161 research outputs found

    Characterization of Two Soybean (Glycine max L.) LEA IV Proteins by Circular Dichroism and Fourier Transform Infrared Spectrometry

    Get PDF
    Late embryogenesis-abundant (LEA) proteins, accumulating to a high level during the late stages of seed development, may play a role as osmoprotectants. However, the functions and mechanisms of LEA proteins remained to be elucidated. Five major groups of LEA proteins have been described. In the present study, we report on the characterization of two members of soybean LEA IV proteins, basic GmPM1 and acidic GmPM28, by circular dichroism and Fourier transform infrared spectroscopy. The spectra of both proteins revealed limited defined secondary structures in the fully hydrated state. Thus, the soybean LEA IV proteins are members of ‘natively unfolded proteins’. GmPM1 or GmPM28 proteins showed a conformational change under hydrophobic or dry conditions. After fast or slow drying, the two proteins showed slightly increased proportions of defined secondary structures (α-helix and β-sheet), from 30 to 49% and from 34 to 42% for GmPM1 and GmPm28, respectively. In the dehydrated state, GmPM1 and GmPM28 interact with non-reducing sugars to improve the transition temperature of cellular glass, with poly-l-lysine to prevent dehydration-induced aggregation and with phospholipids to maintain the liquid crystal phase over a wide temperature range. Our work suggests that soybean LEA IV proteins are functional in the dry state. They are one of the important components in cellular glasses and may stabilize desiccation-sensitive proteins and plasma membranes during dehydration

    Computational and Statistical Analyses of Amino Acid Usage and Physico-Chemical Properties of the Twelve Late Embryogenesis Abundant Protein Classes

    Get PDF
    Late Embryogenesis Abundant Proteins (LEAPs) are ubiquitous proteins expected to play major roles in desiccation tolerance. Little is known about their structure - function relationships because of the scarcity of 3-D structures for LEAPs. The previous building of LEAPdb, a database dedicated to LEAPs from plants and other organisms, led to the classification of 710 LEAPs into 12 non-overlapping classes with distinct properties. Using this resource, numerous physico-chemical properties of LEAPs and amino acid usage by LEAPs have been computed and statistically analyzed, revealing distinctive features for each class. This unprecedented analysis allowed a rigorous characterization of the 12 LEAP classes, which differed also in multiple structural and physico-chemical features. Although most LEAPs can be predicted as intrinsically disordered proteins, the analysis indicates that LEAP class 7 (PF03168) and probably LEAP class 11 (PF04927) are natively folded proteins. This study thus provides a detailed description of the structural properties of this protein family opening the path toward further LEAP structure - function analysis. Finally, since each LEAP class can be clearly characterized by a unique set of physico-chemical properties, this will allow development of software to predict proteins as LEAPs

    Nuclear localization of the dehydrin OpsDHN1 is determined by histidine-rich domain

    Get PDF
    The cactus OpsDHN1 dehydrin belongs to a large family of disordered and highly hydrophilic proteins known as Late Embryogenesis Abundant (LEA) proteins, which accumulate during the late stages of embryogenesis and in response to abiotic stresses. Herein, we present the in vivo OpsDHN1 subcellular localization by N-terminal GFP translational fusion; our results revealed a cytoplasmic and nuclear localization of the GFP::OpsDHN1 protein in Nicotiana benthamiana epidermal cells. In addition, dimer assembly of OpsDHN1 in planta using a Bimolecular Fluorescence Complementation (BiFC) approach was demonstrated. In order to understand the in vivo role of the histidine-rich motif, the OpsDHN1 - Delta His version was produced and assayed for its subcellular localization and dimer capability by GFP fusion and BiFC assays, respectively. We found that deletion of the OpsDHN1 histidine-rich motif restricted its localization to cytoplasm, but did not affect dimer formation. In addition, the deletion of the S-segment in the OpsDHN1 protein affected its nuclear localization. Our data suggest that the deletion of histidine-rich motif and S-segment show similar effects, preventing OpsDHN1 from getting into the nucleus. Based on these results, the histidine-rich motif is proposed as a targeting element for OpsDHN1 nuclear localization.This work was supported by the CONACYT (Investigacion Ciencia Basica CB-2013-221075) funding to JJ, NSERC Discovery Grant to SG, and funding from the Spanish MICINN/MINECO (BIO2011-23828) to AF and MICINN (BIO2011-23828) to JC. The authors acknowledge to Marisol Gascon Irian from Institut de Biologia Molecular y Celular de Plantas and Nydia Hernandez-Rios from Neurology Institute-UNAM for their technical assistance using the confocal laser-scanning microscope.Hernández-Sánchez, I.; Maruri-López, I.; Ferrando Monleón, AR.; Carbonell Gisbert, J.; Graether, S.; Jimenez-Bremont, J. (2015). Nuclear localization of the dehydrin OpsDHN1 is determined by histidine-rich domain. Frontiers in Plant Science. 6(702):1-8. https://doi.org/10.3389/fpls.2015.00702S186702Alsheikh, M. K., Heyen, B. J., & Randall, S. K. (2003). Ion Binding Properties of the Dehydrin ERD14 Are Dependent upon Phosphorylation. Journal of Biological Chemistry, 278(42), 40882-40889. doi:10.1074/jbc.m307151200Belda-Palazón, B., Ruiz, L., Martí, E., Tárraga, S., Tiburcio, A. F., Culiáñez, F., … Ferrando, A. (2012). Aminopropyltransferases Involved in Polyamine Biosynthesis Localize Preferentially in the Nucleus of Plant Cells. PLoS ONE, 7(10), e46907. doi:10.1371/journal.pone.0046907Briesemeister, S., Rahnenf�hrer, J., & Kohlbacher, O. (2010). YLoc—an interpretable web server for predicting subcellular localization. Nucleic Acids Research, 38(suppl_2), W497-W502. doi:10.1093/nar/gkq477Carjuzaa, P., Castellión, M., Distéfano, A. J., del Vas, M., & Maldonado, S. (2008). Detection and subcellular localization of dehydrin-like proteins in quinoa (Chenopodium quinoa Willd.) embryos. Protoplasma, 233(1-2), 149-156. doi:10.1007/s00709-008-0300-4Close, T. J. (1996). Dehydrins: Emergence of a biochemical role of a family of plant dehydration proteins. Physiologia Plantarum, 97(4), 795-803. doi:10.1111/j.1399-3054.1996.tb00546.xCurtis, M. D., & Grossniklaus, U. (2003). A Gateway Cloning Vector Set for High-Throughput Functional Analysis of Genes in Planta. Plant Physiology, 133(2), 462-469. doi:10.1104/pp.103.027979Ferrando, A. (2001). Detection of in vivo protein interactions between Snf1-related kinase subunits with intron-tagged epitope-labelling in plants cells. Nucleic Acids Research, 29(17), 3685-3693. doi:10.1093/nar/29.17.3685Goday, A., Jensen, A. B., Culiáñez-Macià, F. A., Mar Albà, M., Figueras, M., Serratosa, J., … Pagès, M. (1994). The maize abscisic acid-responsive protein Rab17 is located in the nucleus and interacts with nuclear localization signals. The Plant Cell, 6(3), 351-360. doi:10.1105/tpc.6.3.351Godoy, J. A., Lunar, R., Torres-Schumann, S., Moreno, J., Rodrigo, R. M., & Pintor-Toro, J. A. (1994). Expression, tissue distribution and subcellular localization of dehydrin TAS14 in salt-stressed tomato plants. Plant Molecular Biology, 26(6), 1921-1934. doi:10.1007/bf00019503Graether, S. P., & Boddington, K. F. (2014). Disorder and function: a review of the dehydrin protein family. Frontiers in Plant Science, 5. doi:10.3389/fpls.2014.00576Hanin, M., Brini, F., Ebel, C., Toda, Y., Takeda, S., & Masmoudi, K. (2011). Plant dehydrins and stress tolerance. Plant Signaling & Behavior, 6(10), 1503-1509. doi:10.4161/psb.6.10.17088Hara, M. (2010). The multifunctionality of dehydrins: An overview. Plant Signaling & Behavior, 5(5), 503-508. doi:10.4161/psb.11085Hara, M., Fujinaga, M., & Kuboi, T. (2005). Metal binding by citrus dehydrin with histidine-rich domains. Journal of Experimental Botany, 56(420), 2695-2703. doi:10.1093/jxb/eri262Hara, M., Kondo, M., & Kato, T. (2013). A KS-type dehydrin and its related domains reduce Cu-promoted radical generation and the histidine residues contribute to the radical-reducing activities. Journal of Experimental Botany, 64(6), 1615-1624. doi:10.1093/jxb/ert016HARA, M., SHINODA, Y., TANAKA, Y., & KUBOI, T. (2009). DNA binding of citrus dehydrin promoted by zinc ion. Plant, Cell & Environment, 32(5), 532-541. doi:10.1111/j.1365-3040.2009.01947.xHara, M., Terashima, S., & Kuboi, T. (2001). Characterization and cryoprotective activity of cold-responsive dehydrin from Citrus unshiu. Journal of Plant Physiology, 158(10), 1333-1339. doi:10.1078/0176-1617-00600Hernández-Sánchez, I. E., Martynowicz, D. M., Rodríguez-Hernández, A. A., Pérez-Morales, M. B., Graether, S. P., & Jiménez-Bremont, J. F. (2014). A dehydrin-dehydrin interaction: the case of SK3 from Opuntia streptacantha. Frontiers in Plant Science, 5. doi:10.3389/fpls.2014.00520Heyen, B. J., Alsheikh, M. K., Smith, E. A., Torvik, C. F., Seals, D. F., & Randall, S. K. (2002). The Calcium-Binding Activity of a Vacuole-Associated, Dehydrin-Like Protein Is Regulated by Phosphorylation. Plant Physiology, 130(2), 675-687. doi:10.1104/pp.002550Houde, M., Daniel, C., Lachapelle, M., Allard, F., Laliberte, S., & Sarhan, F. (1995). Immunolocalization of freezing-tolerance-associated proteins in the cytoplasm and nucleoplasm of wheat crown tissues. The Plant Journal, 8(4), 583-593. doi:10.1046/j.1365-313x.1995.8040583.xHwang, I. S., Choi, D. S., Kim, N. H., Kim, D. S., & Hwang, B. K. (2013). The pepper cysteine/histidine-rich DC1 domain protein CaDC1 binds both RNA and DNA and is required for plant cell death and defense response. New Phytologist, 201(2), 518-530. doi:10.1111/nph.12521Jensen, A. B., Goday, A., Figueras, M., Jessop, A. C., & Pagès, M. (1998). Phosphorylation mediates the nuclear targeting of the maize Rab17 protein. The Plant Journal, 13(5), 691-697. doi:10.1046/j.1365-313x.1998.00069.xJiménez-Bremont, J. F., Maruri-López, I., Ochoa-Alfaro, A. E., Delgado-Sánchez, P., Bravo, J., & Rodríguez-Kessler, M. (2012). LEA Gene Introns: is the Intron of Dehydrin Genes a Characteristic of the Serine-Segment? Plant Molecular Biology Reporter, 31(1), 128-140. doi:10.1007/s11105-012-0483-xKoag, M.-C., Wilkens, S., Fenton, R. D., Resnik, J., Vo, E., & Close, T. J. (2009). The K-Segment of Maize DHN1 Mediates Binding to Anionic Phospholipid Vesicles and Concomitant Structural Changes. Plant Physiology, 150(3), 1503-1514. doi:10.1104/pp.109.136697Kosugi, S., Hasebe, M., Matsumura, N., Takashima, H., Miyamoto-Sato, E., Tomita, M., & Yanagawa, H. (2008). Six Classes of Nuclear Localization Signals Specific to Different Binding Grooves of Importin α. Journal of Biological Chemistry, 284(1), 478-485. doi:10.1074/jbc.m807017200Mueller, J. K., Heckathorn, S. A., & Fernando, D. (2003). Identification of a Chloroplast Dehydrin in Leaves of Mature Plants. International Journal of Plant Sciences, 164(4), 535-542. doi:10.1086/375376Nylander, M., Svensson, J., Palva, E. T., & Welin, B. V. (2001). Plant Molecular Biology, 45(3), 263-279. doi:10.1023/a:1006469128280Ochoa-Alfaro, A. E., Rodríguez-Kessler, M., Pérez-Morales, M. B., Delgado-Sánchez, P., Cuevas-Velazquez, C. L., Gómez-Anduro, G., & Jiménez-Bremont, J. F. (2011). Functional characterization of an acidic SK3 dehydrin isolated from an Opuntia streptacantha cDNA library. Planta, 235(3), 565-578. doi:10.1007/s00425-011-1531-8Puhakainen, T., Hess, M. W., Mäkelä, P., Svensson, J., Heino, P., & Palva, E. T. (2004). Overexpression of Multiple Dehydrin Genes Enhances Tolerance to Freezing Stress in Arabidopsis. Plant Molecular Biology, 54(5), 743-753. doi:10.1023/b:plan.0000040903.66496.a4Rahman, L. N., McKay, F., Giuliani, M., Quirk, A., Moffatt, B. A., Harauz, G., & Dutcher, J. R. (2013). Interactions of Thellungiella salsuginea dehydrins TsDHN-1 and TsDHN-2 with membranes at cold and ambient temperatures—Surface morphology and single-molecule force measurements show phase separation, and reveal tertiary and quaternary associations. Biochimica et Biophysica Acta (BBA) - Biomembranes, 1828(3), 967-980. doi:10.1016/j.bbamem.2012.11.031Ricardi, M. M., Guaimas, F. F., González, R. M., Burrieza, H. P., López-Fernández, M. P., Jares-Erijman, E. A., … Iusem, N. D. (2012). Nuclear Import and Dimerization of Tomato ASR1, a Water Stress-Inducible Protein Exclusive to Plants. PLoS ONE, 7(8), e41008. doi:10.1371/journal.pone.0041008Riera, M., Figueras, M., Lopez, C., Goday, A., & Pages, M. (2004). Protein kinase CK2 modulates developmental functions of the abscisic acid responsive protein Rab17 from maize. Proceedings of the National Academy of Sciences, 101(26), 9879-9884. doi:10.1073/pnas.0306154101Rorat, T. (2006). Plant dehydrins — Tissue location, structure and function. Cellular and Molecular Biology Letters, 11(4). doi:10.2478/s11658-006-0044-0Tompa, P., & Kovacs, D. (2010). Intrinsically disordered chaperones in plants and animalsThis paper is one of a selection of papers published in this special issue entitled «Canadian Society of Biochemistry, Molecular & Cellular Biology 52nd Annual Meeting — Protein Folding: Principles and Diseases» and has undergone the Journal’s usual peer review process. Biochemistry and Cell Biology, 88(2), 167-174. doi:10.1139/o09-163Voinnet, O., Rivas, S., Mestre, P., & Baulcombe, D. (2003). Retracted: An enhanced transient expression system in plants based on suppression of gene silencing by the p19 protein of tomato bushy stunt virus. The Plant Journal, 33(5), 949-956. doi:10.1046/j.1365-313x.2003.01676.xWisniewski, M., Webb, R., Balsamo, R., Close, T. J., Yu, X.-M., & Griffith, M. (1999). Purification, immunolocalization, cryoprotective, and antifreeze activity of PCA60: A dehydrin from peach (Prunus persica). Physiologia Plantarum, 105(4), 600-608. doi:10.1034/j.1399-3054.1999.105402.xXie, C., Zhang, R., Qu, Y., Miao, Z., Zhang, Y., Shen, X., … Dong, J. (2012). Overexpression of MtCAS31 enhances drought tolerance in transgenic Arabidopsis by reducing stomatal density. New Phytologist, 195(1), 124-135. doi:10.1111/j.1469-8137.2012.04136.xXu, J., Zhang, Y. X., Wei, W., Han, L., Guan, Z. Q., Wang, Z., & Chai, T. Y. (2007). BjDHNs Confer Heavy-metal Tolerance in Plants. Molecular Biotechnology, 38(2), 91-98. doi:10.1007/s12033-007-9005-

    Hv-CBF2A overexpression in barley accelerates COR gene transcript accumulation and acquisition of freezing tolerance during cold acclimation

    Get PDF
    Abstract C-Repeat Binding Factors (CBFs) are DNAbinding transcriptional activators of gene pathways imparting freezing tolerance. Poaceae contain three CBF subfamilies, two of which, HvCBF3/CBFIII and HvCBF4/CBFIV, are unique to this taxon. To gain mechanistic insight into HvCBF4/CBFIV CBFs we overexpressed Hv-CBF2A in spring barley (Hordeum vulgare) cultivar ‘Golden Promise’. The Hv-CBF2A overexpressing lines exhibited stunted growth, poor yield, and greater freezing tolerance compared to non-transformed ‘Golden Promise’. Differences in freezing tolerance were apparent only upon cold acclimation. During cold acclimation freezing tolerance of the Hv-CBF2A overexpressing lines increased more rapidly than that of ‘Golden Promise’ and paralleled the freezing tolerance of the winter hardy barley ‘Dicktoo’. Transcript levels of candidate CBF target genes, COR14B and DHN5 were increased in the overexpressor lines at warm temperatures, and at cold temperatures they accumulated to much higher levels in the Hv-CBF2A overexpressors than in ‘Golden Promise’. Hv-CBF2A overexpression also increased transcript levels of other CBF genes at FROST RESISTANCE-H2-H2 (FR-H2) possessing CRT/DRE sites in their upstream regions, the most notable of which was CBF12. CBF12 transcript levels exhibited a relatively constant incremental increase above levels in ‘Golden Promise’ both at warm and cold. These data indicate that Hv-CBF2A activates target genes at warm temperatures and that transcript accumulation for some of these targets is greatly enhanced by cold temperatures

    Promutagenicity of 8-Chloroguanine, A Major Inflammation-Induced Halogenated DNA Lesion

    No full text
    Chronic inflammation is closely associated with cancer development. One possible mechanism for inflammation-induced carcinogenesis is DNA damage caused by reactive halogen species, such as hypochlorous acid, which is released by myeloperoxidase to kill pathogens. Hypochlorous acid can attack genomic DNA to produce 8-chloro-2′-deoxyguanosine (ClG) as a major lesion. It has been postulated that ClG promotes mutagenic replication using its syn conformer; yet, the structural basis for ClG-induced mutagenesis is unknown. We obtained crystal structures and kinetics data for nucleotide incorporation past a templating ClG using human DNA polymerase β (polβ) as a model enzyme for high-fidelity DNA polymerases. The structures showed that ClG formed base pairs with incoming dCTP and dGTP using its anti and syn conformers, respectively. Kinetic studies showed that polβ incorporated dGTP only 15-fold less efficiently than dCTP, suggesting that replication across ClG is promutagenic. Two hydrogen bonds between syn-ClG and anti-dGTP and a water-mediated hydrogen bond appeared to facilitate mutagenic replication opposite the major halogenated guanine lesion. These results suggest that ClG in DNA promotes G to C transversion mutations by forming Hoogsteen base pairing between syn-ClG and anti-G during DNA synthesis
    corecore