首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 15 毫秒
1.
Sister chromatid cohesion, which is mediated by the cohesin complex, is essential for the proper segregation of chromosomes during mitosis and meiosis. Stable binding of cohesin with chromosomes is regulated in part by the opposing actions of CTF7 (CHROMOSOME TRANSMISSION FIDELITY7) and WAPL (WINGS APART-LIKE). In this study, we characterized the interaction between Arabidopsis thaliana CTF7 and WAPL by conducting a detailed analysis of wapl1-1 wapl2 ctf7 plants. ctf7 plants exhibit major defects in vegetative growth and development and are completely sterile. Inactivation of WAPL restores normal growth, mitosis, and some fertility to ctf7 plants. This shows that the CTF7/WAPL cohesin system is not essential for mitosis in vegetative cells and suggests that plants may contain a second mechanism to regulate mitotic cohesin. WAPL inactivation restores cohesin binding and suppresses ctf7-associated meiotic cohesion defects, demonstrating that WAPL and CTF7 function as antagonists to regulate meiotic sister chromatid cohesion. The ctf7 mutation only had a minor effect on wapl-associated defects in chromosome condensation and centromere association. These results demonstrate that WAPL has additional roles that are independent of its role in regulating chromatin-bound cohesin.  相似文献   

2.
Plants defend themselves against microbial pathogens through a range of highly sophisticated and integrated molecular systems. Recognition of pathogen-secreted effector proteins often triggers the hypersensitive response (HR), a complex multicellular defense reaction where programmed cell death of cells surrounding the primary site of infection is a prominent feature. Even though the HR was described almost a century ago, cell-to-cell factors acting at the local level generating the full defense reaction have remained obscure. In this study, we sought to identify diffusible molecules produced during the HR that could induce cell death in naive tissue. We found that 4-methylsulfinylbutyl isothiocyanate (sulforaphane) is released by Arabidopsis (Arabidopsis thaliana) leaf tissue undergoing the HR and that this compound induces cell death as well as primes defense in naive tissue. Two different mutants impaired in the pathogen-induced accumulation of sulforaphane displayed attenuated programmed cell death upon bacterial and oomycete effector recognition as well as decreased resistance to several isolates of the plant pathogen Hyaloperonospora arabidopsidis. Treatment with sulforaphane provided protection against a virulent H. arabidopsidis isolate. Glucosinolate breakdown products are recognized as antifeeding compounds toward insects and recently also as intracellular signaling and bacteriostatic molecules in Arabidopsis. The data presented here indicate that these compounds also trigger local defense responses in Arabidopsis tissue.Plants are constantly challenged by pathogenic microorganisms and have developed several detection and defense systems to protect themselves against the invaders. Preformed defenses include the waxy cuticle, thick cell walls, and antimicrobial compounds. After recognition of microbe-associated patterns, defense responses are induced, which include the fortification of cell walls and the production of phytoalexins (Monaghan and Zipfel, 2012). Overcoming the preformed and induced defenses of the plant hosts requires adaptation by the pathogen. Pathogenic bacteria use type III secretion to inject proteins (so-called effectors) into the host cytosol in order to overcome plant defense responses (Bent and Mackey, 2007). In turn, plants have developed systems to recognize the pathogenic effectors and mount defense. Recognition of type III effectors by plant resistance (R) proteins induces robust defense responses that frequently include the hypersensitive response (HR).The HR is a complex defense reaction characterized by the induction of programmed cell death (PCD) in the local host tissue as well as the activation of other defense responses in both local and systemic tissue (Mur et al., 2008; Shah, 2009). Oomycetes and true fungi also secrete proteinaceous effectors that can be recognized by host R proteins (Coates and Beynon, 2010; Hückelhoven and Panstruga, 2011; Feng and Zhou, 2012). The lesions formed during the HR vary in size between different host-pathogen pairs; however, a lesion induced at one or a few cells can spread to surrounding cells (Mur et al., 2008). Since pathogens inducing HR typically fail to proliferate, the first infected cell likely releases a compound that promotes PCD in surrounding cells. This is especially clear in models with oomycete and fungal pathogens, where the localization of the pathogen and the spread of cell death around the infection site can be clearly visualized (Mur et al., 2008; Coates and Beynon, 2010). Trailing necrosis is an incomplete resistance phenotype characterized by cell death that trails, but fails to contain, the filamentous growth of the pathogen. One explanation for trailing necrosis is a failure of infected cells to produce a putative mobile defense signal required to enhance defense in neighboring cells. Farther from the site of PCD, other defense pathways are activated and systemic tissue is primed for defense.The hunt for systemically acting compounds has been intense, and several candidates for this signal have been presented (Dempsey and Klessig, 2012). In contrast, even though the phenomenon of HR as a defense reaction was described almost a century ago (Stakman, 1915; Mur et al., 2008), compounds acting on the local tissue scale of the HR have attracted little attention. We set out to find substances released from cells undergoing the HR that could induce cell death in naive tissue. We report that leaf tissue of the model plant Arabidopsis (Arabidopsis thaliana) releases the reactive electrophilic compound sulforaphane after bacterial effector recognition. Mutants affected in sulforaphane production as well as other glucosinolate breakdown products showed delayed or reduced cell death after the recognition of pathogenic effectors and decreased resistance to an oomycete pathogen. Moreover, pretreatment of plants with sulforaphane enhanced resistance against a virulent oomycete isolate. Thus, we interpret this as that sulforaphane and likely similar compounds might both possess direct antimicrobial properties and, through a cytotoxic mechanism, act directly on plant cells to trigger defense responses.  相似文献   

3.
4.
Mcl-1 is a member of the Bcl2-related protein family that is a critical mediator of cell survival. Exposure of cells to stress causes inhibition of Mcl-1 mRNA translation and rapid destruction of Mcl-1 protein by proteasomal degradation mediated by a phosphodegron created by glycogen synthase kinase 3 (GSK3) phosphorylation of Mcl-1. Here we demonstrate that prior phosphorylation of Mcl-1 by the c-Jun N-terminal protein kinase (JNK) is essential for Mcl-1 phosphorylation by GSK3. Stress-induced Mcl-1 degradation therefore requires the coordinated activity of JNK and GSK3. Together, these data establish that Mcl-1 functions as a site of signal integration between the proapoptotic activity of JNK and the prosurvival activity of the AKT pathway that inhibits GSK3.Mcl-1 is an antiapoptotic member of the Bcl2 family. Gene knockout studies of mice demonstrate that Mcl-1 is essential for embryonic development and for the survival of hematopoietic cells (28-30). Studies of the stress response have demonstrated that Mcl-1 plays an important role in the sensitization of cells to apoptotic signals (1, 11, 25). Thus, exposure to UV radiation causes the rapid degradation of Mcl-1 and the release of proapoptotic partner proteins from Mcl-1 complexes (e.g., Bim). The mechanism of rapid Mcl-1 destruction is mediated by the combined actions of two different pathways. First, the exposure to stress causes phosphorylation of the alpha subunit of eukaryotic initiation factor 2 (eIF-2α) on the inhibitory site Ser-51 that prevents translation of Mcl-1 mRNA (1, 11, 25). Second, Mcl-1 is rapidly degraded by the ubiquitin-dependent proteasome pathway (27). Together, these pathways cause a rapid reduction in Mcl-1 expression. This loss of Mcl-1 may be a required initial response for the apoptosis of cells exposed to stress (25).The E3 ubiquitin protein ligase Mule/ARF-BP1 contains a BH3 domain that interacts with Mcl-1 and can initiate ubiquitin-dependent degradation of Mcl-1 (39). Recent studies have demonstrated that rapid stress-induced degradation of Mcl-1 is mediated by an alternative pathway involving the E3 ubiquitin protein ligase β-TrCP, which binds a stress-induced phosphodegron created by the phosphorylation of Mcl-1 by glycogen synthase kinase 3 (GSK3) (7, 21). How the exposure to stress causes GSK3-mediated phosphorylation of Mcl-1 is unclear, but GSK3 has been shown to directly phosphorylate Mcl-1 (7, 21). Mcl-1 phosphorylation and degradation may therefore be controlled by the prosurvival AKT pathway, which can negatively regulate GSK3 (7, 21).Mcl-1 is critically involved in the regulation of cell survival and is therefore subject to regulation by multiple mechanisms (26). Thus, Mcl-1 gene expression is regulated by many growth factors and cytokines (26), and Mcl-1 mRNA is regulated by microRNA pathways (24). The Mcl-1 protein is stabilized by binding TCTP (20) and the BH3-only protein Bim (4). In contrast, the BH3-only protein Noxa binds and destabilizes Mcl-1 (4, 36). Moreover, it is established that Mcl-1 is phosphorylated by several protein kinases on sites that may regulate Mcl-1 function. Phosphorylation of human Mcl-1 (hMcl-1) on Ser-64 (a site that is not conserved in other species) may enhance antiapoptotic activity by increasing the interaction of Mcl-1 with Bim, Noxa, and Bak (18). Phosphorylation on Ser-121 and Thr-163 may inhibit the antiapoptotic activity of hMcl-1 (15), and phosphorylation on Thr-163 may increase hMcl-1 protein stability (9). The conserved GSK3 phosphorylation site Ser-159 (and possibly Ser-155) can initiate rapid proteasomal degradation of hMcl-1 (7, 21). Together, these findings suggest that the function of Mcl-1 is very tightly regulated.The results of previous studies have implicated the c-Jun N-terminal protein kinase (JNK) in the regulation of Mcl-1 (15, 18). The purpose of this study was to test whether Mcl-1 is a target of signal transduction by JNK. We demonstrate that a key function of JNK is to prime Mcl-1 for phosphorylation by GSK3. JNK is required for GSK3-mediated degradation of Mcl-1 in response to stress. Coordinated regulation of the stress-activated JNK pathway and the AKT-inhibited GSK3 pathway is therefore required for stress-induced Mcl-1 degradation.  相似文献   

5.
将拟南芥BAK1基因采用Gateway方法连接到植物表达载体,通过侵花粉管进行转化,从基因和蛋白表达水平检测转化是否成功。以不同BAK1表达水平植株作为试验材料,分析BAK1在芜菁缩叶病毒(Turnip crinkle virus,TCV)-拟南芥(Col-0)亲和互作系统中对植株防御的影响。结果显示,在接种TCV后,BAK1缺陷型植株对TCV较为感病,衰老相关基因表达水平增加,表明BAK1能够增强宿主对病毒的防御作用。  相似文献   

6.
7.
A novel protein elicitor (PaNie(234)) from Pythium aphanidermatum (Edson) Fitzp. was purified, microsequenced, and the corresponding cDNA was cloned. The deduced amino acid sequence contains a putative eukaryotic secretion signal with a proteinase cleavage site. The heterologously expressed elicitor protein without the secretion signal of 21 amino acids (PaNie(213)) triggered programmed cell death and de novo formation of 4-hydroxybenzoic acid in cultured cells of carrot (Daucus carota). Programmed cell death was determined using the tetrazolium assay and DNA laddering. Infiltration of PaNie(213) into the intercellular space of leaves of Arabidopsis (Columbia-0, wild type) resulted in necroses and deposition of callose on the cell walls of spongy parenchyma cells surrounding the necrotic mesophyll cells. Necroses were also formed in tobacco (Nicotiana tabacum cv Wisconsin W38, wild type) and tomato (Lycopersicon esculentum Mill.) but not in maize (Zea mays), oat (Avena sativa), and Tradescantia zebrina (Bosse), indicating that monocotyledonous plants are unable to perceive the signal. The reactions observed after treatment with the purified PaNie(213) were identical to responses measured after treatment with a crude elicitor preparation from the culture medium of P. aphanidermatum, described previously. The availability of the pure protein offers the possibility to isolate the corresponding receptor and its connection to downstream signaling-inducing defense reactions.  相似文献   

8.
Action spectra were measured for suppression of germination of Poa pratensis L. and Amaranthus arenicola I. M. Johnston seed under prolonged or continuous irradiation. The action maxima for both types of seeds are near 720 nm. The maxima are unchanged in position or magnitude in the presence of radiation in the region of 600 to 670 nm adequate to maintain phytochrome predominantly in the far-red-absorbing form. A reversible potentiation of germination to change in form of phytochrome was observed for both seeds. The bearing of these findings on a high-energy regulatory light response is discussed.  相似文献   

9.

Background

The ventral horn is a major substrate in mediating the immobilizing properties of the volatile anesthetic sevoflurane in the spinal cord. In this neuronal network, action potential firing is controlled by GABAA and glycine receptors. Both types of ion channels are sensitive to volatile anesthetics, but their role in mediating anesthetic-induced inhibition of spinal locomotor networks is not fully understood.

Methodology/Principal Findings

To compare the effects of sevoflurane on GABAergic and glycinergic inhibitory postsynaptic currents (IPSCs) whole-cell voltage-clamp recordings from ventral horn interneurons were carried out in organotypic spinal cultures. At concentrations close to MAC (minimum alveolar concentration), decay times of both types of IPSCs were significantly prolonged. However, at 1.5 MAC equivalents, GABAergic IPSCs were decreased in amplitude and reduced in frequency. These effects counteracted the prolongation of the decay time, thereby decreasing the time-averaged GABAergic inhibition. In contrast, amplitudes and frequency of glycinergic IPSCs were not significantly altered by sevoflurane. Furthermore, selective GABAA and glycine receptor antagonists were tested for their potency to reverse sevoflurane-induced inhibition of spontaneous action potential firing in the ventral horn. These experiments confirmed a weak impact of GABAA receptors and a prominent role of glycine receptors at a high sevoflurane concentration.

Conclusions

At high concentrations, sevoflurane mediates neuronal inhibition in the spinal ventral horn primarily via glycine receptors, and less via GABAA receptors. Our results support the hypothesis that the impact of GABAA receptors in mediating the immobilizing properties of volatile anesthetics is less essential in comparison to glycine receptors.  相似文献   

10.
Sphingolipids are emerging as second messengers in programmed cell death and plant defense mechanisms. However, their role in plant defense is far from being understood, especially against necrotrophic pathogens. Sphingolipidomics and plant defense responses during pathogenic infection were evaluated in the mutant of long-chain base phosphate (LCB-P) lyase, encoded by the dihydrosphingosine-1-phosphate lyase1 (AtDPL1) gene and regulating long-chain base/LCB-P homeostasis. Atdpl1 mutants exhibit tolerance to the necrotrophic fungus Botrytis cinerea but susceptibility to the hemibiotrophic bacterium Pseudomonas syringae pv tomato (Pst). Here, a direct comparison of sphingolipid profiles in Arabidopsis (Arabidopsis thaliana) during infection with pathogens differing in lifestyles is described. In contrast to long-chain bases (dihydrosphingosine [d18:0] and 4,8-sphingadienine [d18:2]), hydroxyceramide and LCB-P (phytosphingosine-1-phosphate [t18:0-P] and 4-hydroxy-8-sphingenine-1-phosphate [t18:1-P]) levels are higher in Atdpl1-1 than in wild-type plants in response to B. cinerea. Following Pst infection, t18:0-P accumulates more strongly in Atdpl1-1 than in wild-type plants. Moreover, d18:0 and t18:0-P appear as key players in Pst- and B. cinerea-induced cell death and reactive oxygen species accumulation. Salicylic acid levels are similar in both types of plants, independent of the pathogen. In addition, salicylic acid-dependent gene expression is similar in both types of B. cinerea-infected plants but is repressed in Atdpl1-1 after treatment with Pst. Infection with both pathogens triggers higher jasmonic acid, jasmonoyl-isoleucine accumulation, and jasmonic acid-dependent gene expression in Atdpl1-1 mutants. Our results demonstrate that sphingolipids play an important role in plant defense, especially toward necrotrophic pathogens, and highlight a novel connection between the jasmonate signaling pathway, cell death, and sphingolipids.Plants have evolved a complex array of defenses when attacked by microbial pathogens. The success of plant resistance first relies on the capacity of the plant to recognize its invader. Among early events, a transient production of reactive oxygen species (ROS), known as the oxidative burst, is characteristic of successful pathogen recognition (Torres, 2010). Perception of pathogen attack then initiates a large array of immune responses, including modification of cell walls, as well as the production of antimicrobial proteins and metabolites like pathogenesis-related (PR) proteins and phytoalexins, respectively (Schwessinger and Ronald, 2012). The plant hormones salicylic acid (SA), jasmonic acid (JA), and ethylene (ET) are key players in the signaling networks involved in plant resistance (Bari and Jones, 2009; Tsuda and Katagiri, 2010; Robert-Seilaniantz et al., 2011). Interactions between these signal molecules allow the plant to activate and/or modulate an appropriate array of defense responses, depending on the pathogen lifestyle, necrotroph or biotroph (Glazebrook, 2005; Koornneef and Pieterse, 2008). Whereas SA is considered essential for resistance to (hemi)biotrophic pathogens, it is assumed that JA and ET signaling pathways are important for resistance to necrotrophic pathogens in Arabidopsis (Arabidopsis thaliana; Thomma et al., 2001; Glazebrook, 2005). A successful innate immune response often includes the so-called hypersensitive response (HR), a form of rapid programmed cell death (PCD) occurring in a limited area at the site of infection. This suicide of infected cells is thought to limit the spread of biotrophic pathogens, including viruses, bacteria, fungi, and oomycetes (Mur et al., 2008).During the past decade, significant progress has been made in our understanding of the cellular function of plant sphingolipids. Besides being structural components of cell membranes, sphingolipids are bioactive metabolites that regulate important cellular processes such as cell survival and PCD, occurring during either plant development or plant defense (Dunn et al., 2004; Berkey et al., 2012; Markham et al., 2013). The first evidence of the role of sphingolipids in these processes came from the use of the fungal toxins fumonisin B1 (FB1) and AAL, produced by the necrotrophic agent Alternaria alternata f. sp. lycopersici. These toxins are structural sphingosine (d18:1) analogs and function as ceramide synthase inhibitors. They triggered PCD when exogenously applied to plants. Mutant strains in which the production of such toxins is abrogated failed to infect the host plant, implying that toxin accumulation is required for pathogenicity and that the induction of plant PCD could be considered a virulence tool used by necrotrophic pathogens (Berkey et al., 2012). Moreover, several studies revealed that ceramides (Cers) and long-chain bases (LCBs) are also potent inducers of PCD in plants. For example, exogenously applied Cers and LCBs (d18:0, d18:1, or t18:0) induced PCD either in cell suspension cultures (Liang et al., 2003; Lachaud et al., 2010, 2011; Alden et al., 2011) or in whole seedlings (Shi et al., 2007; Takahashi et al., 2009; Saucedo-García et al., 2011). AAL- and FB1-induced PCD seemed to be due to the accumulation of free sphingoid bases (dihydrosphingosine [d18:0] and phytosphingosine [t18:0]; Abbas et al., 1994; Brandwagt et al., 2000; Shi et al., 2007). Spontaneous cell death in lag one homolog1 or l-myoinositol1-phosphate synthase mutant could be due to trihydroxy-LCB and/or Cer accumulation (Donahue et al., 2010; Ternes et al., 2011). Deciphering of Cer participation in the induction of HR and associated PCD also came from studies on accelerated cell death5 (acd5) and enhancing resistance to powdery mildew8 (RPW8)-mediated hypersensitive response (erh1) mutants, which displayed overaccumulation of Cers. These mutants exhibited spontaneous cell death and resistance to biotrophic pathogens, which seemed to be linked with SA and PR protein accumulation (Liang et al., 2003; Wang et al., 2008).Altogether, these data provide evidence of a link between PCD, defense, and sphingolipid metabolism. However, the fatty acid hydroxylase1/2 (atfah1/atfah2) double mutant that accumulates SA and Cers was more tolerant to the obligate biotrophic fungus Golovinomyces cichoracearum but did not display a PCD-like phenotype, suggesting that Cers alone are not involved in the induction of PCD (König et al., 2012). Moreover, Saucedo-García et al. (2011) postulated that dihydroxy-LCBs, but not trihydroxy-LCBs, might be primary mediators for LCB-induced PCD. The sphingoid base hydroxylase sbh1/sbh2 double mutant completely lacking trihydroxy-LCBs showed enhanced expression of PCD marker genes (Chen et al., 2008). On the contrary, increase in t18:0 was specifically sustained in plant interaction with the avirulent Pseudomonas syringae pv tomato (Pst) strain and correlated with a strong PCD induction in leaves (Peer et al., 2010). Thus, the nature of sphingolipids able to induce PCD is still under debate and may evolve depending on plants and their environment. The phosphorylated form of LCBs (LCB-Ps) could abrogate PCD induced by LCBs, Cers, or heat stress in a dose-dependent manner (Shi et al., 2007; Alden et al., 2011). Furthermore, blocking the conversion of LCBs to LCB-Ps by using specific inhibitors induced PCD in cell suspension culture (Alden et al., 2011). Recently, overexpression of rice (Oryza sativa) LCB kinase in transgenic tobacco (Nicotiana tabacum) plants reduced PCD after treatment with FB1 (Zhang et al., 2013). Genetic mutation on LCB-P lyase encoded by the AtDPL1 gene, modifying the LCB-LCB-P ratio, could impact PCD levels after treatment with FB1 (Tsegaye et al., 2007). Altogether, these data point to the existence of a rheostat between LCBs and their phosphorylated forms that controls plant cell fate toward cell death or survival.Data on plant sphingolipid functions are still fragmentary. Only a few reports have described interconnections between sphingolipids, cell death, and plant defense responses, almost exclusively in response to (hemi)biotrophic pathogens. Knowledge about such relations in response to necrotrophic pathogens is still in its infancy (Rivas-San Vicente et al., 2013; Bi et al., 2014). In this report, the link between sphingolipids, cell death, and plant defense has been explored in response to Botrytis cinerea infection and in comparison with Pst infection. For this purpose, Atdpl1 mutant plants, disturbed in LCB/LCB-P accumulation without displaying any phenotype under standard growth conditions (Tsegaye et al., 2007), have been analyzed after pathogen infection. Our results revealed that modification of sphingolipid contents not only impacted plant tolerance to hemibiotrophs but also greatly affected resistance to necrotrophs. Whereas the SA signaling pathway is globally repressed in Atdpl1-1 compared with wild-type plants, the JA signaling pathway is significantly enhanced. Cell death and ROS accumulation are markedly modified in Atdpl1-1 mutant plants. We further demonstrated that phytosphingosine-1-phosphate (t18:0-P) and d18:0 are key players in pathogen-induced cell death and ROS generation. Here, we thus established a link between JA signaling, PCD, and sphingolipid metabolism.  相似文献   

11.
High fluence-rate blue light (BL) rapidly inhibits hypocotyl growth in Arabidopsis, as in other species, after a lag time of 30 s. This growth inhibition is always preceded by the activation of anion channels. The membrane depolarization that results from the activation of anion channels by BL was only 30% of the wild-type magnitude in hy4, a mutant lacking the HY4 BL receptor. High-resolution measurements of growth made with a computer-linked displacement transducer or digitized images revealed that BL caused a rapid inhibition of growth in wild-type and hy4 seedlings. This inhibition persisted in wild-type seedlings during more than 40 h of continuous BL. By contrast, hy4 escaped from the initial inhibition after approximately 1 h of BL and grew faster than wild type for approximately 30 h. Wild-type seedlings treated with 5-nitro-2-(3-phenylpropylamino)-benzoic acid, a potent blocker of the BL-activated anion channel, displayed rapid growth inhibition, but, similar to hy4, these seedlings escaped from inhibition after approximately 1 h of BL and phenocopied the mutant for at least 2.5 h. The effects of 5-nitro-2-(3-phenylpropylamino)-benzoic acid and the HY4 mutation were not additive. Taken together, the results indicate that BL acts through HY4 to activate anion channels at the plasma membrane, causing growth inhibition that begins after approximately 1 h. Neither HY4 nor anion channels appear to participate greatly in the initial phase of inhibition.  相似文献   

12.
Application of crab shell chitin or pentamer chitin oligosaccharide to Arabidopsis seedlings increased tolerance to salinity in wild-type but not in knockout mutants of the LysM Receptor-Like Kinase1 (CERK1/LysM RLK1) gene, known to play a critical role in signaling defense responses induced by exogenous chitin. Arabidopsis plants overexpressing the endochitinase chit36 and hexoaminidase excy1 genes from the fungus Trichoderma asperelleoides T203 showed increased tolerance to salinity, heavy-metal stresses, and Botrytis cinerea infection. Resistant lines, overexpressing fungal chitinases at different levels, were outcrossed to lysm rlk1 mutants. Independent homozygous hybrids lost resistance to biotic and abiotic stresses, despite enhanced chitinase activity. Expression analysis of 270 stress-related genes, including those induced by reactive oxygen species (ROS) and chitin, revealed constant up-regulation (at least twofold) of 10 genes in the chitinase-overexpressing line and an additional 76 salt-induced genes whose expression was not elevated in the lysm rlk1 knockout mutant or the hybrids harboring the mutation. These findings elucidate that chitin-induced signaling mediated by LysM RLK1 receptor is not limited to biotic stress response but also encompasses abiotic-stress signaling and can be conveyed by ectopic expression of chitinases in plants.  相似文献   

13.
拟南芥幼苗用1-萘氨甲酰苯甲酸(NPA)、IAA和GA3处理后测定根的伸长和向重力性弯曲的结果表明,低浓度(0.001μmol·L-1)IAA和(0.01~1μmol·L-1)GA3促进根的伸长和向重力性弯曲,高浓度(0.01~10μmol·L-1)IAA和(10~100μmol·L-1)GA3的则相反。NPA在总体上是抑制根的伸长和向重力性弯曲,但低浓度(0.4μmol·L-1)的NPA有促进根伸长的趋势。低浓度的IAA和GA3均拮抗NPA对根伸长的影响,且低浓度的GA3对根伸长的促进作用并不依赖IAA。  相似文献   

14.
Abstract: Amyloid β-peptide (Aβ) is the principal component of neuritic plaques in the brain in Alzheimer's disease (AD). Recent studies revealed that Aβ can be neurotoxic by a mechanism involving free radical production and loss of cellular ion homeostasis, thus implicating Aβ as a key factor in the pathogenesis of AD. However, other proteins are present in plaques in AD, including the protease thrombin and protease nexin-1 (PN1), a thrombin inhibitor. We therefore tested the hypothesis that thrombin and PN1 modify neuronal vulnerability to Aβ toxicity. In dissociated rat hippocampal cell cultures the toxicity of Aβ was significantly enhanced by coincubation with thrombin, whereas PN1 protected neurons against Aβ toxicity. Aβ induced an increase in levels of intracellular peroxides and calcium. Thrombin enhanced, and PN1 attenuated, the accumulation of peroxides and calcium induced by Aβ. Taken together, these data demonstrate that thrombin and PN1 have opposing effects on neuronal vulnerability to Aβ and suggest that thrombin and PN1 play roles in the pathogenesis of neuronal injury in AD.  相似文献   

15.
Cell growth is a highly regulated, plastic process. Its control involves balancing positive regulation of anabolic processes with negative regulation of catabolic processes. Although target of rapamycin (TOR) is a major promoter of growth in response to nutrients and growth factors, AMP-activated protein kinase (AMPK) suppresses anabolic processes in response to energy stress. Both TOR and AMPK are conserved throughout eukaryotic evolution. Here, we review the fundamentally important roles of these two kinases in the regulation of cell growth with particular emphasis on their mutually antagonistic signaling.An efficient homeostatic response to maintain cellular energy despite a noncontinuous supply of nutrients is crucial for the survival of organisms. Cells have, therefore, evolved a host of molecular pathways to sense both intra- and extracellular nutrients and thereby quickly adapt their metabolism to changing conditions. The target of rapamycin (TOR) and AMP-activated protein kinase (AMPK) signaling pathways control growth and metabolism in a complementary manner with TOR promoting anabolic processes under nutrient- and energy-rich conditions, whereas AMPK promotes a catabolic response when cells are low on nutrients and energy. Both pathways are highly conserved from yeast to human. This review summarizes the cross talk between TOR and AMPK in different organisms.  相似文献   

16.
17.
植物细胞命运决定机制的解析一直以来都是植物发育生物学研究的核心.模式植物拟南芥的表皮毛形成过程是研究植物细胞命运决定的优良模式系统.为了筛选和鉴定控制拟南芥表皮毛形成的新因子,我们进行了大规模的正向遗传筛选,获得了两株莲座叶表皮毛不能形成或数量显著减少的突变体f08-01和vat002-07.通过对突变基因的克隆和遗传...  相似文献   

18.
Kaposi’s sarcoma-associated herpesvirus (KSHV), also known as human herpesvirus-8, is the causative agent of three hyperproliferative disorders: Kaposi’s sarcoma, primary effusion lymphoma (PEL) and multicentric Castleman’s disease. During viral latency a small subset of viral genes are produced, including KSHV latency-associated nuclear antigen (LANA), which help the virus thwart cellular defense responses. We found that exposure of KSHV-infected cells to oxidative stress, or other inducers of apoptosis and caspase activation, led to processing of LANA and that this processing could be inhibited with the pan-caspase inhibitor Z-VAD-FMK. Using sequence, peptide, and mutational analysis, two caspase cleavage sites within LANA were identified: a site for caspase-3 type caspases at the N-terminus and a site for caspase-1 and-3 type caspases at the C-terminus. Using LANA expression plasmids, we demonstrated that mutation of these cleavage sites prevents caspase-1 and caspase-3 processing of LANA. This indicates that these are the principal sites that are susceptible to caspase cleavage. Using peptides spanning the identified LANA cleavage sites, we show that caspase activity can be inhibited in vitro and that a cell-permeable peptide spanning the C-terminal cleavage site could inhibit cleavage of poly (ADP-ribose) polymerase and increase viability in cells undergoing etoposide-induced apoptosis. The C-terminal peptide of LANA also inhibited interleukin-1beta (IL-1β) production from lipopolysaccharide-treated THP-1 cells by more than 50%. Furthermore, mutation of the two cleavage sites in LANA led to a significant increase in IL-1β production in transfected THP-1 cells; this provides evidence that these sites function to blunt the inflammasome, which is known to be activated in latently infected PEL cells. These results suggest that specific caspase cleavage sites in KSHV LANA function to blunt apoptosis as well as interfere with the caspase-1-mediated inflammasome, thus thwarting key cellular defense mechanisms.  相似文献   

19.
20.
SDD1是气孔发育过程中的关键调控基因,编码一个类枯草杆菌(Bacillus subtilis)蛋白酶的丝氨酸蛋白酶。从EMS诱变的拟南芥(Arabidopsis thaliana)中筛选到2株类似sdd1-1的气孔密度突变体,即e281和g204。其气孔密度和指数均比野生型增加约1.5倍,气孔成簇。遗传分析和基因测序证实它们是2个不同的SDD1新等位基因,其突变分别导致了底物结合位点N区域和催化三联体之一--S区域的氨基酸变化,分别为S变成T及S变为F。形态学和生理学研究表明,SDD1基因不同位点发生突变可导致不同的生物学效应;而且SDD1等位基因间存在拮抗作用,其可能属于基因转应作用中的负效应。  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