F/EF-1567R, respectively (White et al.1990; Glass & Donaldson 1995; O'Donnell & Cigelnik 1997; Carbone &, Kohn.1999). The sequences were deposited in NCBI GenBank YF-24, [ITS; ON204233: TUB2; ON304156: TEF1; ON400075]: YF-5, [ITS; OP379570: TUB2; OP413495: TEF1; OP413496]: YF-13, [ITS; OP379589: TUB2; OP413494: TEF1; OP413497]. Which revealed a 95% similarity to the Ps. theae NTUCC 18-067 [ITS; MT322086: TUB2; MT321888: TEF1; MT321987] ex-type sequences. Based on morphology and multilocus phylogenetic analysis, representative strains were identified as Pseudopestalotiopsis theae. For Koch's postulates, wiped the leaves of six healthy plants of E. japonicus (two-year-old) grown in pots with sterile water, 10 µL of spore suspension (106 spores/ mL) was brushed on five leaves per plant (three plants in total) with a sterile brush, and the other three plants were treated with sterile water instead of spore suspension as control, the plants were placed in a greenhouse at 28°C and 95±1% relative humidity. Seven days after inoculation, brown lesions appeared, similar to those observed in infected plants. Black dots surrounded by a brown halo reappear on the lesions after 12 days, whereas control plants remained healthy. Ps. theae culture was re-isolated from the infected leaves and identified using morphological characteristics and DNA sequence analysis. To our knowledge, Ps. theae can cause diseases on tea plants and has been found in Japan, Thailand and China, this is the first report of leaf spot infection of E. japonicus caused by Ps. theae in China. This disease is reducing the ornamental value of E. japonicus. Our results will contribute to the prevention and cure of leaf spot disease in E. japonicus.

17.
Mitochondrial DNA B Resour ; 7(8): 1445-1447, 2022.
Artigo em Inglês | MEDLINE | ID: mdl-35965647

RESUMO

Mallotus paniculatus (Lam.) Müll. Arg. 1865 (Euphorbiaceae) is a shrub or small tree with medicinal properties that is distributed across Southeast Asia. In this study, we sequenced the complete chloroplast genome of M. paniculatus to study phylogenetic relationships within the family Euphorbiaceae Juss. The complete chloroplast genome of M. paniculatus was 164,455 bp in length, with an overall GC content of 35.3%. It was found to consist of a long single copy region of 89,021 bp, a small single copy region of 18,524 bp, and a pair of inverted repeats of 28,455 bp. Results indicated that the chloroplast genome contains a total of 131 genes, including 78 protein-coding genes, 37 tRNA genes, eight rRNA genes, and eight pseudogenes. The phylogenetic tree showed that M. paniculatus is closely related to Mallotus japonicus and Mallotus peltatus.

18.
Mitochondrial DNA B Resour ; 7(7): 1329-1331, 2022.
Artigo em Inglês | MEDLINE | ID: mdl-35898663

RESUMO

In this study, the chloroplast (Cp) genome of Zanthoxylum avicennae (Lam.) DC was sequenced by high-throughput sequencing technology. The length of the Cp genome of Zanthoxylum avicennae was 158,568 bp, and the total GC content was 38.4%, including a large single-copy (LSC) region of 86,318 bp, a small single-copy (SSC) region of 18,250 bp, and 27,000 bp of inverted repeats (IRs). The Cp genome encoded 131 genes, including 88 protein-coding, 37 tRNA, and six rRNA genes. Phylogenetic analysis of the genome sequence showed that Zanthoxylum avicennae was closely related to Zanthoxylum nitidum, Zanthoxylum esquirolii and Zanthoxylum motuoense of the Rutaceae family.

19.
Mitochondrial DNA B Resour ; 7(6): 1129-1130, 2022.
Artigo em Inglês | MEDLINE | ID: mdl-35783059

RESUMO

Euphorbia micractina Boiss is a plant with high medicinal value. Yet, its molecular biology is not fully understood. In this study, we sequenced the whole chloroplast genome (CP) sequence of E. micractina to study its phylogenetic relationship in Euphorbiaceae. The total length of the chloroplast genome of E. micractina is 162,056 bp, including a large single-copy (LSC) region of 89,936bp bp, a small single-copy (SSC) region of 18,376 bp, and a pair of identical inverted repeat regions (IRs) of 11,367 bp. The genome has 128 genes, including 84 protein-coding genes, 36 transfer RNA (tRNA) genes, and 8 ribosomal RNA (rRNA) genes. The overall GC content of the plastome is 35.7%. The phylogenetic analysis of E. micractina with 30 related species suggested a closest taxonomical relationship with Euphorbia pekinensis in the Euphorbiaceae family.

20.
Mitochondrial DNA B Resour ; 7(6): 961-963, 2022.
Artigo em Inglês | MEDLINE | ID: mdl-35712542

RESUMO

Disporum viridescens is a medicinal plant in three provinces of Northeast China. In this paper, we report the characteristics of the complete chloroplast genome (CP) of Disporum viridescens. We discuss mainly the phylogenetic relationship between this species and its relatives. The length of its sequence was 156,645 bp, with a total GC content of 37.6% A large single-copy region (LSC) of 85,103 bp, a small single-copy region (SSC) of 17,964 bp, and a pair of inverted repeat (IR) regions of 26,789 bp were detected in this study. The complete chloroplast genome sequence contained 127 genes, including 81 encoding, 38 transfer RNA (tRNA), and 8 ribosomal RNA (rRNA) genes. Our phylogenetic analysis results showed that Disporum viridescens is closely phylogenetically related to Disporum cantoniense of the family Colchicaceae.

SELEÇÃO DE REFERÊNCIAS
DETALHE DA PESQUISA
...