首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到10条相似文献,搜索用时 187 毫秒
1.
Tenhaken R  Rubel C 《Plant physiology》1997,115(1):291-298
The function of salicylic acid (SA) in hypersensitive cell death was studied in a soybean (Glycine max)-Pseudomonas syringae pv glycinea system. The infection of cell cultures with bacteria leads to a hypersensitive reaction (HR), which is dependent on an appropriate avirulence gene and on low concentrations of SA. The requirement for SA is essential for a process shortly before the onset of the HR-caused cell death 5 to 6 h after infection with bacteria. SA given 10 to 12 h after infection or preincubation cannot rescue the completion of the cell death program. SA does not inhibit catalase or ascorbate peroxidase in soybean. In addition, the in vivo capacity of the cell culture for the rapid metabolism of H2O2 is not altered by SA. This clearly shows that SA is needed for the HR-caused cell death for a reaction downstream of the oxidative burst. Lipid peroxides accumulate during the HR, but the loss of membrane control precedes the generation of lipid peroxides. The accumulation of lipid peroxides in the HR can be prevented by lipid antioxidants. Nevertheless, cell death kinetics remain unaltered in the presence of antioxidants. It is concluded that lipid peroxides are a consequence of cell death, but not the primary cause of it.  相似文献   

2.
Previous studies suggest that salicylic acid (SA) plays an important role in influencing plant resistance to ozone (O3). To further define the role of SA in O3-induced responses, we compared the responses of two Arabidopsis genotypes that accumulate different amounts of SA in response to O3 and a SA-deficient transgenic Col-0 line expressing salicylate hydroxylase (NahG). The differences observed in O3-induced changes in SA levels, the accumulation of active oxygen species, defense gene expression, and the kinetics and severity of lesion formation indicate that SA influences O3 tolerance via two distinct mechanisms. Detailed analyses indicated that features associated with a hypersensitive response (HR) were significantly greater in O3-exposed Cvi-0 than in Col-0, and that NahG plants failed to exhibit these HR-like responses. Furthermore, O3-induced antioxidant defenses, including the redox state of glutathione, were greatly reduced in NahG plants compared to Col-0 and Cvi-0. This suggests that O3-induced cell death in NahG plants is due to the loss of SA-mediated potentiation of antioxidant defenses, while O3-induced cell death in Cvi-0 is due to activation of a HR. This hypothesis is supported by the observation that inhibition of NADPH-oxidases reduced O3-induced H2O2 levels and the O3-induced cell death in Cvi-0, while no major changes were observed in NahG plants. We conclude that although SA is required to maintain the cellular redox state and potentiate defense responses in O3 exposed plants, high levels of SA also potentiate activation of an oxidative burst and a cell death pathway that results in apparent O3 sensitivity.  相似文献   

3.
Controlled cellular suicide is an important process that can be observed in various organs during plant development. From the generation of proper sexual organs in monoecious plants to the hypersensitive response (HR) that occurs during incompatible pathogen interactions, programmed cell death (PCD) can be readily observed. Although several biochemical and morphological parameters have been described for various types of cell death in plants, the relationships existing between those different types of PCD events remain unclear. In this work, we set out to examine if two early molecular markers of HR cell death (HIN1 and HSR203J) as well as a senescence marker (SAG12) are coordinately induced during these processes. Our result indicates that although there is evidence of some cross-talk between both cell death pathways, spatial and temporal characteristics of activation for these markers during hypersensitive response and senescence are distinct. These observations indicate that these markers are relatively specific for different cell death programs. Interestingly, they also revealed that a senescence-like process seems to be triggered at the periphery of the HR necrotic lesion. This suggests that cells committed to die during the HR might release a signal able to induce senescence in the neighboring cells. This phenomenon could correspond to the establishment of a second barrier against pathogens. Lastly, we used those cell death markers to better characterize cell death induced by copper and we showed that this abiotic induced cell death presents similarities with HR cell death.  相似文献   

4.
Systemic acquired resistance (SAR) is an inducible plant defense response in which a prior foliar pathogen infection activates resistance in noninfected foliar tissues. Salicylic acid (SA) accumulation is essential for the establishment of SAR. While SA is probably not the long‐distance systemic signal instrumental for SAR activation, it is required for transduction of the signal in noninfected tissues. Although SAR was first described as a response to necrogenic pathogen infection, synthetic chemicals have been identified that effectively activate SAR. Elucidation of SAR signal transduction has been facilitated by the identification and characterization of Arabidopsis mutants. Disease lesion mimic mutants exhibit constitutive SAR as well as spontaneous lesion formation similar to pathogen‐associated hypersensitive cell death. Some disease lesion mimic mutants do not exhibit a lesioned phenotype when SA accumulation is prevented, thereby providing evidence for a feedback loop in SAR signal transduction. Moreover, characterization of mutants compromised for SAR activation has provided additional evidence for common signaling components between SAR and gene‐for‐gene resistance.  相似文献   

5.
Soybean cell cultures (cv. Williams 82) respond to Pseudomonas syringae bacteria expressing the avirulence gene AvrA with a hypersensitive reaction, a programmed cell death (PCD) of plant cells to pathogen attack. This PCD is under control of salicylic acid (SA) via an unknown mechanism. In the presence of low concentrations of SA, the cells undergo a very rapid cell death, which needs only half of the time required for the normal hypersensitive reaction (HR). Northern blot studies for defence-related genes show that the expression of many of these genes is tightly linked to the status of the cell death program rather than to pathogen-derived elicitors. Thus the expression is much faster in the SA-accelerated PCD than in the normal hypersensitive reaction. In contrast, other pathogen-responsive genes are induced independently of the speed of PCD, indicating a divergent signalling mechanism. The production of reactive oxygen species during the oxidative burst of bacteria-inoculated soybean cells is slightly enhanced in the presence of SA but occurs at the same time as in untreated cells, suggesting that SA exhibits the control of the PCD downstream of the oxidative burst. Consistent with these findings a HR-specific marker gene is neither directly induced by H2O2 or SA. However, this gene shows a high expression in the regular HR and is induced much faster in the SA-accelerated PCD.  相似文献   

6.
7.
H(2)O(2) from the oxidative burst, cell death, and defense responses such as the production of phenylalanine ammonia lyase (PAL), salicylic acid (SA), and scopoletin were analyzed in cultured tobacco (Nicotiana tabacum) cells treated with three proteinaceous elicitors: two elicitins (alpha-megaspermin and beta-megaspermin) and one glycoprotein. These three proteins have been isolated from Phytophthora megasperma H20 and have been previously shown to be equally efficient in inducing a hypersensitive response (HR) upon infiltration into tobacco leaves. However, in cultured tobacco cells these elicitors exhibited strikingly different biological activities. beta-Megaspermin was the only elicitor that caused cell death and induced a strong, biphasic H(2)O(2) burst. Both elicitins stimulated PAL activity similarly and strongly, while the glycoprotein caused only a slight increase. Only elicitins induced SA accumulation and scopoletin consumption, and beta-megaspermin was more efficient. To assess the role of H(2)O(2) in HR cell death and defense response expression in elicitin-treated cells, a gain and loss of function strategy was used. Our results indicated that H(2)O(2) was neither necessary nor sufficient for HR cell death, PAL activation, or SA accumulation, and that extracellular H(2)O(2) was not a direct cause of intracellular scopoletin consumption.  相似文献   

8.
To investigate the role of salicylic acid (SA) in the hypersensitive response (HR) its accumulation was compromised during different phases of lesion development by differential expression of a salicylate hydroxylase gene (SH-L). Constitutive suppression of SA accumulation was achieved by expression of a gene fusion between the CaMV35S promoter (35S) and SH-L. Using the H2O2-responsive AoPR1 promoter to drive SH-L SA accumulation could be compromised at an early stage, on lesion formation and possibly prior to visible necrosis, whilst use of the salicylate-responsive PR1a promoter reduced SA accumulation at a later stage as lesions expand. TMV infection of 35S-SH-L and AoPR1-SH-L, but not PR1a-SH-L, tobacco resulted in significantly greater rates of lesion growth than in wild-type tobacco. TMV was detected in asymptomatic tissue surrounding lesions only in 35S-SH-L and AoPR1-SH-L lines; subsequently these transgenic lines exhibited a ‘spreading-necrosis’ originating from the lesion which entered the stem and eventually other leaves, a phenotype which could be correlated with the presence of TMV particles. Analysis of TMV-infected and ‘temperature-shifted’ tobacco indicated that both 35S-SH-L and AoPR1-SH-L, but not PR1a-SH-L, transgenics exhibited delayed cell-death compared to wild-type infections. We propose that the SH-L phenotypes indicate that early SA accumulation is a major factor in preventing viral escape, via mechanism(s) which may include influencing the rate of host-cell death and, possibly, an effect on viral function.  相似文献   

9.
Nitric oxide (NO) acts as a signaling molecule in many cellular responses in plants and animals. Oat plants (Avena sativa L.) evoke the hypersensitive response (HR), which shares morphological and biochemical features with mammalian apoptosis, such as DNA laddering and heterochromatin condensation, in response to the avirulent crown rust fungus (Puccinia coronata f. sp. avenae). We examined the role of NO and reactive oxygen species (ROS) in the initiation of hypersensitive cell death, which is induced by direct contact with the pathogen, and apoptotic cell death in the adjacent cells. Cytofluorimetric analysis using the fluorescent NO probe DAF and the H2O2 probe DCF demonstrated that NO and H2O2 were generated simultaneously in primary leaves at an early stage of the defense response. The NO scavenger 2-(4-carboxyphenyl)-4,4,5,5-tetramethylimidazoline-1-oxyl-3-oxide (cPTIO) markedly enhanced H2O2 accumulation detected by 3,3-diaminobenzidine staining and DCF, whereas treatment with the NO donor S-nitroso-N-acetylpenicillamine (SNAP) strongly suppressed it. Superoxide dismutase (SOD) increased NO accumulation, suggesting that endogenous NO may modulate the level of H2O2 by interacting with O2- in the HR lesion. Cytological observation showed that administration of cPTIO, SNAP, or SOD had no effect on elicitation of hypersensitive cell death, but clearly reduced heterochromatin condensation in the nearby cells and DNA laddering. These findings indicate that NO and ROS are not essential mediators for the initiation of hypersensitive cell death. However, NO and O2- but not H2O2 are required for the onset of apoptotic cell death in the adjacent cells, where excess NO may exert its anti-apoptotic function by regulating cellular redox state.  相似文献   

10.
A novel Arabidopsis mutant has been identified with constitutive expression of GST1-GUS using plants with a pathogen-responsive reporter transgene containing the beta-glucuronidase (GUS) coding region driven by the GST1 promoter. The recessive mutant, called agd2 (aberrant growth and death2), has salicylic acid (SA)-dependent increased resistance to virulent and avirulent strains of the bacterial pathogen Pseudomonas syringae, elevated SA levels, a low level of spontaneous cell death, callose deposition, and enlarged cells in leaves. The enhanced resistance of agd2 to virulent P. syringae requires the SA signaling component NONEXPRESSOR OF PR1 (NPR1). However, agd2 renders the resistance response to P. syringae carrying avrRpt2 NPR1-independent. Thus agd2 affects both an SA- and NPR1-dependent general defense pathway and an SA-dependent, NPR1-independent pathway that is active during the recognition of avirulent P. syringae. agd2 plants also fail to show a hypersensitive cell death response (HR) unless NPR1 is removed. This novel function for NPR1 is also apparent in otherwise wild-type plants: npr1 mutants show a stronger HR, while NPR1-overproducing plants show a weaker HR when infected with P. syringae carrying the avrRpm1 gene. Spontaneous cell death in agd2 is partially suppressed by npr1, indicating that NPR1 can suppress or enhance cell death depending on the cellular context. agd2 plants depleted of SA show a dramatic exacerbation of the cell-growth phenotype and increased callose deposition, suggesting a role for SA in regulating growth and this cell-wall modification. AGD2 may function in cell death and/or growth control as well as the defense response, similarly to what has been described in animals for the functions of NFkappaB.  相似文献   

设为首页 | 免责声明 | 关于勤云 | 加入收藏

Copyright©北京勤云科技发展有限公司  京ICP备09084417号