49 research outputs found

    Silenciamiento de genes diana de Delottococcus aberiae DeLotto (Hemiptera: Pseudococcidae) mediante tecnologĂ­a RNAi

    Get PDF
    El RNA de interferencia (RNAi) es una de las herramientas biotecnolĂłgicas con mayor potencial en la protecciĂłn de cultivos frente a insectos plagas. Con esta tĂ©cnica de silenciamiento postranscripcional se puede bloquear la expresiĂłn de genes diana de interĂ©s al destruir eficazmente su correspondiente RNA mensajero (mRNA). En los Ășltimos años se han logrado avances significativos en varios aspectos del mecanismo de silenciamiento gĂ©nico postranscripcional mediado por molĂ©culas de RNA de doble cadena (dsRNA). En la actualidad el mundo cientĂ­fico se encuentra en disposiciĂłn de poder transferir y aplicar de manera prĂĄctica esta potente herramienta al control biolĂłgico de plagas

    Characterization of hemizygous deletions in Citrus using array-Comparative Genomic Hybridization and microsynteny comparisons with the poplar genome

    Get PDF
    <p>Abstract</p> <p>Background</p> <p>Many fruit-tree species, including relevant <it>Citrus </it>spp varieties exhibit a reproductive biology that impairs breeding and strongly constrains genetic improvements. In citrus, juvenility increases the generation time while sexual sterility, inbreeding depression and self-incompatibility prevent the production of homozygous cultivars. Genomic technology may provide citrus researchers with a new set of tools to address these various restrictions. In this work, we report a valuable genomics-based protocol for the structural analysis of deletion mutations on an heterozygous background.</p> <p>Results</p> <p>Two independent fast neutron mutants of self-incompatible clementine (<it>Citrus clementina </it>Hort. Ex Tan. cv. Clemenules) were the subject of the study. Both mutants, named 39B3 and 39E7, were expected to carry DNA deletions in hemizygous dosage. Array-based Comparative Genomic Hybridization (array-CGH) using a <it>Citrus </it>cDNA microarray allowed the identification of underrepresented genes in these two mutants. Subsequent comparison of citrus deleted genes with annotated plant genomes, especially poplar, made possible to predict the presence of a large deletion in 39B3 of about 700 kb and at least two deletions of approximately 100 and 500 kb in 39E7. The deletion in 39B3 was further characterized by PCR on available <it>Citrus </it>BACs, which helped us to build a partial physical map of the deletion. Among the deleted genes, <it>ClpC</it>-like gene coding for a putative subunit of a multifunctional chloroplastic protease involved in the regulation of chlorophyll <it>b </it>synthesis was directly related to the mutated phenotype since the mutant showed a reduced chlorophyll <it>a</it>/<it>b </it>ratio in green tissues.</p> <p>Conclusion</p> <p>In this work, we report the use of array-CGH for the successful identification of genes included in a hemizygous deletion induced by fast neutron irradiation on <it>Citrus clementina</it>. The study of gene content and order into the 39B3 deletion also led to the unexpected conclusion that microsynteny and local gene colinearity in this species were higher with <it>Populus trichocarpa </it>than with the phylogenetically closer <it>Arabidopsis thaliana</it>. This work corroborates the potential of <it>Citrus </it>genomic resources to assist mutagenesis-based approaches for functional genetics, structural studies and comparative genomics, and hence to facilitate citrus variety improvement.</p

    The Altered Expression of microRNA408 Influences the Arabidopsis Response to Iron Deficiency

    Get PDF
    MicroRNAs contribute to the adaptation of plants to varying environmental conditions by affecting systemic mineral nutrient homeostasis. Copper and iron deficiencies antagonistically control the expression of Arabidopsis thaliana microRNA408 (miR408), which post-transcriptionally regulates laccase-like multicopper oxidase family members LAC3, LAC12, and LAC13. In this work, we used miR408 T-DNA insertion mutants (408-KO1 and 408-KO2) and a previously characterized transgenic line overexpressing miR408 (35S:408-14) to explore how miR408 influences copper- and iron-dependent metabolism. We observed that the altered expression of miR408 diminished plant performance and the activation of the iron-regulated genes under iron-deficient conditions. Consistently with the low expression of the miR408-target laccases, we showed that the vascular bundle lignification of the 35S:408-14 plants diminished. The decrease in the phenoloxidase and ferroxidase activities exhibited by wild-type plants under iron deficiency did not occur in the 408-KO1 plants, probably due to the higher expression of laccases. Finally, we observed that the hydrogen peroxide levels under iron starvation were altered in both the 408-KO1 and 35S:408-14 lines. Taken together, these results suggest that Arabidopsis plants with modified miR408 levels undergo multiple deregulations under iron-deficient conditions

    Characterization of the procera

    Full text link

    Identification of stress associated microRNAs in Solanum lycopersicum by high-throughput Sequencing

    Get PDF
    Tomato (Solanum lycopersicum) is one of the most important crops around the world and also a model plant to study response to stress. High-throughput sequencing was used to analyse the microRNA (miRNA) profile of tomato plants undergoing five biotic and abiotic stress conditions (drought, heat, P. syringae infection, B. cinerea infection, and herbivore insect attack with Leptinotarsa decemlineata larvae) and one chemical treatment with a plant defence inducer, hexanoic acid. We identified 104 conserved miRNAs belonging to 37 families and we predicted 61 novel tomato miRNAs. Among those 165 miRNAs, 41 were stress-responsive. Reverse transcription quantitative PCR (RT-qPCR) was used to validate high-throughput expression analysis data, confirming the expression profiles of 10 out of 11 randomly selected miRNAs. Most of the differentially expressed miRNAs were stress-specific, except for sly-miR167c-3p upregulated in B. cinerea and P. syringae infection, sly-newmiR26-3p upregulated in drought and Hx treatment samples, and sly-newmiR33-3p, sly-newmiR6-3p and sly-newmiR8-3p differentially expressed both in biotic and abiotic stresses. From mature miRNAs sequences of the 41 stress-responsive miRNAs 279 targets were predicted. An inverse correlation between the expression profiles of 4 selected miRNAs (sly-miR171a, sly-miR172c, sly-newmiR22-3p and sly-miR167c-3p) and their target genes (Kinesin, PPR, GRAS40, ABC transporter, GDP and RLP1) was confirmed by RT-qPCR. Altogether, our analysis of miRNAs in different biotic and abiotic stress conditions highlight the interest to understand the functional role of miRNAs in tomato stress response as well as their putative targets which could help to elucidate plants molecular and physiological adaptation to stress

    Regulation of ovule initiation by gibberellins and brassinosteroids in tomato and Arabidopsis: two plant species, two molecular mechanisms

    Full text link
    This is the peer reviewed version of the following article: BarroÂżTrastoy, D., Carrera, E., Baños, J., Palau-RodrĂ­guez, J., Ruiz-Rivero, O., Tornero, P., Alonso, J.M., LĂłpez-DĂ­az, I., GĂłmez, M.D. and PĂ©rez-Amador, M.A. (2020), Regulation of ovule initiation by gibberellins and brassinosteroids in tomato and Arabidopsis: two plant species, two molecular mechanisms. Plant J, 102: 1026-1041, which has been published in final form at https://doi.org/10.1111/tpj.14684. This article may be used for non-commercial purposes in accordance with Wiley Terms and Conditions for Self-Archiving.[EN] Ovule primordia formation is a complex developmental process with a strong impact on the production of seeds. In Arabidopsis this process is controlled by a gene network, including components of the signalling pathways of auxin, brassinosteroids (BRs) and cytokinins. Recently, we have shown that gibberellins (GAs) also play an important role in ovule primordia initiation, inhibiting ovule formation in both Arabidopsis and tomato. Here we reveal that BRs also participate in the control of ovule initiation in tomato, by promoting an increase on ovule primordia formation. Moreover, molecular and genetic analyses of the co-regulation by GAs and BRs of the control of ovule initiation indicate that two different mechanisms occur in tomato and Arabidopsis. In tomato, GAs act downstream of BRs. BRs regulate ovule number through the downregulation of GA biosynthesis, which provokes stabilization of DELLA proteins that will finally promote ovule primordia initiation. In contrast, in Arabidopsis both GAs and BRs regulate ovule number independently of the activity levels of the other hormone. Taken together, our data strongly suggest that different molecular mechanisms could operate in different plant species to regulate identical developmental processes even, as for ovule primordia initiation, if the same set of hormones trigger similar responses, adding a new level of complexity.We wish to thank B. Janssen (Horticulture and Food Research Institute, New Zealand) for the pBJ60 shuttle vector, C. Ferrandiz and M. Colombo (IBMCP, CSIC-UPV, Valencia, Spain) for their help in the generation of 35S:ANT lines and L.E.P. Peres (Universidade de Sao Paulo, Brazil) for the tomato mutant lines. Our thanks also go to C. Fuster for technical assistance. This work was supported by grants from the Spanish Ministry of Economy and Competitiveness-FEDER (BIO2017-83138R) to MAPA and from NSF (DBI-0820755, MCB-1158181, and IOS-1444561) to JMA.Barro-Trastoy, D.; Carrera, E.; Baños, J.; Palau-RodrĂ­guez, J.; Ruiz-Rivero, O.; Tornero Feliciano, P.; Alonso, JM.... (2020). Regulation of ovule initiation by gibberellins and brassinosteroids in tomato and Arabidopsis: two plant species, two molecular mechanisms. The Plant Journal. 102(5):1026-1041. https://doi.org/10.1111/tpj.14684S102610411025Azhakanandam, S., Nole-Wilson, S., Bao, F., & Franks, R. G. (2008). SEUSSandAINTEGUMENTAMediate Patterning and Ovule Initiation during Gynoecium Medial Domain Development    . Plant Physiology, 146(3), 1165-1181. doi:10.1104/pp.107.114751Bai, M.-Y., Shang, J.-X., Oh, E., Fan, M., Bai, Y., Zentella, R., 
 Wang, Z.-Y. (2012). Brassinosteroid, gibberellin and phytochrome impinge on a common transcription module in Arabidopsis. Nature Cell Biology, 14(8), 810-817. doi:10.1038/ncb2546Baker, S. C., Robinson-Beers, K., Villanueva, J. M., Gaiser, J. C., & Gasser, C. S. (1997). Interactions Among Genes Regulating Ovule Development in Arabidopsis thaliana. Genetics, 145(4), 1109-1124. doi:10.1093/genetics/145.4.1109Bartrina, I., Otto, E., Strnad, M., Werner, T., & SchmĂŒlling, T. (2011). Cytokinin Regulates the Activity of Reproductive Meristems, Flower Organ Size, Ovule Formation, and Thus Seed Yield in Arabidopsis thaliana      . The Plant Cell, 23(1), 69-80. doi:10.1105/tpc.110.079079Belkhadir, Y., & Jaillais, Y. (2015). The molecular circuitry of brassinosteroid signaling. New Phytologist, 206(2), 522-540. doi:10.1111/nph.13269Bencivenga, S., Simonini, S., BenkovĂĄ, E., & Colombo, L. (2012). The Transcription Factors BEL1 and SPL Are Required for Cytokinin and Auxin Signaling During Ovule Development in Arabidopsis. The Plant Cell, 24(7), 2886-2897. doi:10.1105/tpc.112.100164Brumos, J., Zhao, C., Gong, Y., Soriano, D., Patel, A. P., Perez-Amador, M. A., 
 Alonso, J. M. (2019). An Improved Recombineering Toolset for Plants. The Plant Cell, 32(1), 100-122. doi:10.1105/tpc.19.00431Carrera, E., Ruiz-Rivero, O., Peres, L. E. P., Atares, A., & Garcia-Martinez, J. L. (2012). Characterization of the procera Tomato Mutant Shows Novel Functions of the SlDELLA Protein in the Control of Flower Morphology, Cell Division and Expansion, and the Auxin-Signaling Pathway during Fruit-Set and Development    . Plant Physiology, 160(3), 1581-1596. doi:10.1104/pp.112.204552Carvalho, R. F., Campos, M. L., Pino, L. E., Crestana, S. L., Zsögön, A., Lima, J. E., 
 Peres, L. E. (2011). Convergence of developmental mutants into a single tomato model system: «Micro-Tom» as an effective toolkit for plant development research. Plant Methods, 7(1). doi:10.1186/1746-4811-7-18Chory, J., Nagpal, P., & Peto, C. A. (1991). Phenotypic and Genetic Analysis of det2, a New Mutant That Affects Light-Regulated Seedling Development in Arabidopsis. The Plant Cell, 3(5), 445. doi:10.2307/3869351Clough, S. J., & Bent, A. F. (1998). Floral dip: a simplified method forAgrobacterium-mediated transformation ofArabidopsis thaliana. The Plant Journal, 16(6), 735-743. doi:10.1046/j.1365-313x.1998.00343.xClouse, S. D. (2011). Brassinosteroid Signal Transduction: From Receptor Kinase Activation to Transcriptional Networks Regulating Plant Development. The Plant Cell, 23(4), 1219-1230. doi:10.1105/tpc.111.084475Cucinotta, M., Colombo, L., & Roig-Villanova, I. (2014). Ovule development, a new model for lateral organ formation. Frontiers in Plant Science, 5. doi:10.3389/fpls.2014.00117Cucinotta, M., Manrique, S., Guazzotti, A., Quadrelli, N. E., Mendes, M. A., Benkova, E., & Colombo, L. (2016). Cytokinin response factors integrate auxin and cytokinin pathways for female reproductive organ development. Development. doi:10.1242/dev.143545DaviĂšre, J.-M., & Achard, P. (2016). A Pivotal Role of DELLAs in Regulating Multiple Hormone Signals. Molecular Plant, 9(1), 10-20. doi:10.1016/j.molp.2015.09.011De Vleesschauwer, D., Van Buyten, E., Satoh, K., Balidion, J., Mauleon, R., Choi, I.-R., 
 Höfte, M. (2012). Brassinosteroids Antagonize Gibberellin- and Salicylate-Mediated Root Immunity in Rice      . Plant Physiology, 158(4), 1833-1846. doi:10.1104/pp.112.193672Dorcey, E., Urbez, C., BlĂĄzquez, M. A., Carbonell, J., & Perez-Amador, M. A. (2009). Fertilization-dependent auxin response in ovules triggers fruit development through the modulation of gibberellin metabolism in Arabidopsis. The Plant Journal, 58(2), 318-332. doi:10.1111/j.1365-313x.2008.03781.xFujioka, S., Li, J., Choi, Y. H., Seto, H., Takatsuto, S., Noguchi, T., 
 Sakurai, A. (1997). The Arabidopsis deetiolated2 mutant is blocked early in brassinosteroid biosynthesis. The Plant Cell, 9(11), 1951-1962. doi:10.1105/tpc.9.11.1951Galbiati, F., Sinha Roy, D., Simonini, S., Cucinotta, M., Ceccato, L., Cuesta, C., 
 Colombo, L. (2013). An integrative model of the control of ovule primordia formation. The Plant Journal, 76(3), 446-455. doi:10.1111/tpj.12309Gallego-Bartolome, J., Minguet, E. G., Grau-Enguix, F., Abbas, M., Locascio, A., Thomas, S. G., 
 Blazquez, M. A. (2012). Molecular mechanism for the interaction between gibberellin and brassinosteroid signaling pathways in Arabidopsis. Proceedings of the National Academy of Sciences, 109(33), 13446-13451. doi:10.1073/pnas.1119992109GarcĂ­a-Hurtado, N., Carrera, E., Ruiz-Rivero, O., LĂłpez-Gresa, M. P., Hedden, P., Gong, F., & GarcĂ­a-MartĂ­nez, J. L. (2012). The characterization of transgenic tomato overexpressing gibberellin 20-oxidase reveals induction of parthenocarpic fruit growth, higher yield, and alteration of the gibberellin biosynthetic pathway. Journal of Experimental Botany, 63(16), 5803-5813. doi:10.1093/jxb/ers229Gleave, A. P. (1992). A versatile binary vector system with a T-DNA organisational structure conducive to efficient integration of cloned DNA into the plant genome. Plant Molecular Biology, 20(6), 1203-1207. doi:10.1007/bf00028910Gomez, M. D., Ventimilla, D., Sacristan, R., & Perez-Amador, M. A. (2016). Gibberellins Regulate Ovule Integument Development by Interfering with the Transcription Factor ATS. Plant Physiology, 172(4), 2403-2415. doi:10.1104/pp.16.01231He, J.-X., Gendron, J. M., Sun, Y., Gampala, S. S. L., Gendron, N., Sun, C. Q., & Wang, Z.-Y. (2005). BZR1 Is a Transcriptional Repressor with Dual Roles in Brassinosteroid Homeostasis and Growth Responses. Science, 307(5715), 1634-1638. doi:10.1126/science.1107580Huang, H.-Y., Jiang, W.-B., Hu, Y.-W., Wu, P., Zhu, J.-Y., Liang, W.-Q., 
 Lin, W.-H. (2013). BR Signal Influences Arabidopsis Ovule and Seed Number through Regulating Related Genes Expression by BZR1. Molecular Plant, 6(2), 456-469. doi:10.1093/mp/sss070Kurepin, L. V., Joo, S.-H., Kim, S.-K., Pharis, R. P., & Back, T. G. (2011). Interaction of Brassinosteroids with Light Quality and Plant Hormones in Regulating Shoot Growth of Young Sunflower and Arabidopsis Seedlings. Journal of Plant Growth Regulation, 31(2), 156-164. doi:10.1007/s00344-011-9227-7Li, Q.-F., Wang, C., Jiang, L., Li, S., Sun, S. S. M., & He, J.-X. (2012). An Interaction Between BZR1 and DELLAs Mediates Direct Signaling Crosstalk Between Brassinosteroids and Gibberellins in Arabidopsis. Science Signaling, 5(244). doi:10.1126/scisignal.2002908Li, X.-J., Chen, X.-J., Guo, X., Yin, L.-L., Ahammed, G. J., Xu, C.-J., 
 Yu, J.-Q. (2015). DWARFoverexpression induces alteration in phytohormone homeostasis, development, architecture and carotenoid accumulation in tomato. Plant Biotechnology Journal, 14(3), 1021-1033. doi:10.1111/pbi.12474Liu, Z., Franks, R. G., & Klink, V. P. (2000). Regulation of Gynoecium Marginal Tissue Formation by LEUNIG and AINTEGUMENTA. The Plant Cell, 12(10), 1879-1891. doi:10.1105/tpc.12.10.1879Marti, E. (2006). Genetic and physiological characterization of tomato cv. Micro-Tom. Journal of Experimental Botany, 57(9), 2037-2047. doi:10.1093/jxb/erj154Mizukami, Y., & Fischer, R. L. (2000). Plant organ size control: AINTEGUMENTA regulates growth and cell numbers during organogenesis. Proceedings of the National Academy of Sciences, 97(2), 942-947. doi:10.1073/pnas.97.2.942Montoya, T., Nomura, T., Yokota, T., Farrar, K., Harrison, K., Jones, J. G. D., 
 Bishop, G. J. (2005). Patterns of Dwarf expression and brassinosteroid accumulation in tomato reveal the importance of brassinosteroid synthesis during fruit development. The Plant Journal, 42(2), 262-269. doi:10.1111/j.1365-313x.2005.02376.xMĂŒller, C. J., Larsson, E., SpĂ­chal, L., & Sundberg, E. (2017). Cytokinin-Auxin Crosstalk in the Gynoecial Primordium Ensures Correct Domain Patterning. Plant Physiology, 175(3), 1144-1157. doi:10.1104/pp.17.00805Murashige, T., & Skoog, F. (1962). A Revised Medium for Rapid Growth and Bio Assays with Tobacco Tissue Cultures. Physiologia Plantarum, 15(3), 473-497. doi:10.1111/j.1399-3054.1962.tb08052.xOlimpieri, I., Siligato, F., Caccia, R., Soressi, G. P., Mazzucato, A., Mariotti, L., & Ceccarelli, N. (2007). Tomato fruit set driven by pollination or by the parthenocarpic fruit allele are mediated by transcriptionally regulated gibberellin biosynthesis. Planta, 226(4), 877-888. doi:10.1007/s00425-007-0533-zPaz-Ares, J., & The REGIA Consortium. (2002). REGIA, An EU Project on Functional Genomics of Transcription Factors fromArabidopsis thaliana. Comparative and Functional Genomics, 3(2), 102-108. doi:10.1002/cfg.146Peng, J., Carol, P., Richards, D. E., King, K. E., Cowling, R. J., Murphy, G. P., & Harberd, N. P. (1997). The Arabidopsis GAI gene defines a signaling pathway that negatively regulates gibberellin responses . Genes & Development, 11(23), 3194-3205. doi:10.1101/gad.11.23.3194Reyes-Olalde, J. I., Zuñiga-Mayo, V. M., ChĂĄvez Montes, R. A., Marsch-MartĂ­nez, N., & de Folter, S. (2013). Inside the gynoecium: at the carpel margin. Trends in Plant Science, 18(11), 644-655. doi:10.1016/j.tplants.2013.08.002Sabelli, P. A., & Larkins, B. A. (2009). The Development of Endosperm in Grasses. Plant Physiology, 149(1), 14-26. doi:10.1104/pp.108.129437Schneitz, K., Baker, S. C., Gasser, C. S., & Redweik, A. (1998). Pattern formation and growth during floral organogenesis: HUELLENLOS and AINTEGUMENTA are required for the formation of the proximal region of the ovule primordium in Arabidopsis thaliana. Development, 125(14), 2555-2563. doi:10.1242/dev.125.14.2555Schneitz, K., Hulskamp, M., & Pruitt, R. E. (1995). Wild-type ovule development in Arabidopsis thaliana: a light microscope study of cleared whole-mount tissue. The Plant Journal, 7(5), 731-749. doi:10.1046/j.1365-313x.1995.07050731.xSeo, M., Jikumaru, Y., & Kamiya, Y. (2011). Profiling of Hormones and Related Metabolites in Seed Dormancy and Germination Studies. Methods in Molecular Biology, 99-111. doi:10.1007/978-1-61779-231-1_7Serrani, J. C., Sanjuán, R., Ruiz-Rivero, O., Fos, M., & García-Martínez, J. L. (2007). Gibberellin Regulation of Fruit Set and Growth in Tomato. Plant Physiology, 145(1), 246-257. doi:10.1104/pp.107.098335Serrani, J. C., Carrera, E., Ruiz-Rivero, O., Gallego-Giraldo, L., Peres, LĂĄ. E. P., & GarcĂ­a-MartĂ­nez, J. L. (2010). Inhibition of Auxin Transport from the Ovary or from the Apical Shoot Induces Parthenocarpic Fruit-Set in Tomato Mediated by Gibberellins    . Plant Physiology, 153(2), 851-862. doi:10.1104/pp.110.155424Sun, T. (2010). Gibberellin-GID1-DELLA: A Pivotal Regulatory Module for Plant Growth and Development. Plant Physiology, 154(2), 567-570. doi:10.1104/pp.110.161554Sun, T. (2011). The Molecular Mechanism and Evolution of the GA–GID1–DELLA Signaling Module in Plants. Current Biology, 21(9), R338-R345. doi:10.1016/j.cub.2011.02.036Tanaka, K., Nakamura, Y., Asami, T., Yoshida, S., Matsuo, T., & Okamoto, S. (2003). Physiological Roles of Brassinosteroids in Early Growth of Arabidopsis: Brassinosteroids Have a Synergistic Relationship with Gibberellin as well as Auxin in Light-Grown Hypocotyl Elongation. Journal of Plant Growth Regulation, 22(3), 259-271. doi:10.1007/s00344-003-0119-3Tang, Y., Liu, H., Guo, S., Wang, B., Li, Z., Chong, K., & Xu, Y. (2017). OsmiR396d Affects Gibberellin and Brassinosteroid Signaling to Regulate Plant Architecture in Rice. Plant Physiology, 176(1), 946-959. doi:10.1104/pp.17.00964Tong, H., Xiao, Y., Liu, D., Gao, S., Liu, L., Yin, Y., 
 Chu, C. (2014). Brassinosteroid Regulates Cell Elongation by Modulating Gibberellin Metabolism in Rice    . The Plant Cell, 26(11), 4376-4393. doi:10.1105/tpc.114.132092Truernit, E., Bauby, H., Dubreucq, B., Grandjean, O., Runions, J., Barthélémy, J., & Palauqui, J.-C. (2008). High-Resolution Whole-Mount Imaging of Three-Dimensional Tissue Organization and Gene Expression Enables the Study of Phloem Development and Structure inArabidopsis . The Plant Cell, 20(6), 1494-1503. doi:10.1105/tpc.107.056069Tursun, B., Cochella, L., Carrera, I., & Hobert, O. (2009). A Toolkit and Robust Pipeline for the Generation of Fosmid-Based Reporter Genes in C. elegans. PLoS ONE, 4(3), e4625. doi:10.1371/journal.pone.0004625Unterholzner, S. J., Rozhon, W., Papacek, M., Ciomas, J., Lange, T., Kugler, K. G., 
 Poppenberger, B. (2015). Brassinosteroids Are Master Regulators of Gibberellin Biosynthesis in Arabidopsis. The Plant Cell, 27(8), 2261-2272. doi:10.1105/tpc.15.00433Wang, Z.-Y., Nakano, T., Gendron, J., He, J., Chen, M., Vafeados, D., 
 Chory, J. (2002). Nuclear-Localized BZR1 Mediates Brassinosteroid-Induced Growth and Feedback Suppression of Brassinosteroid Biosynthesis. Developmental Cell, 2(4), 505-513. doi:10.1016/s1534-5807(02)00153-3Xiao, H., Radovich, C., Welty, N., Hsu, J., Li, D., Meulia, T., & van der Knaap, E. (2009). Integration of tomato reproductive developmental landmarks and expression profiles, and the effect of SUN on fruit shape. BMC Plant Biology, 9(1). doi:10.1186/1471-2229-9-49Xiao, Y., Liu, D., Zhang, G., Tong, H., & Chu, C. (2017). Brassinosteroids Regulate OFP1, a DLT Interacting Protein, to Modulate Plant Architecture and Grain Morphology in Rice. Frontiers in Plant Science, 8. doi:10.3389/fpls.2017.0169

    Effects of hospital facilities on patient outcomes after cancer surgery: an international, prospective, observational study

    Get PDF
    Background Early death after cancer surgery is higher in low-income and middle-income countries (LMICs) compared with in high-income countries, yet the impact of facility characteristics on early postoperative outcomes is unknown. The aim of this study was to examine the association between hospital infrastructure, resource availability, and processes on early outcomes after cancer surgery worldwide.Methods A multimethods analysis was performed as part of the GlobalSurg 3 study-a multicentre, international, prospective cohort study of patients who had surgery for breast, colorectal, or gastric cancer. The primary outcomes were 30-day mortality and 30-day major complication rates. Potentially beneficial hospital facilities were identified by variable selection to select those associated with 30-day mortality. Adjusted outcomes were determined using generalised estimating equations to account for patient characteristics and country-income group, with population stratification by hospital.Findings Between April 1, 2018, and April 23, 2019, facility-level data were collected for 9685 patients across 238 hospitals in 66 countries (91 hospitals in 20 high-income countries; 57 hospitals in 19 upper-middle-income countries; and 90 hospitals in 27 low-income to lower-middle-income countries). The availability of five hospital facilities was inversely associated with mortality: ultrasound, CT scanner, critical care unit, opioid analgesia, and oncologist. After adjustment for case-mix and country income group, hospitals with three or fewer of these facilities (62 hospitals, 1294 patients) had higher mortality compared with those with four or five (adjusted odds ratio [OR] 3.85 [95% CI 2.58-5.75]; p&lt;0.0001), with excess mortality predominantly explained by a limited capacity to rescue following the development of major complications (63.0% vs 82.7%; OR 0.35 [0.23-0.53]; p&lt;0.0001). Across LMICs, improvements in hospital facilities would prevent one to three deaths for every 100 patients undergoing surgery for cancer.Interpretation Hospitals with higher levels of infrastructure and resources have better outcomes after cancer surgery, independent of country income. Without urgent strengthening of hospital infrastructure and resources, the reductions in cancer-associated mortality associated with improved access will not be realised

    Use of anticoagulants and antiplatelet agents in stable outpatients with coronary artery disease and atrial fibrillation. International CLARIFY registry

    Get PDF

    Omecamtiv mecarbil in chronic heart failure with reduced ejection fraction, GALACTIC‐HF: baseline characteristics and comparison with contemporary clinical trials

    Get PDF
    Aims: The safety and efficacy of the novel selective cardiac myosin activator, omecamtiv mecarbil, in patients with heart failure with reduced ejection fraction (HFrEF) is tested in the Global Approach to Lowering Adverse Cardiac outcomes Through Improving Contractility in Heart Failure (GALACTIC‐HF) trial. Here we describe the baseline characteristics of participants in GALACTIC‐HF and how these compare with other contemporary trials. Methods and Results: Adults with established HFrEF, New York Heart Association functional class (NYHA) ≄ II, EF ≀35%, elevated natriuretic peptides and either current hospitalization for HF or history of hospitalization/ emergency department visit for HF within a year were randomized to either placebo or omecamtiv mecarbil (pharmacokinetic‐guided dosing: 25, 37.5 or 50 mg bid). 8256 patients [male (79%), non‐white (22%), mean age 65 years] were enrolled with a mean EF 27%, ischemic etiology in 54%, NYHA II 53% and III/IV 47%, and median NT‐proBNP 1971 pg/mL. HF therapies at baseline were among the most effectively employed in contemporary HF trials. GALACTIC‐HF randomized patients representative of recent HF registries and trials with substantial numbers of patients also having characteristics understudied in previous trials including more from North America (n = 1386), enrolled as inpatients (n = 2084), systolic blood pressure &lt; 100 mmHg (n = 1127), estimated glomerular filtration rate &lt; 30 mL/min/1.73 m2 (n = 528), and treated with sacubitril‐valsartan at baseline (n = 1594). Conclusions: GALACTIC‐HF enrolled a well‐treated, high‐risk population from both inpatient and outpatient settings, which will provide a definitive evaluation of the efficacy and safety of this novel therapy, as well as informing its potential future implementation
    corecore