99 research outputs found

    Empowering Small Enterprises by Driving Value and Flow Through Systemic Strategic Planning: An exploratory study into the strategic planning of Small Enterprises in the Canadian Provinces of Ontario and Quebec

    Get PDF
    The journey takes us at the heart of strategic planning in small enterprises (SEs). We explore the role of strategy and business models in SEs and how the two are adopted by SE managers, the level of involvement of their employees, and the outcomes on the enterprise. In the increasingly fast-paced and highly competitive business landscape systemic strategic planning provides small enterprises (SEs) with valuable opportunities to understand and thoroughly examine their enterprises’ internal and external environments, and identify ways to enhance the likelihood of steering the company towards a preferred future. In this in situ research initiative, which focuses on enterprises in the Provinces of Ontario and Quebec, we seek primarily to gather a better understanding of small enterprise managers’ views on strategic planning. Secondly, we explore tools, techniques, and practices used. Thirdly, we explore the growing trend of business modeling and how it is welcomed by SEs. Our research team has uncovered a limited use or absence of systemic strategic planning amongst SEs. As a result, many SE managers have a preconceived notion that strategic planning is not for them and thus, at a cost, turn their attention to what is inherently most intuitive to them: operational planning. SE managers often focus on an isolated aspect of the process, such as sales or financial performance, rather than the integrated process, and thus fail to close the loop. Our research findings have identified a combination of key factors that impact the adoption of strategic planning. To that affect, several recommendations are proposed to the stackeholders as means of increasing the systemic adoption of strategic planning in SEs

    Family adjustments to large scale financial crisis: US and Malaysian Cases

    Full text link
    Economicshttp://deepblue.lib.umich.edu/bitstream/2027.42/85265/1/alazem.pd

    Thrombocytopenia is more Frequent in Gram negative Neonatal Septicemia

    Get PDF
    Background: Sepsis is one of the major causes of neonatal thrombocytopenia. Aim of the work: To identify the frequency, severity, and clinical outcome of thrombocytopenia associated with culture-proven neonatal septicemia in the Neonatal Intensive Care Units (NICUs) of Cairo University Children's Hospitals. Methods: We conducted a retrospective cohort study that included all neonates with culture-proven sepsis and thrombocytopenia who were admitted to the NICUs over a one-year period (from January 2017 to December 2017). Thrombocytopenia was defined as platelet count less than 150x103/µL. The thrombocytopenic neonates were divided into two groups according to the type of cultured bacteria (gram-positive and gram-negative). Both groups were compared regarding maternal and neonatal risk factors, onset and severity of thrombocytopenia, complications, and patient survival. Results: A total of 316 out of 2172 (total number of NICU admissions) newborns were found to have culture proven-sepsis (14.5%). The frequency of thrombocytopenia in neonates with culture proven-sepsis was 30.3% (n = 96/316). Prematurity is a risk factor for early onset sepsis with thrombocytopenia (p= 0.001). The frequency of severe thrombocytopenia is more in gram-negative sepsis than that in gram-positive sepsis at the onset of sepsis and at the lowest platelet count (p= 0.014, 0.015) respectively. The frequency of hemorrhage in neonates with sepsis and thrombocytopenia was 20.8 % (n = 20/96) and it was mainly pulmonary hemorrhage 10.4 % (n=10). The overall mortality among the study group was 40.6% (n=39/96), with a higher mortality rate (46.3%) in gram-negative sepsis with thrombocytopenia (OR 2.65, p= 0.042). Conclusion: Neonatal thrombocytopenia is a common finding in neonatal sepsis, and the frequency of severe thrombocytopenia is more in gram negative sepsis. Pulmonary hemorrhage is a common type of bleeding in thrombocytopenic neonates with sepsis. Gram-positive sepsis associated thrombocytopenia has a better prognosis than gram-negative sepsis

    The coat protein of Alfalfa mosaic virus interacts and interferes with the transcriptional activity of the bHLH transcription factor ILR3 promoting salicylic acid-dependent defence signalling response

    Full text link
    [EN] During virus infection, specific viral component-host factor interaction elicits the transcriptional reprogramming of diverse cellular pathways. Alfalfa mosaic virus (AMV) can establish a compatible interaction in tobacco and Arabidopsis hosts. We show that the coat protein (CP) of AMV interacts directly with transcription factor (TF) ILR3 of both species. ILR3 is a basic helix-loop-helix (bHLH) family member of TFs, previously proposed to participate in diverse metabolic pathways. ILR3 has been shown to regulate NEET in Arabidopsis, a critical protein in plant development, senescence, iron metabolism and reactive oxygen species (ROS) homeostasis. We show that the AMV CP-ILR3 interaction causes a fraction of this TF to relocate from the nucleus to the nucleolus. ROS, pathogenesis-related protein 1 (PR1) mRNAs, salicylic acid (SA) and jasmonic acid (JA) contents are increased in healthy Arabidopsis loss-of-function ILR3 mutant (ilr3.2) plants, which implicates ILR3 in the regulation of plant defence responses. In AMV-infected wild-type (wt) plants, NEET expression is reduced slightly, but is induced significantly in ilr3.2 mutant plants. Furthermore, the accumulation of SA and JA is induced in Arabidopsis wt-infected plants. AMV infection in ilr3.2 plants increases JA by over 10-fold, and SA is reduced significantly, indicating an antagonist crosstalk effect. The accumulation levels of viral RNAs are decreased significantly in ilr3.2 mutants, but the virus can still systemically invade the plant. The AMV CP-ILR3 interaction may down-regulate a host factor, NEET, leading to the activation of plant hormone responses to obtain a hormonal equilibrium state, where infection remains at a level that does not affect plant viability.F.A. was the recipient of a contract Ramon y Cajal (RYC-2010-06169) program of the Ministerio de Educacion, Cultura y Deporte of Spain. We thank L. Corachan for excellent technical assistance. This work was supported by Grants BIO2014-54862-R from the Spanish grant agency Direccion General de Investigacion Cientifica y Tecnica (DGICT) the Prometeo Program GV2015/010 from the Generalitat Valenciana and PAID-06-10-1496 from the Universitat Politecnica de Valencia (Spain).Aparicio Herrero, F.; Pallás Benet, V. (2017). The coat protein of Alfalfa mosaic virus interacts and interferes with the transcriptional activity of the bHLH transcription factor ILR3 promoting salicylic acid-dependent defence signalling response. Molecular Plant Pathology. 18(2):173-186. https://doi.org/10.1111/mpp.12388S173186182Abbink, T. E. M., Peart, J. R., Mos, T. N. M., Baulcombe, D. C., Bol, J. F., & Linthorst, H. J. M. (2002). Silencing of a Gene Encoding a Protein Component of the Oxygen-Evolving Complex of Photosystem II Enhances Virus Replication in Plants. Virology, 295(2), 307-319. doi:10.1006/viro.2002.1332Alazem, M., & Lin, N. (2014). Roles of plant hormones in the regulation of host–virus interactions. Molecular Plant Pathology, 16(5), 529-540. doi:10.1111/mpp.12204Aparicio, F., Vilar, M., Perez-Payá, E., & Pallás, V. (2003). The coat protein of prunus necrotic ringspot virus specifically binds to and regulates the conformation of its genomic RNA. Virology, 313(1), 213-223. doi:10.1016/s0042-6822(03)00284-8Aparicio, F., Thomas, C. L., Lederer, C., Niu, Y., Wang, D., & Maule, A. J. (2005). Virus Induction of Heat Shock Protein 70 Reflects a General Response to Protein Accumulation in the Plant Cytosol. Plant Physiology, 138(1), 529-536. doi:10.1104/pp.104.058958Aparicio, F., Sánchez-Navarro, J. A., & Pallás, V. (2006). In vitro and in vivo mapping of the Prunus necrotic ringspot virus coat protein C-terminal dimerization domain by bimolecular fluorescence complementation. Journal of General Virology, 87(6), 1745-1750. doi:10.1099/vir.0.81696-0Balasubramaniam, M., Kim, B.-S., Hutchens-Williams, H. M., & Loesch-Fries, L. S. (2014). The Photosystem II Oxygen-Evolving Complex Protein PsbP Interacts With the Coat Protein of Alfalfa mosaic virus and Inhibits Virus Replication. Molecular Plant-Microbe Interactions®, 27(10), 1107-1118. doi:10.1094/mpmi-02-14-0035-rBhat, S., Folimonova, S. Y., Cole, A. B., Ballard, K. D., Lei, Z., Watson, B. S., … Nelson, R. S. (2012). Influence of Host Chloroplast Proteins on Tobacco mosaic virus Accumulation and Intercellular Movement. Plant Physiology, 161(1), 134-147. doi:10.1104/pp.112.207860Bol, J. F. (2005). Replication of Alfamo- and Ilarviruses: Role of the Coat Protein. Annual Review of Phytopathology, 43(1), 39-62. doi:10.1146/annurev.phyto.43.101804.120505Callaway, A., Giesman-Cookmeyer, D., Gillock, E. T., Sit, T. L., & Lommel, S. A. (2001). THEMULTIFUNCTIONALCAPSIDPROTEINS OFPLANTRNA VIRUSES. Annual Review of Phytopathology, 39(1), 419-460. doi:10.1146/annurev.phyto.39.1.419Collum, T. D., & Culver, J. N. (2016). The impact of phytohormones on virus infection and disease. Current Opinion in Virology, 17, 25-31. doi:10.1016/j.coviro.2015.11.003Culver, J. N., & Padmanabhan, M. S. (2007). Virus-Induced Disease: Altering Host Physiology One Interaction at a Time. Annual Review of Phytopathology, 45(1), 221-243. doi:10.1146/annurev.phyto.45.062806.094422Donze, T., Qu, F., Twigg, P., & Morris, T. J. (2014). Turnip crinkle virus coat protein inhibits the basal immune response to virus invasion in Arabidopsis by binding to the NAC transcription factor TIP. Virology, 449, 207-214. doi:10.1016/j.virol.2013.11.018Fryer, M. J., Ball, L., Oxborough, K., Karpinski, S., Mullineaux, P. M., & Baker, N. R. (2003). Control of Ascorbate Peroxidase 2 expression by hydrogen peroxide and leaf water status during excess light stress reveals a functional organisation of Arabidopsis leaves. The Plant Journal, 33(4), 691-705. doi:10.1046/j.1365-313x.2003.01656.xGarcía, J. A., & Pallás, V. (2015). Viral factors involved in plant pathogenesis. Current Opinion in Virology, 11, 21-30. doi:10.1016/j.coviro.2015.01.001Heim, M. A. (2003). The Basic Helix-Loop-Helix Transcription Factor Family in Plants: A Genome-Wide Study of Protein Structure and Functional Diversity. Molecular Biology and Evolution, 20(5), 735-747. doi:10.1093/molbev/msg088Herranz, M. C., Pallas, V., & Aparicio, F. (2012). Multifunctional Roles for the N-Terminal Basic Motif of Alfalfa mosaic virus Coat Protein: Nucleolar/Cytoplasmic Shuttling, Modulation of RNA-Binding Activity, and Virion Formation. Molecular Plant-Microbe Interactions®, 25(8), 1093-1103. doi:10.1094/mpmi-04-12-0079-rHuang, Z., Yeakley, J. M., Garcia, E. W., Holdridge, J. D., Fan, J.-B., & Whitham, S. A. (2005). Salicylic Acid-Dependent Expression of Host Genes in Compatible Arabidopsis-Virus Interactions. Plant Physiology, 137(3), 1147-1159. doi:10.1104/pp.104.056028Inaba, J., Kim, B. M., Shimura, H., & Masuta, C. (2011). Virus-Induced Necrosis Is a Consequence of Direct Protein-Protein Interaction between a Viral RNA-Silencing Suppressor and a Host Catalase. Plant Physiology, 156(4), 2026-2036. doi:10.1104/pp.111.180042Jiménez, I., López, L., Alamillo, J. M., Valli, A., & García, J. A. (2006). Identification of a Plum pox virus CI-Interacting Protein from Chloroplast That Has a Negative Effect in Virus Infection. Molecular Plant-Microbe Interactions®, 19(3), 350-358. doi:10.1094/mpmi-19-0350Kim, K.-C., Lai, Z., Fan, B., & Chen, Z. (2008). Arabidopsis WRKY38 and WRKY62 Transcription Factors Interact with Histone Deacetylase 19 in Basal Defense. The Plant Cell, 20(9), 2357-2371. doi:10.1105/tpc.107.055566Kim, S. A., Punshon, T., Lanzirotti, A., Li, L., Alonso, J. M., Ecker, J. R., … Guerinot, M. L. (2006). Localization of Iron in Arabidopsis Seed Requires the Vacuolar Membrane Transporter VIT1. Science, 314(5803), 1295-1298. doi:10.1126/science.1132563Liu, Z., Zhang, Z., Faris, J. D., Oliver, R. P., Syme, R., McDonald, M. C., … Friesen, T. L. (2012). The Cysteine Rich Necrotrophic Effector SnTox1 Produced by Stagonospora nodorum Triggers Susceptibility of Wheat Lines Harboring Snn1. PLoS Pathogens, 8(1), e1002467. doi:10.1371/journal.ppat.1002467Long, T. A., Tsukagoshi, H., Busch, W., Lahner, B., Salt, D. E., & Benfey, P. N. (2010). The bHLH Transcription Factor POPEYE Regulates Response to Iron Deficiency in Arabidopsis Roots. The Plant Cell, 22(7), 2219-2236. doi:10.1105/tpc.110.074096Lukhovitskaya, N. I., Solovieva, A. D., Boddeti, S. K., Thaduri, S., Solovyev, A. G., & Savenkov, E. I. (2013). An RNA Virus-Encoded Zinc-Finger Protein Acts as a Plant Transcription Factor and Induces a Regulator of Cell Size and Proliferation in Two Tobacco Species. The Plant Cell, 25(3), 960-973. doi:10.1105/tpc.112.106476Mandadi, K. K., & Scholthof, K.-B. G. (2013). Plant Immune Responses Against Viruses: How Does a Virus Cause Disease? The Plant Cell, 25(5), 1489-1505. doi:10.1105/tpc.113.111658Maule, A. J., Escaler, M., & Aranda, M. A. (2000). Programmed responses to virus replication in plants. Molecular Plant Pathology, 1(1), 9-15. doi:10.1046/j.1364-3703.2000.00002.xNechushtai, R., Conlan, A. R., Harir, Y., Song, L., Yogev, O., Eisenberg-Domovich, Y., … Mittler, R. (2012). Characterization of Arabidopsis NEET Reveals an Ancient Role for NEET Proteins in Iron Metabolism. The Plant Cell, 24(5), 2139-2154. doi:10.1105/tpc.112.097634Nelson, B. K., Cai, X., & Nebenführ, A. (2007). A multicolored set of in vivo organelle markers for co-localization studies in Arabidopsis and other plants. The Plant Journal, 51(6), 1126-1136. doi:10.1111/j.1365-313x.2007.03212.xNemeth, K., Salchert, K., Putnoky, P., Bhalerao, R., Koncz-Kalman, Z., Stankovic-Stangeland, B., … Koncz, C. (1998). Pleiotropic control of glucose and hormone responses by PRL1, a nuclear WD protein, in Arabidopsis. Genes & Development, 12(19), 3059-3073. doi:10.1101/gad.12.19.3059Ni, P., & Cheng Kao, C. (2013). Non-encapsidation activities of the capsid proteins of positive-strand RNA viruses. Virology, 446(1-2), 123-132. doi:10.1016/j.virol.2013.07.023Olsen, A. N., Ernst, H. A., Leggio, L. L., & Skriver, K. (2005). NAC transcription factors: structurally distinct, functionally diverse. Trends in Plant Science, 10(2), 79-87. doi:10.1016/j.tplants.2004.12.010Paddock, M. L., Wiley, S. E., Axelrod, H. L., Cohen, A. E., Roy, M., Abresch, E. C., … Jennings, P. A. (2007). MitoNEET is a uniquely folded 2Fe 2S outer mitochondrial membrane protein stabilized by pioglitazone. Proceedings of the National Academy of Sciences, 104(36), 14342-14347. doi:10.1073/pnas.0707189104Pallas, V., & García, J. A. (2011). How do plant viruses induce disease? Interactions and interference with host components. Journal of General Virology, 92(12), 2691-2705. doi:10.1099/vir.0.034603-0Pallas, V., Aparicio, F., Herranz, M. C., Sanchez-Navarro, J. A., & Scott, S. W. (2013). The Molecular Biology of Ilarviruses. Advances in Virus Research, 139-181. doi:10.1016/b978-0-12-407698-3.00005-3Palukaitis, P., Groen, S. C., & Carr, J. P. (2013). The Rumsfeld paradox: some of the things we know that we don’t know about plant virus infection. Current Opinion in Plant Biology, 16(4), 513-519. doi:10.1016/j.pbi.2013.06.004Peng, X., Hu, Y., Tang, X., Zhou, P., Deng, X., Wang, H., & Guo, Z. (2012). Constitutive expression of rice WRKY30 gene increases the endogenous jasmonic acid accumulation, PR gene expression and resistance to fungal pathogens in rice. Planta, 236(5), 1485-1498. doi:10.1007/s00425-012-1698-7Pieterse, C. M. J., Van der Does, D., Zamioudis, C., Leon-Reyes, A., & Van Wees, S. C. M. (2012). Hormonal Modulation of Plant Immunity. Annual Review of Cell and Developmental Biology, 28(1), 489-521. doi:10.1146/annurev-cellbio-092910-154055Puranik, S., Sahu, P. P., Srivastava, P. S., & Prasad, M. (2012). NAC proteins: regulation and role in stress tolerance. Trends in Plant Science, 17(6), 369-381. doi:10.1016/j.tplants.2012.02.004Rampey, R. A., Woodward, A. W., Hobbs, B. N., Tierney, M. P., Lahner, B., Salt, D. E., & Bartel, B. (2006). An Arabidopsis Basic Helix-Loop-Helix Leucine Zipper Protein Modulates Metal Homeostasis and Auxin Conjugate Responsiveness. Genetics, 174(4), 1841-1857. doi:10.1534/genetics.106.061044Ren, T., Qu, F., & Morris, T. J. (2005). The nuclear localization of the Arabidopsis transcription factor TIP is blocked by its interaction with the coat protein of Turnip crinkle virus. Virology, 331(2), 316-324. doi:10.1016/j.virol.2004.10.039Rodrigo, G., Carrera, J., Ruiz-Ferrer, V., del Toro, F. J., Llave, C., Voinnet, O., & Elena, S. F. (2012). A Meta-Analysis Reveals the Commonalities and Differences in Arabidopsis thaliana Response to Different Viral Pathogens. PLoS ONE, 7(7), e40526. doi:10.1371/journal.pone.0040526Sanchez-Navarro, J., Miglino, R., Ragozzino, A., & Bol, J. F. (2001). Engineering of Alfalfa mosaic virus RNA 3 into an expression vector. Archives of Virology, 146(5), 923-939. doi:10.1007/s007050170125Sánchez-Navarro, J. A., Carmen Herranz, M., & Pallás, V. (2006). Cell-to-cell movement of Alfalfa mosaic virus can be mediated by the movement proteins of Ilar-, bromo-, cucumo-, tobamo- and comoviruses and does not require virion formation. Virology, 346(1), 66-73. doi:10.1016/j.virol.2005.10.024Selth, L. A., Dogra, S. C., Rasheed, M. S., Healy, H., Randles, J. W., & Rezaian, M. A. (2004). A NAC Domain Protein Interacts with Tomato leaf curl virus Replication Accessory Protein and Enhances Viral Replication. The Plant Cell, 17(1), 311-325. doi:10.1105/tpc.104.027235Seo, 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_7Su, L.-W., Chang, S. H., Li, M.-Y., Huang, H.-Y., Jane, W.-N., & Yang, J.-Y. (2013). Purification and biochemical characterization of Arabidopsis At-NEET, an ancient iron-sulfur protein, reveals a conserved cleavage motif for subcellular localization. Plant Science, 213, 46-54. doi:10.1016/j.plantsci.2013.09.001Taschner, P. E. M., Van Der Kuyl, A. C., Neeleman, L., & Bol, J. F. (1991). Replication of an incomplete alfalfa mosaic virus genome in plants transformed with viral replicase genes. Virology, 181(2), 445-450. doi:10.1016/0042-6822(91)90876-dThompson, J. D., Higgins, D. G., & Gibson, T. J. (1994). CLUSTAL W: improving the sensitivity of progressive multiple sequence alignment through sequence weighting, position-specific gap penalties and weight matrix choice. Nucleic Acids Research, 22(22), 4673-4680. doi:10.1093/nar/22.22.4673Toledo-Ortiz, G., Huq, E., & Quail, P. H. (2003). The Arabidopsis Basic/Helix-Loop-Helix Transcription Factor Family. The Plant Cell, 15(8), 1749-1770. doi:10.1105/tpc.013839Torres, M. A. (2010). ROS in biotic interactions. Physiologia Plantarum, 138(4), 414-429. doi:10.1111/j.1399-3054.2009.01326.xUhrig, J. F., Canto, T., Marshall, D., & MacFarlane, S. A. (2004). Relocalization of Nuclear ALY Proteins to the Cytoplasm by the Tomato Bushy Stunt Virus P19 Pathogenicity Protein. Plant Physiology, 135(4), 2411-2423. doi:10.1104/pp.104.046086Weber, P. H., & Bujarski, J. J. (2015). Multiple functions of capsid proteins in (+) stranded RNA viruses during plant–virus interactions. Virus Research, 196, 140-149. doi:10.1016/j.virusres.2014.11.014Whitham, S. A., Quan, S., Chang, H.-S., Cooper, B., Estes, B., Zhu, T., … Hou, Y.-M. (2003). Diverse RNA viruses elicit the expression of common sets of genes in susceptibleArabidopsis thalianaplants. The Plant Journal, 33(2), 271-283. doi:10.1046/j.1365-313x.2003.01625.xWhitham, S. A., Yang, C., & Goodin, M. M. (2006). Global Impact: Elucidating Plant Responses to Viral Infection. Molecular Plant-Microbe Interactions®, 19(11), 1207-1215. doi:10.1094/mpmi-19-120

    Hexanoic Acid Treatment Prevents Systemic MNSV Movement in Cucumis melo Plants by Priming Callose Deposition Correlating SA and OPDA Accumulation

    Get PDF
    [EN] Unlike fungal and bacterial diseases, no direct method is available to control viral diseases. The use of resistance-inducing compounds can be an alternative strategy for plant viruses. Here we studied the basal response of melon to Melon necrotic spot virus (MNSV) and demonstrated the efficacy of hexanoic acid (Hx) priming, which prevents the virus from systemically spreading. We analysed callose deposition and the hormonal profile and gene expression at the whole plant level. This allowed us to determine hormonal homeostasis in the melon roots, cotyledons, hypocotyls, stems and leaves involved in basal and hexanoic acid-induced resistance (Hx-IR) to MNSV. Our data indicate important roles of salicylic acid (SA), 12-oxo-phytodienoic acid (OPDA), jasmonic-isoleucine, and ferulic acid in both responses to MNSV. The hormonal and metabolites balance, depending on the time and location associated with basal and Hx-IR, demonstrated the reprogramming of plant metabolism in MNSV-inoculated plants. The treatment with both SA and OPDA prior to virus infection significantly reduced MNSV systemic movement by inducing callose deposition. This demonstrates their relevance in Hx-IR against MNSV and a high correlation with callose deposition. Our data also provide valuable evidence to unravel priming mechanisms by natural compounds.This work has been supported by grants from the Spanish Ministry of Science and Innovation (AGL2010-22300-C03-01-02, AGL2013-49023-C03-01-02-R and BIO2014-54862-R), co-funded by the European Regional Development Fund.Fernandez-Crespo, E.; Navarro Bohigues, JA.; Serra Soriano, M.; Finiti, I.; García Agustín, P.; Pallás Benet, V.; Gonzalez-Bosch, C. (2017). Hexanoic Acid Treatment Prevents Systemic MNSV Movement in Cucumis melo Plants by Priming Callose Deposition Correlating SA and OPDA Accumulation. Frontiers in Plant Science. 8:1-15. https://doi.org/10.3389/fpls.2017.01793S1158Alazem, M., & Lin, N. (2014). Roles of plant hormones in the regulation of host–virus interactions. Molecular Plant Pathology, 16(5), 529-540. doi:10.1111/mpp.12204Ando, S., Obinata, A., & Takahashi, H. (2014). WRKY70 interacting with RCY1 disease resistance protein is required for resistance to Cucumber mosaic virus in Arabidopsis thaliana. Physiological and Molecular Plant Pathology, 85, 8-14. doi:10.1016/j.pmpp.2013.11.001Anfoka, G. H. (2000). Benzo-(1,2,3)-thiadiazole-7-carbothioic acid S-methyl ester induces systemic resistance in tomato (Lycopersicon esculentum. Mill cv. Vollendung) to Cucumber mosaic virus. Crop Protection, 19(6), 401-405. doi:10.1016/s0261-2194(00)00031-4Aranega-Bou, P., de la O Leyva, M., Finiti, I., García-Agustín, P., & González-Bosch, C. (2014). Priming of plant resistance by natural compounds. Hexanoic acid as a model. Frontiers in Plant Science, 5. doi:10.3389/fpls.2014.00488Bellés, J. M., López-Gresa, M. P., Fayos, J., Pallás, V., Rodrigo, I., & Conejero, V. (2008). Induction of cinnamate 4-hydroxylase and phenylpropanoids in virus-infected cucumber and melon plants. Plant Science, 174(5), 524-533. doi:10.1016/j.plantsci.2008.02.008Bolwell, G. P., Davies, D. R., Gerrish, C., Auh, C.-K., & Murphy, T. M. (1998). Comparative Biochemistry of the Oxidative Burst Produced by Rose and French Bean Cells Reveals Two Distinct Mechanisms. Plant Physiology, 116(4), 1379-1385. doi:10.1104/pp.116.4.1379Camañes, G., Scalschi, L., Vicedo, B., González-Bosch, C., & García-Agustín, P. (2015). An untargeted global metabolomic analysis reveals the biochemical changes underlying basal resistance and priming in Solanum lycopersicum, and identifies 1-methyltryptophan as a metabolite involved in plant responses to Botrytis cinerea and Pseudomonas sy. The Plant Journal, 84(1), 125-139. doi:10.1111/tpj.12964Clarke, S. F., Guy, P. L., Burritt, D. J., & Jameson, P. E. (2002). Changes in the activities of antioxidant enzymes in response to virus infection and hormone treatment. Physiologia Plantarum, 114(2), 157-164. doi:10.1034/j.1399-3054.2002.1140201.xCollum, T. D., & Culver, J. N. (2016). The impact of phytohormones on virus infection and disease. Current Opinion in Virology, 17, 25-31. doi:10.1016/j.coviro.2015.11.003Conti, G., Rodriguez, M. C., Venturuzzi, A. L., & Asurmendi, S. (2016). Modulation of host plant immunity by Tobamovirus proteins. Annals of Botany, mcw216. doi:10.1093/aob/mcw216Culver, J. N., & Padmanabhan, M. S. (2007). Virus-Induced Disease: Altering Host Physiology One Interaction at a Time. Annual Review of Phytopathology, 45(1), 221-243. doi:10.1146/annurev.phyto.45.062806.094422Dong, C.-J., Li, L., Shang, Q.-M., Liu, X.-Y., & Zhang, Z.-G. (2014). Endogenous salicylic acid accumulation is required for chilling tolerance in cucumber (Cucumis sativus L.) seedlings. Planta, 240(4), 687-700. doi:10.1007/s00425-014-2115-1Ellinger, D., Naumann, M., Falter, C., Zwikowics, C., Jamrow, T., Manisseri, C., … Voigt, C. A. (2013). Elevated Early Callose Deposition Results in Complete Penetration Resistance to Powdery Mildew in Arabidopsis. Plant Physiology, 161(3), 1433-1444. doi:10.1104/pp.112.211011Finiti, I., de la O. Leyva, M., Vicedo, B., Gómez-Pastor, R., López-Cruz, J., García-Agustín, P., … González-Bosch, C. (2014). Hexanoic acid protects tomato plants againstBotrytis cinereaby priming defence responses and reducing oxidative stress. Molecular Plant Pathology, 15(6), 550-562. doi:10.1111/mpp.12112Flors, V., Leyva, M. de la O., Vicedo, B., Finiti, I., Real, M. D., García-Agustín, P., … González-Bosch, C. (2007). Absence of the endo-β-1,4-glucanases Cel1 and Cel2 reduces susceptibility toBotrytis cinereain tomato. The Plant Journal, 52(6), 1027-1040. doi:10.1111/j.1365-313x.2007.03299.xFlors, V., Ton, J., Van Doorn, R., Jakab, G., García-Agustín, P., & Mauch-Mani, B. (2007). Interplay between JA, SA and ABA signalling during basal and induced resistance against Pseudomonas syringae and Alternaria brassicicola. The Plant Journal, 54(1), 81-92. doi:10.1111/j.1365-313x.2007.03397.xFriedrich, L., Lawton, K., Ruess, W., Masner, P., Specker, N., Rella, M. G., … Ryals, J. (1996). A benzothiadiazole derivative induces systemic acquired resistance in tobacco. The Plant Journal, 10(1), 61-70. doi:10.1046/j.1365-313x.1996.10010061.xFurch, A. C. U., Zimmermann, M. R., Kogel, K.-H., Reichelt, M., & Mithöfer, A. (2014). Direct and individual analysis of stress-related phytohormone dispersion in the vascular system ofCucurbita maximaafter flagellin 22 treatment. New Phytologist, 201(4), 1176-1182. doi:10.1111/nph.12661García, J. A., & Pallás, V. (2015). Viral factors involved in plant pathogenesis. Current Opinion in Virology, 11, 21-30. doi:10.1016/j.coviro.2015.01.001Garcia-Marcos, A., Pacheco, R., Manzano, A., Aguilar, E., & Tenllado, F. (2013). Oxylipin Biosynthesis Genes Positively Regulate Programmed Cell Death during Compatible Infections with the Synergistic Pair Potato Virus X-Potato Virus Y and Tomato Spotted Wilt Virus. Journal of Virology, 87(10), 5769-5783. doi:10.1128/jvi.03573-12Genovés, A., Navarro, J. A., & Pallás, V. (2006). Functional analysis of the five melon necrotic spot virus genome-encoded proteins. Journal of General Virology, 87(8), 2371-2380. doi:10.1099/vir.0.81793-0Genovés, A., Navarro, J. A., & Pallás, V. (2009). A self-interacting carmovirus movement protein plays a role in binding of viral RNA during the cell-to-cell movement and shows an actin cytoskeleton dependent location in cell periphery. Virology, 395(1), 133-142. doi:10.1016/j.virol.2009.08.042Ghoshroy, S., Freedman, K., Lartey, R., & Citovsky, V. (1998). Inhibition of plant viral systemic infection by non‐toxic concentrations of cadmium. The Plant Journal, 13(5), 591-602. doi:10.1046/j.1365-313x.1998.00061.xGosalvez, B., Navarro, J. ., Lorca, A., Botella, F., Sánchez-Pina, M. ., & Pallas, V. (2003). Detection of Melon necrotic spot virus in water samples and melon plants by molecular methods. Journal of Virological Methods, 113(2), 87-93. doi:10.1016/s0166-0934(03)00224-6GOSALVEZ‐BERNAL, B., GENOVES, A., ANTONIO NAVARRO, J., PALLAS, V., & SANCHEZ‐PINA, M. A. (2008). Distribution and pathway for phloem‐dependent movement of Melon necrotic spot virus in melon plants. Molecular Plant Pathology, 9(4), 447-461. doi:10.1111/j.1364-3703.2008.00474.xHanley-Bowdoin, L., Bejarano, E. R., Robertson, D., & Mansoor, S. (2013). Geminiviruses: masters at redirecting and reprogramming plant processes. Nature Reviews Microbiology, 11(11), 777-788. doi:10.1038/nrmicro3117Hernandez, J. A., Diaz-Vivancos, P., Rubio, M., Olmos, E., Ros-Barcelo, A., & Martinez-Gomez, P. (2006). Long-term plum pox virus infection produces an oxidative stress in a susceptible apricot, Prunus armeniaca, cultivar but not in a resistant cultivar. Physiologia Plantarum, 126(1), 140-152. doi:10.1111/j.1399-3054.2005.00581.xHipper, C., Brault, V., Ziegler-Graff, V., & Revers, F. (2013). Viral and Cellular Factors Involved in Phloem Transport of Plant Viruses. Frontiers in Plant Science, 4. doi:10.3389/fpls.2013.00154Inaba, J., Kim, B. M., Shimura, H., & Masuta, C. (2011). Virus-Induced Necrosis Is a Consequence of Direct Protein-Protein Interaction between a Viral RNA-Silencing Suppressor and a Host Catalase. Plant Physiology, 156(4), 2026-2036. doi:10.1104/pp.111.180042Lange, L., & Insunza, V. (1977). Root-inhabiting Olpidium species: The O. radicale complex. Transactions of the British Mycological Society, 69(3), 377-384. doi:10.1016/s0007-1536(77)80074-0Lee, J.-Y., Wang, X., Cui, W., Sager, R., Modla, S., Czymmek, K., … Lakshmanan, V. (2011). A Plasmodesmata-Localized Protein Mediates Crosstalk between Cell-to-Cell Communication and Innate Immunity in Arabidopsis. The Plant Cell, 23(9), 3353-3373. doi:10.1105/tpc.111.087742Leyva, M. O., Vicedo, B., Finiti, I., Flors, V., Del Amo, G., Real, M. D., … González-Bosch, C. (2008). Preventive and post-infection control ofBotrytis cinereain tomato plants by hexanoic acid. Plant Pathology, 57(6), 1038-1046. doi:10.1111/j.1365-3059.2008.01891.xLi, J., Brader, G., Kariola, T., & Tapio Palva, E. (2006). WRKY70 modulates the selection of signaling pathways in plant defense. The Plant Journal, 46(3), 477-491. doi:10.1111/j.1365-313x.2006.02712.xManacorda, C. A., Mansilla, C., Debat, H. J., Zavallo, D., Sánchez, F., Ponz, F., & Asurmendi, S. (2013). Salicylic Acid Determines Differential Senescence Produced by Two Turnip mosaic virus Strains Involving Reactive Oxygen Species and Early Transcriptomic Changes. Molecular Plant-Microbe Interactions®, 26(12), 1486-1498. doi:10.1094/mpmi-07-13-0190-rMandadi, K. K., & Scholthof, K.-B. G. (2013). Plant Immune Responses Against Viruses: How Does a Virus Cause Disease? The Plant Cell, 25(5), 1489-1505. doi:10.1105/tpc.113.111658Mauch-Mani, B., & Mauch, F. (2005). The role of abscisic acid in plant–pathogen interactions. Current Opinion in Plant Biology, 8(4), 409-414. doi:10.1016/j.pbi.2005.05.015Mayers, C. N., Lee, K.-C., Moore, C. A., Wong, S.-M., & Carr, J. P. (2005). Salicylic Acid-Induced Resistance to Cucumber mosaic virus in Squash and Arabidopsis thaliana: Contrasting Mechanisms of Induction and Antiviral Action. Molecular Plant-Microbe Interactions®, 18(5), 428-434. doi:10.1094/mpmi-18-0428Mittler, R. (2017). ROS Are Good. Trends in Plant Science, 22(1), 11-19. doi:10.1016/j.tplants.2016.08.002Miura, K., & Tada, Y. (2014). Regulation of water, salinity, and cold stress responses by salicylic acid. Frontiers in Plant Science, 5. doi:10.3389/fpls.2014.00004Naumann, M., Somerville, S. C., & Voigt, C. A. (2013). Differences in early callose deposition during adapted and non-adapted powdery mildew infection of resistantArabidopsislines. Plant Signaling & Behavior, 8(6), e24408. doi:10.4161/psb.24408Navarro, J. A., Genovés, A., Climent, J., Saurí, A., Martínez-Gil, L., Mingarro, I., & Pallás, V. (2006). RNA-binding properties and membrane insertion of Melon necrotic spot virus (MNSV) double gene block movement proteins. Virology, 356(1-2), 57-67. doi:10.1016/j.virol.2006.07.040Nicaise, V. (2014). Crop immunity against viruses: outcomes and future challenges. Frontiers in Plant Science, 5. doi:10.3389/fpls.2014.00660Nieto, C., Morales, M., Orjeda, G., Clepet, C., Monfort, A., Sturbois, B., … Bendahmane, A. (2006). AneIF4Eallele confers resistance to an uncapped and non-polyadenylated RNA virus in melon. The Plant Journal, 48(3), 452-462. doi:10.1111/j.1365-313x.2006.02885.xNováková, S., Flores-Ramírez, G., Glasa, M., Danchenko, M., Fiala, R., & Skultety, L. (2015). Partially resistant Cucurbita pepo showed late onset of the Zucchini yellow mosaic virus infection due to rapid activation of defense mechanisms as compared to susceptible cultivar. Frontiers in Plant Science, 6. doi:10.3389/fpls.2015.00263Ohki, T., Akita, F., Mochizuki, T., Kanda, A., Sasaya, T., & Tsuda, S. (2010). The protruding domain of the coat protein of Melon necrotic spot virus is involved in compatibility with and transmission by the fungal vector Olpidium bornovanus. Virology, 402(1), 129-134. doi:10.1016/j.virol.2010.03.020Pacheco, R., García-Marcos, A., Manzano, A., de Lacoba, M. G., Camañes, G., García-Agustín, P., … Tenllado, F. (2012). Comparative Analysis of Transcriptomic and Hormonal Responses to Compatible and Incompatible Plant-Virus Interactions that Lead to Cell Death. Molecular Plant-Microbe Interactions®, 25(5), 709-723. doi:10.1094/mpmi-11-11-0305Padmanabhan, M. S., Shiferaw, H., & Culver, J. N. (2006). The Tobacco mosaic virus Replicase Protein Disrupts the Localization and Function of Interacting Aux/IAA Proteins. Molecular Plant-Microbe Interactions®, 19(8), 864-873. doi:10.1094/mpmi-19-0864Pallas, V., & García, J. A. (2011). How do plant viruses induce disease? Interactions and interference with host components. Journal of General Virology, 92(12), 2691-2705. doi:10.1099/vir.0.034603-0Park, S.-W., Li, W., Viehhauser, A., He, B., Kim, S., Nilsson, A. K., … Lawrence, C. B. (2013). Cyclophilin 20-3 relays a 12-oxo-phytodienoic acid signal during stress responsive regulation of cellular redox homeostasis. Proceedings of the National Academy of Sciences, 110(23), 9559-9564. doi:10.1073/pnas.1218872110Peng, H., Li, S., Wang, L., Li, Y., Li, Y., Zhang, C., & Hou, X. (2013). Turnip mosaic virus induces expression of the LRR II subfamily genes and regulates the salicylic acid signaling pathway in non-heading Chinese cabbage. Physiological and Molecular Plant Pathology, 82, 64-72. doi:10.1016/j.pmpp.2013.01.006Rodrigo, G., Carrera, J., Ruiz-Ferrer, V., del Toro, F. J., Llave, C., Voinnet, O., & Elena, S. F. (2012). A Meta-Analysis Reveals the Commonalities and Differences in Arabidopsis thaliana Response to Different Viral Pathogens. PLoS ONE, 7(7), e40526. doi:10.1371/journal.pone.0040526Rodriguez, M. C., Conti, G., Zavallo, D., Manacorda, C. A., & Asurmendi, S. (2014). TMV-Cg Coat Protein stabilizes DELLA proteins and in turn negatively modulates salicylic acid-mediated defense pathway during Arabidopsis thalianaviral infection. BMC Plant Biology, 14(1). doi:10.1186/s12870-014-0210-xScalschi, L., Sanmartín, M., Camañes, G., Troncho, P., Sánchez-Serrano, J. J., García-Agustín, P., & Vicedo, B. (2014). Silencing ofOPR3in tomato reveals the role of OPDA in callose deposition during the activation of defense responses againstBotrytis cinerea. The Plant Journal, 81(2), 304-315. doi:10.1111/tpj.12728Scalschi, L., Vicedo, B., Camañes, G., Fernandez-Crespo, E., Lapeña, L., González-Bosch, C., & García-Agustín, P. (2012). Hexanoic acid is a resistance inducer that protects tomato plants againstPseudomonas syringaeby priming the jasmonic acid and salicylic acid pathways. Molecular Plant Pathology, 14(4), 342-355. doi:10.1111/mpp.12010Serra-Soriano, M., Pallás, V., & Navarro, J. A. (2014). A model for transport of a viral membrane protein through the early secretory pathway: minimal sequence and endoplasmic reticulum lateral mobility requirements. The Plant Journal, 77(6), 863-879. doi:10.1111/tpj.12435Taheri, P., & Tarighi, S. (2010). Riboflavin induces resistance in rice against Rhizoctonia solani via jasmonate-mediated priming of phenylpropanoid pathway. Journal of Plant Physiology, 167(3), 201-208. doi:10.1016/j.jplph.2009.08.003Taheri, P., & Tarighi, S. (2011). A survey on basal resistance and riboflavin-induced defense responses of sugar beet against Rhizoctonia solani. Journal of Plant Physiology, 168(10), 1114-1122. doi:10.1016/j.jplph.2011.01.001Tamogami, S., Noge, K., Abe, M., Agrawal, G. K., & Rakwal, R. (2012). Methyl jasmonate is transported to distal leaves via vascular process metabolizing itself into JA-Ile and triggering VOCs emission as defensive metabolites. Plant Signaling & Behavior, 7(11), 1378-1381. doi:10.4161/psb.21762Ueki, S., & Citovsky, V. (2002). The systemic movement of a tobamovirus is inhibited by a cadmium-ion-induced glycine-rich protein. Nature Cell Biology, 4(7), 478-486. doi:10.1038/ncb806Vatén, A., Dettmer, J., Wu, S., Stierhof, Y.-D., Miyashima, S., Yadav, S. R., … Helariutta, Y. (2011). Callose Biosynthesis Regulates Symplastic Trafficking during Root Development. Developmental Cell, 21(6), 1144-1155. doi:10.1016/j.devcel.2011.10.006Vicedo, B., Flors, V., de la O Leyva, M., Finiti, I., Kravchuk, Z., Real, M. D., … González-Bosch, C. (2009). Hexanoic Acid-Induced Resistance Against Botrytis cinerea in Tomato Plants. Molecular Plant-Microbe Interactions®, 22(11), 1455-1465. doi:10.1094/mpmi-22-11-1455Vlot, A. C., Dempsey, D. A., & Klessig, D. F. (2009). Salicylic Acid, a Multifaceted Hormone to Combat Disease. Annual Review of Phytopathology, 47(1), 177-206. doi:10.1146/annurev.phyto.050908.135202Wang, X., Sager, R., Cui, W., Zhang, C., Lu, H., & Lee, J.-Y. (2013). Salicylic Acid Regulates Plasmodesmata Closure during Innate Immune Responses in Arabidopsis. The Plant Cell, 25(6), 2315-2329. doi:10.1105/tpc.113.110676Zhu, F., Xi, D.-H., Yuan, S., Xu, F., Zhang, D.-W., & Lin, H.-H. (2014). Salicylic Acid and Jasmonic Acid Are Essential for Systemic Resistance Against Tobacco mosaic virus in Nicotiana benthamiana. Molecular Plant-Microbe Interactions®, 27(6), 567-577. doi:10.1094/mpmi-11-13-0349-rZhu, S., Gao, F., Cao, X., Chen, M., Ye, G., Wei, C., & Li, Y. (2005). The Rice Dwarf Virus P2 Protein Interacts with ent-Kaurene Oxidases in Vivo, Leading to Reduced Biosynthesis of Gibberellins and Rice Dwarf Symptoms. Plant Physiology, 139(4), 1935-1945. doi:10.1104/pp.105.07230

    Impacting Children’s Physical and Mental Health through Kinesiology Support in Clinical Care: A Randomized Controlled Trial Protocol

    Get PDF
    Objectives To enhance the confidence of children and adolescents with medical conditions and disabilities to engage in healthy, active lifestyles. Children with medical conditions and disabilities often exhibit more sedentary lifestyles relative to peers and are at increased risk of poor health outcomes. Clinical experience suggests physical activity confidence is an important factor influencing physical activity participation. Methods This randomized controlled trial evaluates an evidence-based intervention targeting physical activity confidence among children and adolescents with medical conditions and disabilities. Potential participants, 8 to 18 years of age diagnosed with a medical condition or disability, will be screened for adequate physical activity motivation but a lack confidence. Consenting participants (n=128) will be randomized 1:1 to a 12-week in-person or virtual physical activity intervention (24 hours/week total) led by a Registered Kinesiologist or control (assessments only). The intervention will combine physical activity participation with education about physical activity knowledge, goal setting, motivation and self-management. Primary outcomes are self-reported physical activity confidence and motivation at baseline, post-intervention and three months following intervention completion. A secondary outcome will be daily physical activity minutes assessed by accelerometry. A repeated measures mixed model will be used to compare outcomes between the in-person intervention, virtual intervention, and control groups (alpha=0.05). Conclusions This trial aims to assess the impact of a novel application of behaviour change theory on physical activity confidence among children and adolescents living with medical conditions or disabilities. Increased physical activity confidence, knowledge and skills could enable these youth to lead a more active lifestyle

    Le Canada, terre d’accueil?

    No full text
    corecore